id
stringlengths
27
33
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
13
1.81M
warning/0004/cond-mat0004153.html
ar5iv
text
# Untitled Document Separable Structure of Many-Body Ground-State Wave Function Yeong E. Kim <sup>*</sup><sup>*</sup>* e-mail:yekim$`\mathrm{@}`$physics.purdue.edu and Alexander L. Zubarev e-mail: zubareva$`\mathrm{@}`$physics.purdue.edu Department of Physics, Purdue University West Lafayette, Indiana 47907 Abstract > We have investigated a general structure of the ground-state wave function for the Schrödinger equation for $`N`$ identical interacting particles (bosons or fermions) confined in a harmonic anisotropic trap in the limit of large $`N`$. It is shown that the ground-state wave function can be written in a separable form. As an example of its applications, this form is used to obtain the ground-state wave function describing collective dynamics for $`N`$ trapped bosons interacting via contact forces. The structure of the ground-state wave function for a many-body system is very important for theoretical understanding of recently observed Bose-Einstein condensation (BEC) (the theoretical aspects of the BEC are discussed in recent reviews ) and other many body problems. The Ginzburg-Pitaevskii-Gross (GPG) equation is most widely used to describe the experimental results for the BEC. Recently, an alternative method of equivalent linear two-body (ELTB) equations for many body systems has been developed based on the variational principle . In this paper, we consider $`N`$ identical particles (bosons or fermions) confined in a harmonic anisotropic trap. We show that in the case of large $`N`$ the ground-state wave function can be written in separable form as $$\mathrm{\Psi }(\stackrel{}{r}_1,\stackrel{}{r}_2,\mathrm{}\stackrel{}{r}_N)=\varphi (x,y,z)\chi (\mathrm{\Omega },\sigma ),$$ $`(1)`$ where $$x=\sqrt{\underset{i=1}{\overset{N}{}}x_i^2},y=\sqrt{\underset{i=1}{\overset{N}{}}y_i^2},z=\sqrt{\underset{i=1}{\overset{N}{}}z_i^2},$$ $`(2)`$ $`\mathrm{\Omega }`$ is a set of (3N - 3) angular variables, and $`\sigma `$ is a set of spin variables. We start from a generalization of the hyperspherical expansion of the wave function for the Hamiltonian $$H=\frac{\mathrm{}^2}{2m}\underset{i=1}{\overset{N}{}}\mathrm{\Delta }_i+\frac{1}{2}m\underset{i=1}{\overset{N}{}}(\omega _x^2x_i^2+\omega _y^2y_i^2+\omega _z^2z_i^2)+\underset{i<j}{}V_{int}(𝐫_i𝐫_j)$$ $`(3)`$ in the form $$\mathrm{\Psi }(𝐫_1,\mathrm{}𝐫_N)=\underset{[K]}{}\mathrm{\Phi }_{[K]}(x,y,z)Y_{[K]}(\mathrm{\Omega }_x^N,\mathrm{\Omega }_y^N,\mathrm{\Omega }_z^N,\sigma ),$$ $`(4)`$ where $`Y_{[K]}(\mathrm{\Omega }_x^N,\mathrm{\Omega }_y^N,\mathrm{\Omega }_z^N,\sigma )=Y_{K_x,K_y,K_z}^{\nu _x,\nu _y,\nu _z}(\mathrm{\Omega }_x^N,\mathrm{\Omega }_y^N,\mathrm{\Omega }_z^N,\sigma )`$ is the combination of the hyperspherical harmonics, $`Y_{K_x}^{\nu _x}(\mathrm{\Omega }_x^N),Y_{K_y}^{\nu _y}(\mathrm{\Omega }_y^N),`$ and $`Y_{K_z}^{\nu _z}(\mathrm{\Omega }_z^N)`$, with functions of spin variables $`\sigma `$, which is symmetric or antisymmetric with respect to permutations of particles for bosons or fermions respectively. $`[K]`$ represents a set of numbers $`[K_x,\nu _x,K_y,\nu _y,K_z,\nu _z]`$. The hyperspherical harmonics $`Y_{K_x}^{\nu _x}(\mathrm{\Omega }_x^N)`$, $`Y_{K_y}^{\nu _y}(\mathrm{\Omega }_y^N)`$, and $`Y_{K_z}^{\nu _z}(\mathrm{\Omega }_z^N)`$ are eigenfunctions of the hyperspherical angular parts of the Laplace operators $`_{i=1}^N\frac{^2}{x_i^2}`$, $`_{i=1}^N\frac{^2}{y_i^2}`$, and $`_{i=1}^N\frac{^2}{z_i^2}`$, respectively. The Laplace operators are defined by $$\underset{i=1}{\overset{N}{}}\frac{^2}{x_i^2}=\frac{1}{x^{N1}}\frac{}{x}(x^{N1}\frac{}{x})+\frac{1}{x^2}\mathrm{\Delta }_{\mathrm{\Omega }_x^N},$$ $$\underset{i=1}{\overset{N}{}}\frac{^2}{y_i^2}=\frac{1}{y^{N1}}\frac{}{y}(y^{N1}\frac{}{y})+\frac{1}{y^2}\mathrm{\Delta }_{\mathrm{\Omega }_y^N},$$ $`(5)`$ and $$\underset{i=1}{\overset{N}{}}\frac{^2}{z_i^2}=\frac{1}{z^{N1}}\frac{}{z}(z^{N1}\frac{}{z})+\frac{1}{z^2}\mathrm{\Delta }_{\mathrm{\Omega }_z^N}.$$ The hyperspherical angles $`\theta _1^x,\theta _2^x,\mathrm{}\theta _{N1}^x,\theta _1^y,\theta _2^y,\mathrm{}\theta _{N1}^y,\theta _1^z,\theta _2^z,\mathrm{}\theta _{N1}^z`$ can be chosen in such a way that the hyperspherical angular parts of the Laplace operators $`\mathrm{\Delta }_{\mathrm{\Omega }_t^N}`$ satisfy the recursion relation $$\mathrm{\Delta }_{\mathrm{\Omega }_u^N}=\frac{1}{\mathrm{sin}^{N2}\theta _{N1}^u}\frac{}{\theta _N^u}(\mathrm{sin}^{N2}\theta _{N1}^u\frac{}{\theta _{N1}^u})+\frac{1}{\mathrm{sin}^2\theta _{N1}^u}\mathrm{\Delta }_{\mathrm{\Omega }_u^{N1}}$$ $`(6)`$ with $`u=x,y,`$ or $`z`$. Functions $`\mathrm{\Phi }_{[K]}(x,y,z)`$ satisfy equations $$\underset{[K^{}]}{}h_{[K],[K^{}]}\mathrm{\Phi }_{[K^{}]}(x,y,z)=E\mathrm{\Phi }_{[K]}(x,y,z),$$ $`(7)`$ where $$\begin{array}{ccc}& & h_{[K][K^{}]}=\delta _{K_xK_x^{}}\delta _{K_yK_y^{}}\delta _{K_zK_z^{}}\delta _{\nu _x\nu _x^{}}\delta _{\nu _y\nu _y^{}}\delta _{\nu _z\nu _z^{}}[\frac{\mathrm{}^2}{2m}(\frac{^2}{x^2}+\frac{^2}{y^2}+\frac{^2}{z^2})\hfill \\ & & \\ & & +\frac{m}{2}(\omega _x^2x^2+\omega _y^2y^2+\omega _z^2z^2)+\frac{\mathrm{}^2}{2m}(\frac{(N1+2K_x)(N3+2K_x)}{4x^2}\hfill \\ & & \\ & & +\frac{(N1+2K_y)(N3+2K_y)}{4y^2}+\frac{(N1+2K_z)(N3+2K_z)}{4z^2}]\hfill \\ & & \\ & & +V_{[K][K^{}]}(x,y,z),\hfill \end{array}$$ $`(8)`$ with $$V_{[K][K^{}]}(x,y,z)=<K_x,\nu _x,K_y,\nu _y,K_z,\nu _z\underset{i<j}{}V_{int}(𝐫_i𝐫_j)K_x^{},\nu _x^{},K_y^{},\nu _y^{},K_z^{},\nu _z^{}>.$$ $`(9)`$ We write $`\mathrm{\Phi }_{[K]}(x,y,z)`$ in the form of a Laplace integral $$\mathrm{\Phi }_{[K]}(x,y,z)=f_{[K]}(\alpha _x,\alpha _y,\alpha _z)\varphi _x(x,\alpha _x)\varphi _y(y,\alpha _y)\varphi _z(z,\alpha _z)𝑑\alpha _x𝑑\alpha _y𝑑\alpha _z,$$ $`(10)`$ where $$\varphi _t(t,\alpha _t)=\sqrt{\frac{2}{\mathrm{\Gamma }(N/2)}}(\frac{m\stackrel{~}{\omega }}{\alpha _t^2\mathrm{}})^{N/4}\mathrm{exp}[m\stackrel{~}{\omega }(\frac{t}{\alpha _t})^2/(2\mathrm{})]t^{(N1)/2},$$ $`(11)`$ and $`\stackrel{~}{\omega }=(\omega _x\omega _y\omega _z)^{1/3}`$. The Hill-Wheeler type equations are obtained by requiring that energy of the system is stationary with respect to the functions $`f_{[K]}(\alpha _x,\alpha _y,\alpha _z)`$ $$\begin{array}{ccc}& & _{[K]}d\alpha _xd\alpha _yd\alpha _zf_{[K]}(\alpha _x,\alpha _y,\alpha _z)[H_{[K][K^{}]}(\alpha _x\alpha _y\alpha _z,\alpha _x^{}\alpha _y^{}\alpha _z^{})\hfill \\ & & \\ & & \delta _{[K][K^{}]}S(\alpha _x\alpha _y\alpha _z,\alpha _x^{}\alpha _y^{}\alpha _z^{})E]=0,\hfill \end{array}$$ $`(12)`$ where $$\begin{array}{ccc}& & H_{[K][K^{}]}(\alpha _x\alpha _y\alpha _z,\alpha _x^{}\alpha _y^{}\alpha _z^{})=<\varphi _x(x,\alpha _x)\varphi _y(y,\alpha _y)\varphi _z(z,\alpha _z)Y_{[K]}\hfill \\ & & \\ & & \times H\varphi _x(x,\alpha _x^{})\varphi _y(y,\alpha _y^{})\varphi _z(z,\alpha _z^{})Y_{[K^{}]}>,\hfill \end{array}$$ $`(13)`$ and $$S(\alpha _x\alpha _y\alpha _z,\alpha _x^{}\alpha _y^{}\alpha _z^{})=<\varphi _x(x,\alpha _x)\varphi _y(y,\alpha _y)\varphi _z(z,\alpha _z)\varphi _x(x,\alpha _x^{})\varphi _y(y,\alpha _y^{})\varphi _z(z,\alpha _z^{})>.$$ $`(14)`$ In order to solve the Hill-Wheeler type equations (12), we assume that the integral in Eq. (10) can be replaced by sum $$\mathrm{\Phi }_{[K]}(x,y,z)=\underset{i,j,k=1}{\overset{\mathrm{}}{}}c_{ijk}^{[K]}\varphi _x(x,\alpha _x^i)\varphi _y(y,\alpha _y^j)\varphi _z(z,\alpha _z^k),$$ $`(15)`$ where $`c_{ijk}^{[K]}`$ are solutions of the following equations $$\underset{\stackrel{i^{},j^{},k^{}}{[K^{}]}}{}[H_{[K][K^{}]}(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha _y^j^{}\alpha _z^k^{})\delta _{[K][K^{}]}S(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha _y^j^{}\alpha _z^k^{})E]c_{i^{}j^{}k^{}}^{[K^{}]}=0.$$ $`(16)`$ For the case of large $`N`$, the overlap, Eq. (14), $$S(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha _y^j^{}\alpha _z^k^{})=[\frac{8\alpha _x^i\alpha _x^i^{}\alpha _y^j\alpha _y^j^{}\alpha _z^k\alpha _z^k^{}}{((\alpha _x^i)^2+(\alpha _x^i^{})^2)((\alpha _y^j)^2+(\alpha _y^j^{})^2)((\alpha _z^k)^2+(\alpha _z^k^{})^2)}]^{N/2}$$ $`(17)`$ reduces to the Kronecker deltas $$S(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha _y^j^{}\alpha _z^k^{})=\delta _{ii^{}}\delta _{jj^{}}\delta _{kk^{}}$$ $`(18)`$ Since the ratio $$H_{[K][K^{}]}(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha _y^j^{}\alpha _z^k^{})/S(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha _y^j^{}\alpha _z^k^{})$$ is a much more slowly varying function of $`\alpha `$ compared to $`S(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha ^j^{}y\alpha _z^k^{})`$ in almost all cases , we have for the case of large $`N`$ $$H_{[K][K^{}]}(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha _y^j^{}\alpha _z^k^{})=\stackrel{~}{H}_{[K][K^{}]}(\stackrel{~}{\alpha _x},\stackrel{~}{\alpha _y},\stackrel{~}{\alpha _z})\delta _{ii^{}}\delta _{jj^{}}\delta _{kk^{}},$$ $`(19)`$ (see Appendix for the case of $`N`$ identical particles interacting via contact repulsive force). Substitution of Eq. (19) into Eq. (16) gives $$\mathrm{\Phi }_{[K]}(x,y,z)=\stackrel{~}{c}_{[K]}\varphi _x(x,\stackrel{~}{\alpha _x})\varphi _y(y,\stackrel{~}{\alpha _y})\varphi _z(z,\stackrel{~}{\alpha _z}),$$ $`(20)`$ where $`\stackrel{~}{c}_{[K]}`$ are solutions of the following equations $$\underset{[K^{}]}{}[\stackrel{~}{H}_{[K][K^{}]}(\stackrel{~}{\alpha _x},\stackrel{~}{\alpha _y},\stackrel{~}{\alpha _z})\delta _{[K][K^{}]}E]\stackrel{~}{c}_{[K^{}]},$$ $`(21)`$ and parameters $`\stackrel{~}{\alpha _x},\stackrel{~}{\alpha _y}`$, and $`\stackrel{~}{\alpha _z}`$ are solutions of $$\frac{E}{\stackrel{~}{\alpha _x}}=\frac{E}{\stackrel{~}{\alpha _y}}=\frac{E}{\stackrel{~}{\alpha _z}}=0.$$ Substitution of Eq. (20) into Eq. (4) yields $`\mathrm{\Psi }(\stackrel{}{r}_1,\stackrel{}{r}_2,\mathrm{}\stackrel{}{r}_N)`$ given by Eq. (1) with $$\varphi (x,y,z)=\varphi _x(x,\stackrel{~}{\alpha _x})\varphi _y(y,\stackrel{~}{\alpha _y})\varphi _z(z,\stackrel{~}{\alpha _z}),$$ and $$\chi (\mathrm{\Omega },\sigma )=\underset{[K]}{}\stackrel{~}{c}^{[K]}Y_{[K]}(\mathrm{\Omega }_x^N,\mathrm{\Omega }_y^N,\mathrm{\Omega }_z^N,\sigma ).$$ We now consider $`N`$ identical particles confined in an anisotropic harmonic trap and interacting via contact force $$V_{int}(\stackrel{}{r}_i\stackrel{}{r}_j)=\frac{4\pi \mathrm{}^2a}{m}\delta (\stackrel{}{r}_i\stackrel{}{r}_j),$$ $`(23)`$ with positive scattering length $`a>0`$. Using factorization (1) we have $$\begin{array}{ccc}& & [\frac{\mathrm{}^2}{2m}(\frac{^2}{x^2}+\frac{^2}{y^2}+\frac{^2}{z^2})+\frac{m}{2}(\omega _x^2x^2+\omega _y^2y^2+\omega _z^2z^2)\frac{\mathrm{}^2}{2m}(\frac{c_x}{x^2}+\frac{c_y}{y^2}+\frac{c_z}{z^2})\hfill \\ & & \\ & & +\frac{\mathrm{}^2}{2m}\frac{(N1)(N3)}{4}(\frac{1}{x^2}+\frac{1}{y^2}+\frac{1}{z^2})+\frac{c}{xyz}]\varphi (x,y,z)=E\varphi (x,y,z),\hfill \end{array}$$ $`(24)`$ where $`c_t=<\chi \mathrm{\Delta }_{\mathrm{\Omega }_t^N}\chi >/<\chi \chi >`$ with $`t=(x,y,z)`$ and $$c=\frac{a\mathrm{}^2N(N1)}{\sqrt{2\pi }m}(\frac{\mathrm{\Gamma }(N/2)}{\mathrm{\Gamma }((N1)/2)})^3\stackrel{~}{c}.$$ In the large $`N`$ limit, parameters $`c_x,c_y,c_z,`$ and $`\stackrel{~}{c}`$ are expected to be slowly varying functions of $`N`$. For $`N`$ identical bosonic atoms with large $`N,`$ an essentially exact expression for the ground state energy can be obtained by neglecting the kinetic energy term in the GPG equation (this is called “Thomas-Fermi approximation” ). From comparison of the ground state solution of Eq. (24) with the Thomas-Fermi approximation , we can fix unknown parameters and find the ground-state solution of Eq. (24) as $$\varphi (x,y,z)=\psi _x(x)\psi _y(y)\psi _z(z),$$ $`(25)`$ $$E=\frac{5N\mathrm{}\stackrel{~}{\omega }}{4}\stackrel{~}{n}^{2/5},$$ $`(26)`$ with $$\begin{array}{ccc}\hfill \psi _x(x)=Ax^{(N1)/2}\mathrm{exp}[m\stackrel{~}{\omega }(x/\alpha )^2/(2\mathrm{})],& & \\ \hfill \psi _y(y)=Ay^{(N1)/2}\mathrm{exp}[m\stackrel{~}{\omega }(y/\beta )^2/(2\mathrm{})],& & \\ \hfill \psi _z(z)=Az^{(N1)/2}\mathrm{exp}[m\stackrel{~}{\omega }(z/\gamma )^2/(2\mathrm{})],& & \end{array}$$ $`(27)`$ where $`A=\sqrt{2/\mathrm{\Gamma }(N/2)}(m\stackrel{~}{\omega }/(\alpha ^2\mathrm{}))^{N/4}`$, $`\alpha =\stackrel{~}{n}^{1/5}\stackrel{~}{\omega }/\omega _x`$, $`\beta =\stackrel{~}{n}^{1/5}\stackrel{~}{\omega }/\omega _y`$, $`\gamma =\stackrel{~}{n}^{1/5}\stackrel{~}{\omega }/\omega _z`$, $`\stackrel{~}{\omega }=(\omega _x\omega _y\omega _z)^{1/3},\stackrel{~}{n}=n\stackrel{~}{c},n=2\sqrt{\stackrel{~}{\omega }m/(2\pi \mathrm{})}Na`$ and $$\stackrel{~}{c}=(\frac{4}{7})^{5/2}\frac{15}{8}\sqrt{\pi }0.82.$$ $`(28)`$ Eqs. (25-28) give the exact ground-state solution of Eq. (24) for large $`N`$. Thus we have found an analytical solution for the ground-state wave function describing collective dynamics in variables (x,y,z) in the large $`N`$ limit. We note that the slope of the Thomas-Fermi wave function becomes infinity at the surface, leading to logarithmic singularity in the kinetic energy. Hence it is necessary to modify the Thomas-Fermi wave function near the surface \[12-14\]. In contrast, we do not have such problems for our solution, Eq. (25-28). It is also interesting to compare our results with the ELTB method . For this situation (contact force, Eq. (23) and large N limit), the ELTB method corresponds to $`\stackrel{~}{c}^{2/5}=1`$. It shows that the ELTB method is a very good approximation with relative error of about $`8\%`$ for parameter $`\stackrel{~}{c}^{2/5}`$. In summary, we have investigated the general structure of the ground-state solution of the Schrödinger equation for $`N`$ identical interacting particles (bosons or fermions) confined in a harmonic anisotropic trap in the large $`N`$ limit. The main results and conclusions are as follows (i) It has been shown that in the case of large $`N`$ the ground-state wave function can be written in separable form, Eq. (1). (ii) Using this form, we have found an analytical solution for the ground-state wave function, Eqs. (25-29), describing collective dynamics in collective variables (x,y,z) for $`N`$ trapped bosons interacting via contact repulsive forces. (iii) Our results can be used for checking various approximations (both existing and future) made for the Schrödinger equation describing $`N`$ identical interacting particles (bosons or fermions) confined in a harmonic anisotropic trap. Appendix To prove Eq. (19) we consider the contact potential case $$V_{int}(𝐫_i𝐫_j,\sigma )=\delta (𝐫_i𝐫_j)\eta (\sigma ),$$ $`(A.1)`$ where $`\eta `$ depends on spin variables. Using Eq. (A.1) we can rewrite Eq. (9) as $$V_{[K][K^{}]}(x,y,z)=\gamma _{[K][K^{}]}\frac{N(N1)}{xyz},$$ $`(A.2)`$ where $`\gamma _{[K][K^{}]}`$ does not depend on $`x,y,z`$. Substitution of Eq. (A.2) into Eq. (13) gives $$\begin{array}{ccc}& & H_{[K][K^{}]}(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha _y^j^{}\alpha _z^k^{})/(\mathrm{}\stackrel{~}{\omega }N)=(1/2)S(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha _y^j^{}\alpha _z^k^{})\hfill \\ & & \\ & & \times [\delta _{[K][K^{}]}(\frac{1+(\alpha _x^i)^2(\alpha _x^i^{})^2\beta _x^2}{(\alpha _x^i)^2+(\alpha _x^i^{})^2}+\frac{1+(\alpha _y^j)^2(\alpha _y^j^{})^2\beta _y^2}{(\alpha _y^j)^2+(\alpha _y^j^{})^2}+\frac{1+(\alpha _z^k)^2(\alpha _z^k^{})^2\beta _z^2}{(\alpha _z^k)^2+(\alpha _z^k^{})^2})\hfill \\ & & \\ & & +\frac{\sqrt{((\alpha _x^i)^2+(\alpha _x^i^{})^2)((\alpha _y^j)^2+(\alpha _y^j^{})^2)((\alpha _z^k)^2+(\alpha _z^k^{})^2)}}{\alpha _x^i\alpha _x^i^{}\alpha _y^j\alpha _y^j^{}\alpha _z^k\alpha _z^k^{}}(\frac{\mathrm{\Gamma }((N1)/2)}{\mathrm{\Gamma }(N/2)})^3\hfill \\ & & \\ & & \times (\frac{m\stackrel{~}{\omega }}{\mathrm{}})^{3/2}\frac{N1}{\sqrt{8}}\gamma _{[K][K^{}]}],\hfill \end{array}$$ $`(A.3)`$ where $`\beta _t=\omega _t/\stackrel{~}{\omega }`$ for $`t=x,y`$, or $`z`$. For large $`N`$, $`S(\alpha _x^i\alpha _y^j\alpha _z^k,\alpha _x^i^{}\alpha _y^j^{}\alpha _z^k^{})`$, Eq. (14), reduces to the Kronecker deltas $`\delta _{ii^{}}\delta _{jj^{}}\delta _{kk^{}}`$, and hence from Eq. (A.3) we obtain Eq. (19). References 1. http://amo.phy.gasou.edu/bec.html, and references therein. 2. K. Burnett, M. Edwards, and C.W. Clark, Phys. Today, 52, 37 (1999); F. Dalfovo, S. Giorgini, L. Pitaevskii, and S. Stringari, Rev. Mod. Phys. 71, 463 (1999). 3. L. Ginzburg and L.P. Pitaevskii, Zh. Eksp. Teor. Fiz. 34, 1240 (1958) \[Sov. Phys. JETP, 858 (1958)\]; E.P. Gross, J. Math. Phys. 4, 195 (1963). 4. A.L. Zubarev and Y.E. Kim, Phys. Lett. A263, 33 (1999). 5. Y.E. Kim and A.L. Zubarev, J. Phys. B: At. Mol. Opt. Phys. 33, 55 (2000). 6. Yu. T. Grin’, Yad. Fiz. 12, 927 (1970) \[Sov. J. Nucl. Phys. 12, 345 (1971)\]. 7. Yu. F. Smirnov and K. V. Shitikova, Fiz. Elem. Chastits At. Yadra 8, 847 (1977) \[Sov. J. Part. Nucl. 8, 344 (1977)\]. 8. D.L. Hill and J.A. Wheeler, Phys. Rev. 89, 1102 (1953). 9. J.J. Griffin and J.A. Wheeler, Phys. Rev. 108, 311 (1957). 10. D.M. Brink and A. Weiguny, Nucl. Phys. A120, 59 (1968). 11. G. Baym and C. J. Pethick, Phys. Rev. Lett. 76, 6 (1996). 12. A.L. Fetter and D.L. Feder, Phys. Rev. A58, 3185 (1998). 13. F. Dalfovo, L.P. Pitaevskii, and S. Stringari, Phys. Rev. A54, 4213 (1996). 14. E. Lundh, C.J. Pethick, and H. Smith, Phys. Rev. A55, 2126 (1997).
warning/0004/quant-ph0004095.html
ar5iv
text
# Teleportation Scheme of 𝑆-level Quantum Pure States by Two-level EPRsSupported by National Natural Science Foundation of China under Grant No. 19975043 ((April 25, 2000)) ## Abstract Unknown quantum pure states of arbitrary but definite $`s`$-level of a particle can be transferred onto a group of remote two-level particles through two-level EPRs as many as the number of those particles in this group. We construct such a kind of teleportation, the realization of which need a nonlocal unitary transformation to the quantum system that is made up of the s-level particle and all the two-level particles at one end of the EPRs, and measurements to all the single particles in this system. The unitary transformation to more than two particles is also written into the product form of two-body unitary transformations. Quantum mechanics offers us the capabilities of transferring information different from the classical case, either for computation or communication. Bennett et.al. , developed a quantum method of teleportation, through which, an unknown quantum pure state of a spin-$`\frac{1}{2}`$ particle (we call it ’qubit’ ) is teleported from the sender ’Alice’ at the sending terminal onto the qubit at the receiving terminal where the receiver ’Bob’ need to perform a unitary transformation on his qubit. At first it is necessary to prepare two spin-$`\frac{1}{2}`$ particles in an Einstein-Podolsky-Rosen (EPR) entangled state or so-called a Bell state and send them to the two different places to establish a quantum channel between Alice and Bob. The second step is that Alice performs a Bell operator measurement to the quantum system involving her share of the two entangled particles together with the particle at an unknown state to be transferred. Then through classical channels, for example, by broadcasting, Alice needs to let Bob know which one she gets of the four possible outcomes of the Bell operator measurement. After Bob performs on his share of the two formerly entangled particles one of four unitary transformations determined by those outcomes, this particle will be in the unknown state. In this way, the unknown state is teleported from one place to another. The new method of teleportation has interested a lot of research groups. They at once started the research work on quantum teleportation and have made great development, theoretical and experimental as well. It was generalized to the case of continuous variables . Sixia Yu et.al., investigated canonical quantum teleportation of finite-level unknown states by introducing a canonically conjugated pair of quantum phase and number . The successful experimental realization of quantum telepartation of unknown polarization states carried on a photon and the succedent experiments about finite-level quantum system teleportation have aroused a series of discussions and further research of this topic from various aspects . Possible applications have been considered in Ref. . The method of teleportation in the case of continuous variables got its experimental realization in 1998 . From a general point of view, no matter what form it is, there are four steps to realize the quantum teleportation, which can be seen clearly in Bennett’s initial scheme : (a) EPR entangled states preparing; (b) Bell operator measurements by the sender; (c) the sender informing the receiver of his outcomes through classical channels; (d) the receiver performing unitary transformation according to the classical information. However, the (b) step is not necessary, for it can be substituted by a nonlocal unitary transformation along with local measurements (Here ’local’ means to single particles). More specifically, the unitary transformation is performed on the sender’s EPR particle and the state-unknown particle to form some sort of entangled state involving the state-unknown particle together with the EPR particles, both Alice’s and Bob’s, while the local measurements are performed one by one on Alice’s particles. These measurements will result in the random collapse of all the sender’s particles onto definite states. At the other end of communication, the receiver will got the same results as in the case of performing Bell operator measurements. In other words, the unitary transformation and local measurements is equivalent to a Bell operator measurement. The unnecessity of the (b) step gets further evidence from Ref. in which Brassard et. al. indicated the possibility of realizing teleportation by controlled NOT gates and single qubit operations used in quantum networks. In this article, it is supposed that the unknown state to be transferred is an arbitrary but definite $`S`$-level pure quantum state carried on one particle labelled with $`C`$. Different from Ref. , in which the shared state is a maximally entangled EPR states of $`S`$-level, we use the multi-channel made up of $`L`$ two-level EPRs. It means that at first Alice and Bob have to prepare this group of EPRs and share each of them, with one particle of each EPR controllable to the sender and the other to the receiver. We shall see how the unknown state of $`S`$-level is teleported from $`C`$ at Alice’s place to the Bob halves of the EPRs. It is necessary here to indicate that the two Hilbert spaces are not the same, one is the single particle’s while the other is the multi-particle’s, but from the Hilbert space with more dimensions (the bigger one) we can always select a subspace equivalent to the other (the smaller one). In our case, $`2^LS`$ is required and therefore we can select $`S`$ normalized orthogonal vectors as the basis of the subspace from the $`L`$ two-level particles’ Hilbert space to make them mapping one by one to the $`S`$ eigenvectors of $`C`$. Two states respectively in the two sorts of Hilbert space will be regarded as the same if the coefficients are the same when expressed as the linear superposition of their own basis. Only in this means can we say that the state on $`C`$ is teleported onto the $`L`$ particles. We label all the EPRs with serial numbers $`0,1,\mathrm{},L1`$, while the corresponding particles at Alice’s place and Bob’s are labelled respectively $`A_0,A_1,\mathrm{},A_{L1}`$ and $`B_0,B_1,\mathrm{},B_{L1}`$. The EPR entangled state of each pair of particles $`A_k`$ and $`B_k`$ ( $`k=0,1,\mathrm{},L1`$ ) can be chosen as follows $$|\mathrm{\Phi }_{A_kB_k}=\frac{1}{\sqrt{2}}\left(|0_{A_k}|0_{B_k}+|1_{A_k}|1_{B_k}\right)$$ (1) where we express the eigenvectors of the two-level particles as $`|0,|1`$ which in the case of $`\frac{1}{2}`$-spin particles, for example, refer to spin-up state and spin-down state respectively.. Moreover, the state of $`C`$ is generally written as $$|\psi _C=\underset{m=0}{\overset{S1}{}}\alpha _m|m_C$$ (2) in which $`\alpha _m`$ $`(m=0,1,\mathrm{},S1)`$ is a complex number satisfying $`\underset{m=0}{\overset{S1}{}}\left|\alpha _m\right|^2=1`$ and $`|0,|1,\mathrm{},|S1`$ denote the $`S`$ eigenvectors of the $`S`$-level particle. It is convenient that we distinguish $`|0`$ and $`|1`$ only with subscript, i.e. $`|0_{A_k}`$ or $`|0_{B_k}`$ is not the same state with $`|0_C`$, and so is $`|1_{A_k}`$ or $`|1_{B_k}`$ with $`|1_C`$. Further restriction $`2^{L1}<S`$ is set on $`L`$, since so many EPRs is the least but enough to realize our teleportation. Any number can be expressed as its binary form above which we will mark the symbol ’$``$’. For example, a number customarily in decimal form $`n`$ is decomposed into $`L`$-bit number $`n=2^{L1}n_{L1}+\mathrm{}+2^1n_1+2^0n_0`$ where $`2^Ln`$ and $`n_k=0`$ or $`1`$ $`(k=0,1,\mathrm{},L1)`$, and is written as $$n=\overline{n_{L1}\mathrm{}n_1n_0}$$ (3) On the other hand, any binary number has its decimal correspondence. If we regard the $`L`$ particles $`A_0,A_1,\mathrm{},A_{L1}`$ or $`B_0,B_1,\mathrm{},B_{L1}`$ as ’qubits’ , each state $`|n_{L1}_{A_{L1}}\mathrm{}|n_1_{A_1}|n_0_{A_0}=|n_{L1}\mathrm{}n_1n_0_A`$ or $`|n_{L1}_{B_{L1}}\mathrm{}|n_1_{B_1}|n_0_{B_0}=|n_{L1}\mathrm{}n_1n_0_B`$ ($`n_k=0`$ or $`1`$, $`k=0,1,\mathrm{}L1`$) will correspond to a binary number $`\overline{n_{L1}\mathrm{}n_1n_0}`$ and we introduce a symbol ’$`|\text{ }`$’ to simplify the denotation of the state as $$|n|n_{L1}\mathrm{}n_1n_0$$ (4) where $`n`$ has the same meaning as in Eq. 3. The quantum state of the composite system made up of $`A`$, $`B`$ and $`C`$ will thus be as follows $$|\mathrm{\Psi }_0_{ABC}=|\psi _C\underset{k=0}{\overset{L1}{}}|\mathrm{\Phi }_{A_kB_k}=\frac{1}{\sqrt{N}}\underset{m=0}{\overset{S1}{}}\underset{n=0}{\overset{N1}{}}\alpha _m|m_C|n_A|n_B$$ (5) where $`N=2^L`$. In principle, Alice is able to perform on the composite system $`AC`$ any quantum operations, including local or nonlocal unitary transformations and measurements. To realize the teleportation, a nonlocal unitary transformation $`U_{AC}`$ to all the bodies included in system $`AC`$ is performed. $`U_{AC}`$ will realize the following transformation $$U_{AC}|m_C|n_A=\frac{1}{\sqrt{S}}\underset{j=0}{\overset{S1}{}}e^{i\frac{2mj\pi }{S}}|j_Cf^n(j,m)_A$$ (6) in which $`m=0,1,\mathrm{},S1`$, and $`f^n(j,m)`$ is a number of decimal form determined by $`j`$, $`m`$ and $`n`$ so that $`f^n(j,m)`$ is one of the $`N`$ eigenstates. If we also express $`j`$, $`m`$ and $`f^n(j,m)`$ as the binary form $`j`$ $``$ $`\overline{j_{L1}\mathrm{}j_1j}_0`$ (7) $`m`$ $``$ $`\overline{m_{L1}\mathrm{}m_1m_0}`$ $`f^n(j,m)`$ $``$ $`\overline{f_{L1}^n(j,m)\mathrm{}f_1^n(j,m)f_0^n(j,m)}`$ $`j_k,m_k,f_k^n(j,m)`$ $`=`$ $`0,1(k=0,1,\mathrm{}.L1)`$ $`f^n(j,m)`$ will be determined by $`f_k^n(j,m)`$s that satisfy $$f_k^n(j,m)=n_kj_km_k$$ (8) where ’$``$’ denotes addition modulo $`2`$. One can easily prove the unitarity of $`U_{AC}`$ and show that when any two among $`j`$, $`m`$ and $`n`$ are definite, $`f^n(j,m)`$s different in the parameter of the rest will be orthogonal mutually. For example, $$f^n(j,m^{})f^n(j,m)=\delta _{m^{}m}$$ (9) Where $`m,m^{}=0,1,\mathrm{},S1`$. Eq. 9 means that any two basis among $`f^n(j,m)`$s with the same $`n`$ and $`j`$ but different $`m`$ will not be the same. After the transformation of $`U_{AC}`$, due to Eq. 68, the quantum state of system $`ABC`$ will change to $$|\mathrm{\Psi }_{ABC}=U_{AC}|\mathrm{\Psi }_0_{ABC}=\frac{1}{\sqrt{S}}\underset{j=0}{\overset{S1}{}}\left\{\right|j_C\frac{1}{\sqrt{N}}\underset{n=1}{\overset{N1}{}}\left(\right|n_A\underset{m=0}{\overset{S1}{}}\alpha _me^{i\frac{2mj\pi }{S}}f^n(j,m)_B)\}$$ (10) which is the entangled quantum state involving all the particles in system $`ABC`$. If now Alice performs measurements to the single particles $`C,A_0,A_1,\mathrm{},A_{L1}`$, with the same possibility of $`\frac{1}{NS}`$, she will acquire one of the outcomes, i.e., the collapse of the state of these particles to the possible eigenstate $`|j_C|n_A`$ ($`j=0,1,\mathrm{},L1`$ and $`n=0,1,\mathrm{},N1`$). Thus the entanglement among $`A`$, $`B`$ and $`C`$ will be destroyed and Bob will acquire the state of $`B`$ $$|\psi ^n(j)_B=_{m=0}^{S1}\alpha _me^{i\frac{2mj\pi }{S}}f^n(j,m)_B$$ (11) which is an entangled quantum state of particles $`B_0,B_1,\mathrm{},B_{L1}`$. If $`n`$ and $`j`$ are definite, Eq. 9 ensure that we can redefine $`S`$ of $`B`$’s basis as $`|m^{}e^{i\frac{2mj\pi }{S}}f^n(j,m)`$ $`|m^{}=e^{i\frac{2mj\pi }{S}}f^n(j,m)`$, where $`|0^{},|1^{},\mathrm{},|S1^{}`$ form the basis of the subspace of system $`B`$’s Hilbert space. Therefore we get $$|\psi ^n(j)_B=\underset{m=0}{\overset{S1}{}}\alpha _m|m_B^{}$$ (12) According to our discussion in paragraph 4 and the comparison of Eq. 2 and 12, we can regard $`|\psi ^n(j)`$ and $`|\psi `$ as the same. However, we need indicate that $`|m^{}`$ lies on $`j`$ and $`n`$, which makes it is still necessary to build the classical channels between Alice and Bob to transfer the information about Alice’s outcomes, or the information of $`j`$ and $`n`$ in the other words, since Bob will not know exactly what the $`|m^{}`$ means without the knowledge of $`j`$ and $`n`$. Just the necessity of classical information transferring makes the faster-than-light communication impossible. We have discussed above the possibility, in principle, the possibility of teleportation of any $`S`$-level quantum states by no less than $`L=\mathrm{log}S`$ two-level EPRs. In our discussion, we use the complicated unitary transformation $`U_{AC}`$, which means the evolution of the quantum state of system $`AC`$ under the interaction of all those particles involved in $`AC`$. The complication of $`U_{AC}`$ leads to the complication of operation. It is even impossible for us to operate such a transformation unless we take further consideration. The method of quantum computational networks has shown out the most feasible way of realizing the operation. The quantum computational networks has been much studied in Ref. . Following their method, we make the transformation more operationable by decomposing $`U_{AC}`$, which is to $`2L+1`$ particles, into a sequence of two-body unitary transformations and a simple single-body unitary transformation. Only two classes of such transformations are used: (a) the discrete Fourier transform modulo $`S`$, denoted $`DFT_S`$, which is a unitary transformation in $`S`$ dimensions.. It is defined relative to the basis $`|0_C,|1_C,\mathrm{},|S1_C`$ by $$DFT_S|m_C=\frac{1}{\sqrt{S}}\underset{j=0}{\overset{S1}{}}e^{i\frac{2mj\pi }{S}}|j_C$$ (13a) (b) a combined unitary transformation $`U_{Ck}`$ to the two particles $`C`$ and $`A_k`$ ($`k=0,1,\mathrm{},L1`$). $`U_{Ck}`$ is defined by $$U_{Ck}|m_C|n_k_{A_k}=|m_C|m_kn_k_{A_k}$$ (14) $`U_{AC}`$ can be decomposed into the product of these two classes of transformation $$U_{AC}=\left(\underset{k=0}{\overset{L1}{}}U_{Ck}\right)DFT_S\left(\underset{k=0}{\overset{L1}{}}U_{Ck}\right)$$ (15) where because $`[U_{Ck^{}},U_{Ck}]=0`$ for any $`k,k^{}=0,1,\mathrm{},L1`$, we need not distinguish their order. By 15 we simplify the problem in operation of $`U_{AC}`$, for the quantum operation on two bodies is far more feasible than on a lot of bodies. In summary, we construct the scheme of transferring an arbitrary $`S`$-level quantum state by using two-level EPRs. The importance of this construction lies not only on the scheme itself, but also on the possibility of further research and application of teleportation. It leads us to more general, more feasible and simultaneously more challenging considerations on the problem of teleportation. A lot of questions, such as probabilistic teleportation and teleportation of unknown quantum states by definite number of EPRs, are thus put forward before us, waiting for us to solving. ACKNOWLEDGMENTS We would like to thank all the other members of our research group for their helpful discussions. They are Guang Hou, Shengjun Wu, Prof. Qiang Wu, Yifan Luo, Minxin Huang, Miss Jie Yang, Guojun Zhu, Ganjun Zhu and Meisheng Zhao. We have also benefited from the cooperation with Prof. Anton Zeilinger’s group at Austria. We are grateful to them especially Dr. Jianwei Pan for their helpful information and suggestions.
warning/0004/cond-mat0004085.html
ar5iv
text
# Motional dressed states in a Bose condensate: Superfluidity at supersonic speed ## ACKNOWLEDGMENTS We acknowledge the support from the Chinese University of Hong Kong Direct Grant No. 2060148. C.K.L. is supported by a postdoctoral fellowship at the Chinese University of Hong Kong.
warning/0004/astro-ph0004234.html
ar5iv
text
# Black Hole Emergence in Supernovae ## 1. Introduction Core-collapse supernovae mark the death of a massive star and the birth of a compact remnant. Theory suggests that the remnant can be either a neutron star or a black hole, depending on the character of the progenitor and the details of the explosion. Radio pulsar emission has allowed to compile a variety of observational evidence associating neutron stars with sites of known supernovae, but similar evidence for a black hole - supernova connection is still mostly unavailable. Interestingly, indirect evidence that the black hole candidate in the X-ray binary system GRO J1655-40 was formed in a supernova explosion has recently been reported by Israelian et al. (1999). The inference is based on detection of high abundances of nitrogen and oxygen on the surface of the companion, which has too low a mass to have produced them by thermonuclear burning, hence indicating that they were deposited there by the supernova that created the black hole. The formation of a compact object in a supernova can be inferred directly if the object causes an observable effect on the total luminosity which follows the explosion – i.e., the light curve. In particular, some material from the base of the expanding envelope may remain bound to the compact object and continuously fall back onto it, thus generating an accretion luminosity. Zampieri et al. (1998a) have recently performed a self-consistent investigation of the accretion luminosity generated by spherically symmetric “fallback” of matter onto a black hole in the wake of a supernova. With the aid of a general-relativistic, radiation hydrodynamic Lagrangian code they showed that, at late times, the fallback evolves quasi-stationarily with the accretion luminosity obeying the analytic expression found by Blondin (1986) for steady state, spherical hypercritical accretion<sup>1</sup><sup>1</sup>1We use the term “hypercritical” to describe an accretion rate $`\dot{M}`$ that largely exceeds the Eddington limit for the accreting object, i.e., $`\dot{M}\dot{M}_{Edd}L_{Edd}/c^2`$. The actual luminosity that arises would still be sub-Eddington if the efficiency of converting energy into radiative luminosity is low enough.. The accretion luminosity declines secularly with a time dependence of $`t^{25/18}`$. This decay is driven by the decrease in the accretion rate due to the continuous expansion of the supernova envelope (Colpi, Shapiro & Wasserman 1996), Such an accretion luminosity will produce a distinct signature on the total light curve if and when it becomes comparable to the output of other power sources, namely the initial internal energy of the envelope and decays of radioactive isotopes synthesized in the explosion. It is well established that the efficiency of converting spherical accretion onto a black hole into radiative luminosity may be quite low (Shapiro 1973; Blondin 1986; Wandel, Yahil & Milgrom 1986; Park 1990; Nobili, Turolla & Zampieri 1991; Zampieri, Miller & Turolla 1996), and, in general, the accretion luminosity will be undetectable in comparison with the luminosity generated by radioactive decays. This is particularly important when compared to the case of accretion onto a neutron star, which is expected to be far more efficient in generating an accretion luminosity. For the last decade this distinction has been specifically applied to the case of SN1987A: Chevalier (1989) and Houck & Chevalier (1991) found that accretion onto a newly formed neutron star in SN1987A would generate an accretion luminosity at the Eddington-rate ($`10^{38}`$ergs$`\text{s}^1`$) within a few months after the explosion, and the luminosity would persist at this rate for several years<sup>2</sup><sup>2</sup>2It is noteworthy that magnetic dipole emission of a Crab-like pulsar would deposit energy in the envelope at a similar rate (Woosley et al., 1989).. The proximity of SN1987A has allowed observations to follow the light curve for over 3000 days (to a present luminosity of $`10^{36}`$ergs$`\text{s}^1`$), and the light curve has been completely consistent with heating by decays of $`{}_{}{}^{56}\mathrm{Co}`$, $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$. The absence of any evidence for accretion luminosity in the case of SN1987A has been used to argue that the compact remnant in SN1987A is a low mass black hole, implying significant consequences regarding the nuclear equation of state (Bethe & Brown, 1995). It has also been argued, however, that this absence could still be consistent with a weakly magnetized neutron star, if the accretion flow has become dynamically unstable (Fryer, Colgate & Pinto 1999). A positive indication of the presence of a black hole would be an actual detection of the accretion luminosity. The power-law time dependence would make the accretion luminosity easily discernible from luminosity generated by radioactive heating, which decreases exponentially with time. Furthermore, the difference in the time dependence implies that the accretion luminosity must eventually become dominant over radioactive decays, so in principle, the black hole always “emerges” in supernovae light curves. However, the potential for observation is usually quite slim: for example, Zampieri et al. (1998a) estimated that if a low mass black hole was formed in SN1987A, the accretion luminosity will not become dominant over that of heating by $`{}_{}{}^{44}\mathrm{Ti}`$decays until $`900`$ years after the explosion, when the luminosity will have dropped to a mere $`10^{32}`$ergs$`\text{s}^1`$. The abundance of radioactive elements and the amount of fallback in SN1987A are rather typical of core collapse supernovae of low to intermediate mass progenitors (12–25 $`M_{}`$) so that, if a black hole is formed, the “emergence” of the accretion luminosity occurs only at such late times and low luminosities that it is of no observational consequence. An important exception may exist, however, in the explosions of more massive stars. In their survey of supernova explosions and nucleosynthesis, Woosley & Weaver (1995) find that more massive stars tend to produce more massive compact remnants: the original iron core is larger, leading to a more massive proto-neutron star and to larger amounts of fallback (see also Fryer 1999). Correspondingly, larger mass progenitors are more likely to produce black holes (as the remnant mass would be larger than the maximum mass of stable neutron stars), and to have larger amounts of material that remain bound to the black hole as a reservoir for late time fallback. Just as important, the final abundance of the key radioactive isotopes ($`{}_{}{}^{56}\mathrm{Ni}`$, <sup>57</sup>Ni and $`{}_{}{}^{44}\mathrm{Ti}`$) in the expanding envelope is quite sensitive to the location of the cutoff that defines the material which settles on the black hole during the early stages (while the explosion is still proceeding). For the more massive progenitors the cutoff is further out, and since the radioactive elements are synthesized in the innermost layers above the proto-neutron star, the total mass of these elements in the ejected envelopes of higher mass stars can be significantly smaller than in “standard” supernovae. In fact, progenitor of masses in the range $`3040M_{}`$, are expected to be practically free of radioactive elements that can power the bolometric light curve (Woosley & Weaver, 1995). For such stars, black hole accretion luminosity should emerge immediately after the decay of the recombination peak (tens of days after the explosions), while it is still relatively powerful, and hence potentially observable. Naturally, the more massive stars are a small minority in the stellar population, as are their explosions among core collapse supernovae. However, one recent supernova, SN1997D in NGC 1536 (deMello & Benetti, 1997), may present a border-line candidate for a direct detection of black hole emergence. Its observed light curve indicates that that the total mass of $`{}_{}{}^{56}\mathrm{Co}`$(the daughter of $`{}_{}{}^{56}\mathrm{Ni}`$) in the envelope is only $`2\times 10^3M_{}`$, a factor of $`35`$ lower than SN1987A. By analyzing the light curve and spectra, Turatto et al. (1998) suggested that SN1997D was a low energy ($`E_{tot}4\times 10^{50}`$ergs) explosion of a $`26M_{}`$ star. The expected mass of the remnant is about $`3M_{}`$ (formed by a $`1.2M_{}`$ early fallback on a $`1.8M_{}`$ collapsed core), hence most likely a black hole. Based on the inferred average properties of the material available for late time fallback and on the late time behavior of the accretion luminosity found by Zampieri et al. (1998a), Zampieri, Shapiro & Colpi (1998b) estimated that accretion luminosity will emerge in SN1997D in the optical band within only a few years after the explosion. In this work we revisit the issue of black hole emergence in supernova due to spherically-symmetric late-time accretion. In particular, we investigate a more robust estimate for our “proto-type” case, SN1997D. We furnish revised analytic estimates and a detailed numerical investigation of the radiation-hydrodynamic evolution of the ejecta, incorporating the effects of a realistic chemical composition, opacities and radioactive heating on the accretion history and luminosity. The numerical study is based on an improved version of the radiation-hydrodynamic code used by Zampieri et al. (1998a), where the main modifications are the inclusion of realistic envelope composition and opacities (rather than pure hydrogen) and of heating by radioactive decays. We find that a $`3M_{}`$ black hole should emerge in the SN1997D light curve at about 1000-1500 days after the explosion – i.e., during late 1999 and 2000 – when the total bolometric luminosity is $`0.53\times 10^{36}`$ergs$`\text{s}^1`$, and is marginally detectable with HST. In the course of our computational study we were compelled to introduce a rescaling scheme in order to numerically follow the accretion history and light curve for several years. The huge dynamical range of pertinent time scales renders a full scale time-integration numerically impractical. Our solution has been to study a scaled model which evolves more quickly and allows a simple relation between the properties and history of the rescaled model and the true one. We note that this scheme, which is presented in detail in the appendix, is potentially suitable for other problems which involve time-dependent radiation hydrodynamics spanning a large dynamic range. We begin our discussion in § 2 with a summary of the characteristic features of supernova light curve evolution when accretion onto a central black hole is occurring. Quantitative relations between the light curve magnitude and the relevant time scales are also reviewed. The principal features of the numerical code of Zampieri et al. (1998a) and the modifications incorporated for this work are described in § 3. The light curve of SN1997D is examined in detail in § 4, where we present analytic and numerical results, including our estimates regarding black hole emergence. Prospects of observing black hole emergence in other supernovae are discussed in § 5, and further discussion and conclusions are offered in § 6. ## 2. The Supernova Light Curve including Accretion The bolometric luminosity that follows the explosion of a massive star is powered by the internal energy of the expanding envelope, emitted as thermal photons. Under “bolometric” we do not include the hard X-ray and $`\gamma `$ray photons that are emitted in radioactive decays and escape from the envelope before thermalization. In most core collapse supernovae, luminosity in the early (up to tens of days) light curve is powered by energy deposited in the envelope by the supernova shock. After this energy is depleted, only ongoing sources of energy can continue to power the later part of the light curve, often referred to as a “tail”. By definition, radiative emission from late fallback qualifies as an ongoing source, as do decays of radioactive elements in the envelope<sup>3</sup><sup>3</sup>3 Young & Branch (1989) pointed out that if the energy generation from radioactive decays is sufficiently high and the progenitor radius is sufficiently small, radioactive decays can become the dominant source of power for the light curve even very early after the explosion. They suggest that this is the case in Type II-linear supernovae.. It is useful to examine the evolution of the light curve in terms of the key characteristic time scales and luminosity scales, which can be estimated on the basis of the explosion energy and the average properties of the progenitor (Arnett 1980, 1996). When accretion onto the central object caused by late-time fallback is taken into account, additional time scales and a luminosity scale must also be included (Zampieri et al., 1998a). We briefly review these basic estimates below. They will serve also as the basis for establishing the rescaling scheme presented in the appendix. ### 2.1. Expansion For an envelope with initial radius $`R_0`$ and outer velocity $`V_0`$ we can define an initial expansion time scale, $`t_0`$, as $$t_0=\frac{R_0}{V_0}.$$ (1) Assuming that after the passage of the shock the envelope settles into free streaming and homologous expansion (as is typical in strong spherical shocks), the initial velocity of a shell at radius $`r`$ in the expanding envelope is simply $`v_0(r)=r/t_0`$. For free streaming the radius of the shell will follow $`rt`$, and since the expansion is nearly adiabatic, the temperature $`T`$ and density $`\rho `$ of the radiation-pressure dominated gas decline according to $`Tt^1`$ and $`\rho t^3`$. These quantities determine the leakage rate of thermal photons from the envelope material, and hence govern the history of the early light curve. ### 2.2. Diffusion During the time that the initial internal energy is the dominant source for radiative luminosity, the evolution of the light curve is governed by the properties of the optically-thick, hydrogen-rich outer layer, that holds most of this energy. We denote the mass, initial average density and initial average temperature of this layer as $`M_H,\rho _{0,H}`$ and $`T_{0,H}`$, respectively. The initial post-shock temperature of the envelope is generally high enough ($`T_{0,H}10^6`$$`{}_{}{}^{}\mathrm{K}`$) so that the envelope material is highly ionized, and the dominant source of opacity, $`\kappa `$, is Thomson scattering. Correspondingly, the thermal photons are mostly trapped in the envelope, and a radiative luminosity occurs through diffusion of thermal energy to the surface. For a homogeneous expanding envelope this diffusion luminosity can be approximated by (Arnett, 1996) $$L_{diff}(t)=L_{diff,0}e^{(t_0t+t^2/2)/(t_0t_{diff,0})},$$ (2) where $$L_{diff,0}=\frac{(\frac{4}{3}\pi \beta ac)T_{0,H}^4R_0^4}{\kappa M_H},t_{diff,0}=\frac{3\kappa \rho _0R_0^2}{\beta c},$$ (3) with $`\beta `$ a numerical factor which depends on the temperature profile. The value $`\beta =13.8`$ (exact for the “radiative-zero” temperature profile, $`T_0(r)=T_0(\mathrm{sin}(\pi \frac{r}{R_0})/(\pi \frac{r}{R_0}))^{1/4}`$) is usually a reasonable estimate (Arnett, 1980). The time dependence of equation (2) includes the expansion time $`t_0`$ and the initial diffusion time $`t_{diff,0}`$, which is the characteristic time scale for radiation to cross the envelope through diffusion at the onset of expansion. The characteristic time for radiation to diffuse out of the expanding envelope is given by (Zampieri et al., 1998a) $`t_{diff}\sqrt{t_0t_{diff,0}}=\left[{\displaystyle \frac{1}{4\pi }}\kappa M_H\right]^{1/2}{\displaystyle \frac{1}{(cV_0)^{1/2}}}`$ (4) $`\left[{\displaystyle \frac{1}{3}}\kappa R_0^3\rho _{0,H}\right]^{1/2}{\displaystyle \frac{1}{(cV_0)^{1/2}}}.`$ The diffusion approximation breaks down when expansion makes the envelope transparent to its thermal photons, after a typical time $$t_{trans}\left[\frac{3}{4\pi }\kappa M_H\right]^{1/2}\frac{1}{V_0}[\kappa R_0^3\rho _{0,H}]^{1/2}\frac{1}{V_0},$$ (5) which is significantly longer than $`t_{diff}`$. However, the onset of recombination in the envelope usually occurs early enough so that the actual value of $`t_{trans}`$ is irrelevant. ### 2.3. Recombination As the temperature of the envelope is gradually degraded by expansion, it eventually reaches the characteristic recombination temperature of the envelope material. In a hydrogen-rich envelope this temperature can be estimated by $`T_{rec,H}`$ (approximately $`10^4`$$`{}_{}{}^{}\mathrm{K}`$), so that for adiabatic expansion we find a recombination time scale $$t_{rec}=t_0\frac{T_{0,H}}{T_{rec,H}}.$$ (6) The main impact of recombination in the context of the light curve is that it imposes a sharp decrease in the opacity to the envelope’s own thermal photons: as the envelope material recombines, it rapidly transforms from opaque to transparent. This process is rapid enough so that the photons cannot readjust to the global thermal profile of the envelope; rather, a recombination “front” sweeps inward through the envelope (but outward in the laboratory frame), essentially liberating all the thermal energy which is still stored in the envelope at the time $`t_{rec}`$. As the envelope material recombines, it rapidly transforms from opaque to transparent. In many Type II supernovae the time until adiabatic expansion cools the envelope to the recombination temperature is usually smaller or equal to the diffusion time scale and significantly shorter than the transparency time scale, so a significant fraction of the initial thermal energy is still available at the time of recombination. This energy is then released over the time required for the hydrogen envelope to recombine, generating what is usually an observable peak in the light curve (SN1987A providing a proto-typical example). Roughly, the average luminosity during the recombination peak can be estimated as $$\overline{L}_{rec}=\frac{E_{thermal}(t_{rec})}{t_{rec}}=\frac{4\pi }{3}R_0^3a\frac{T_{0,H}^2T_{rec,H}^2}{t_0},$$ (7) assuming that energy losses due to photon diffusion have been negligible, which is a good approximation if $`t_{rec,H}<t_{diff}`$. In equation (7) $`a`$ is the radiation constant. Note that the nonlinear nature of the physics governing the propagation of the recombination front through the envelope limits the applicability of using the average quantities of the envelope to fully describe the light curve during the recombination phase (see more detailed estimates in Nomoto et al. 1994, Arnett 1996). ### 2.4. Radioactive Heating A complete analysis of a realistic light curve must also include the effects of heating due to decays of radioactive isotopes synthesized in the explosion. After recombination, we can estimate that the envelope is practically transparent to its thermal photons, so that the luminosity due to radioactive heating is roughly equal to the instantaneous energy deposition rate due to the decays, through thermalization of the decay products. The rate of energy deposition in the envelope by the decays of an initial total mass $`M_X`$ of a given isotope $`X`$ can be expressed as: $$Q_X(t)=M_X\left(\epsilon _{X,\gamma }f_X(t)+\epsilon _{X,e^+}\right)\text{e}^{t/\tau _X},$$ (8) where $`\tau _X`$ is the life-time of the element $`X`$. The form of equation (8) distinguishes between the photons ($`\gamma `$-rays) and positrons emitted in the decays (with energy rates per unit mass of $`\epsilon _{X,\gamma }`$ and $`\epsilon _{X,e^+}`$ respectively) since $`\gamma `$-rays are not totally trapped in the envelope, and their contribution to heating is modified by a trapping factor, $`f(t)`$. On average, the hard photons lose about half their energy in every scattering, until their energy is degraded to tens of keV. At these energies free-bound absorption on heavy elements becomes the dominant source of opacity and the photons are thermalized rapidly. Because the $`\gamma `$-ray opacity rises very rapidly with decreasing energy, we can assume to first order that once a hard photon scatters it is absorbed. The trapping factor can be simply approximated as (Woosley et al., 1989) $$f_X(t)=1\mathrm{exp}\left(\kappa _{\gamma ,X}\mathrm{\Phi }_0\left(\frac{t_0}{t}\right)^2\right),$$ (9) where $`\mathrm{\Phi }_0(t/t_0)^2`$ and $`\kappa _{\gamma ,X}`$ are the $`\gamma `$-ray column depth and opacity for the typical photons emitted in the decays of element $`X`$. Since the inner (more dense) layers are the dominant contributors to the optical depth for $`\gamma `$-rays, the helium-rich layer can be used to estimate the total $`\gamma `$ray optical depth. The time-dependent column depth is $`\mathrm{\Phi }_0\left({\displaystyle \frac{t_0}{t}}\right)^2{\displaystyle \rho 𝑑r}\rho _{He}(t)R_{He}(t)=\rho _{He,0}\left({\displaystyle \frac{t}{t_0}}\right)^3V_{He,0}t`$ (10) $`=\left({\displaystyle \frac{3M_{He}}{4\pi }}\right)^{1/3}\rho _{He,0}^{2/3}\left({\displaystyle \frac{t_0}{t}}\right)^2.`$ In equations (9-10) it is implicitly assumed that the expansion time $`t_0`$ is universal to the entire envelope; otherwise, it must be replaced with the characteristic expansion time of the helium-rich layer. Note that equation (10) allows us to express the trapping factor in terms of the initial (and presumably known) global quantities of the helium-rich layer. The quantity $`\rho _{0,He}t_0^3`$ (which appears here to the power $`\frac{2}{3}`$) is a useful one also in the context of accretion (see Chevalier 1989, and in detail below). In addition to the characteristic time $`\tau _X`$, another characteristic time involving radioactive heating is the $`\gamma `$transparency time, similarly to equation (5). Since $`\gamma `$-rays carry the bulk of the energy emitted in the radioactive decays, it is important to compare this time, $`t_{trns\gamma ,X}=\left[{\displaystyle \frac{3}{4\pi }}\kappa _{\gamma ,X}M_{He}\right]^{1/2}{\displaystyle \frac{1}{V_{0,He}}}`$ (11) $`[\kappa _{\gamma ,X}R_{0,He}^3\rho _{0,He}]^{1/2}{\displaystyle \frac{1}{V_{0,He}}},`$ with the other characteristic times governing the light curve history. At times $`tt_{trns\gamma ,X}`$ some of the hard photons emitted in the decays of element $`X`$ escape before thermalizing and do not contribute to the bolometric light curve. We note again, that equation (11) assumes, as before, that the bulk of the envelope’s opacity to hard photons is contributed by the helium-rich layer. The relevant radioactive isotopes for powering the light curve are $`{}_{}{}^{56}\mathrm{Ni}`$and its daughter nucleus $`{}_{}{}^{56}\mathrm{Co}`$, $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$. We adopt the characteristic parameters for these isotopes from Woosley et al. (1989), listed in table 2.4. Note that in a realistic application, especially when the early history of the envelope is of interest, the $`{}_{}{}^{56}\mathrm{Co}`$mass must be adjusted according to its production rate in $`{}_{}{}^{56}\mathrm{Ni}`$decays (Shigeyama & Nomoto, 1990), $$M_{{}_{}{}^{56}\mathrm{Co}}(t)=M_{{}_{}{}^{56}\mathrm{Ni}}(t=0)\left[\mathrm{exp}(t/\tau _{{}_{}{}^{56}\mathrm{Co}})\mathrm{exp}(t/\tau _{{}_{}{}^{56}\mathrm{Ni}})\right].$$ (12) | TABLE 1 | | | | | | --- | --- | --- | --- | --- | | The relevant radioactive isotopes for powering a supernova light curve: | | | | | | Energy emission rates in $`\gamma `$rays and positrons, life-times and | | | | | | effective $`\gamma `$ray opacities at the characteristic photon energies <sup>a</sup> | | | | | | isotope | $`\epsilon _\gamma `$ | $`\epsilon _{e^+}`$ | $`\tau `$ | $`\kappa _\gamma /Y_e^T^b`$ | | | (ergs$`\text{s}^1`$) | (ergs$`\text{s}^1`$) | (days) | (cm<sup>2</sup>/gm) | | $`{}_{}{}^{56}\mathrm{Ni}`$ | $`3.90\times 10^{10}`$ | 0 | $`8.8`$ | 0.06 | | $`{}_{}{}^{56}\mathrm{Co}`$ | $`6.40\times 10^9`$ | $`2.24\times 10^8`$ | $`111.3`$ | 0.06 | | $`{}_{}{}^{57}\mathrm{Co}`$ | $`6.81\times 10^8`$ | 0 | $`391.0`$ | 0.144 | | $`{}_{}{}^{44}\mathrm{Ti}`$ | $`2.06\times 10^8`$ | $`6.536\times 10^7`$ | $`3.28\times 10^4`$ | 0.073 | <sup>a</sup>adapted from Woosley et al. 1989. <sup>b</sup>$`Y_e^T=`$ the total fraction (bound and free) of electrons per nucleon ### 2.5. Accretion Late-time fallback onto the black hole can be characterized in terms of the ratio of the initial expansion time scale $`t_0`$ and the initial accretion time scale, $$t_{acc,0}=\frac{GM_{BH}}{c_{s,0}^3(\text{He})},$$ (13) where $`M_{BH}`$ is the black hole mass and $`c_{s,0}(\text{He})`$ is the initial sound speed in the helium rich layer, which we assume is the source of material for late-time accretion. The hierarchy of these two time scales determines the hydrodynamic evolution of the accretion flow (Colpi et al., 1996). If $`\stackrel{~}{k}t_{acc,0}/t_01`$, the gas has no time to respond to pressure forces, and accretion proceeds in a dust-like (pressure-free) manner from its onset. Alternatively, if $`\stackrel{~}{k}1`$ initially, accretion is at first almost unaffected by expansion and follows a sequence of Bondi-like (Bondi 1952) quasistationary states with a slowly decreasing density at large distance. Even in this latter case the flow does eventually become dust-like, since expansion continuously decreases the density and pressure in the reservoir of bound material. The transition is expected when the actual expansion and accretion time scales become comparable, yielding a transition time (Colpi et al., 1996) $$\frac{t_{tr}}{t_{acc,0}}\left(\frac{9}{2}\right)^{1/(9\mathrm{\Gamma }11)}\stackrel{~}{k}^{9(\mathrm{\Gamma }1)/(9\mathrm{\Gamma }11)},$$ (14) where $`\mathrm{\Gamma }`$ is the adiabatic index of the gas. Thus $`t_{tr}\frac{9}{2}\stackrel{~}{k}^3t_{acc,0}`$ for $`\mathrm{\Gamma }=4/3`$ and $`t_{tr}\left(\frac{9}{2}\right)^{1/4}\stackrel{~}{k}^{3/2}t_{acc,0}`$ for $`\mathrm{\Gamma }=5/3`$. The hierarchy of time scales reflects the corresponding hierarchy of radii, when comparing the marginally bound radius, $`R_{mb}`$ with the accretion radius, $`R_{acc}`$. Initially $`R_{mb,0}=(2GM_{BH}t_0^2)^{1/3}`$ , $`R_{acc,0}={\displaystyle \frac{GM_{BH}}{c_{s,0}^2(\text{He})}},`$ (15) so that if $`R_{acc,0}R_{mb,0}`$, the initial accretion is Bondi-like. As the envelope expands $`R_{acc}t`$ while $`R_{mb}t^{2/3}`$ (Colpi et al., 1996), so that, as in the case of time scales, eventually the accretion will become dust-like, even if it had begun as a fluid one. The transition time $`t_{tr}`$ roughly corresponds to the time when $`R_{mb}=R_{acc}`$. If $`t_{acc,0}/t_01`$, pressure forces are never dominant and the late time accretion rate from an homologously expanding envelope can be estimated by the dust-like solution of Colpi et al. (1996). They find that the ratio of the dust-accretion rate $`\dot{M}`$ to the Eddington accretion rate $`\dot{M}_{Edd}`$ (see below) is $$\frac{\dot{M}}{\dot{M}_{Edd}}\frac{4\pi ^{2/3}}{9}\rho _{0,He}t_0\kappa c\left(\frac{t}{t_0}\right)^{5/3}.$$ (16) The resulting luminosity produced by late-time fallback can be computed by including radiation in the calculation. A self-consistent analysis of the radiation-hydrodynamic evolution of the accretion flow from an expanding envelope (Zampieri et al., 1998a) shows that, at sufficiently late times, the evolution of the flow always proceeds as a sequence of quasistationary states. This is because the dynamical time scale at the accretion or marginally bound radius is much longer than all the relevant time scales of the flow and the radiation field in the inner accreting region. Both pressure and radiation forces are negligible at late times, and the flow is indeed dust-like and accretion rate declines as $`t^{5/3}`$. If initially $`t_{acc,0}/t_01`$ and radiation pressure is always negligible throughout the evolution, $`\dot{M}`$ is given by equation (16) with the initial characteristic parameters of the bound material. On the other hand, if $`t_{acc,0}/t_01`$ and the accretion begins as a fluid flow, the late time accretion will still settle on a dust-like solution with $`\dot{M}t^{5/3}`$, but with a modified (reduced) absolute magnitude. As a consequence of the quasi-stationary evolution of the flow and the radiation field, the late-time luminosity produced by fallback from the expanding envelope closely follows the steady-state, spherical hypercritical accretion formula derived by Blondin (1986), $`{\displaystyle \frac{L_{acc}}{L_{Edd}}}4\times 10^7\left({\displaystyle \frac{\mu }{0.5}}\right)^{4/3}\left({\displaystyle \frac{\kappa }{0.4\mathrm{cm}^2\mathrm{g}^1}}\right)^{1/3}\times `$ (17) $`\left({\displaystyle \frac{M_{BH}}{M_{}}}\right)^{1/3}\left({\displaystyle \frac{\dot{M}}{\dot{M}_{Edd}}}\right)^{5/6},`$ for a given accretion rate $`\dot{M}`$. In equation (17) $`\kappa `$ is the opacity and $`\mu `$ is the mean molecular weight per electron of the accreting material. The luminosity and accretion rate in equation (17) are measured in units of the appropriate Eddington quantities: the Eddington luminosity, $`L_{Edd}=4\pi GM_{BH}/\kappa =1.3\times 10^{38}(\kappa /0.4\mathrm{cm}^2\mathrm{g}^1)^1(M_{BH}/M_{})`$ ergs$`\text{s}^1`$, and the Eddington accretion rate, $`\dot{M}_{Edd}=L_{Edd}/c^2`$. The evolution of the late time accretion luminosity for a dust-like flow can then be expressed as (Zampieri et al., 1998b) $$L_{acc}(t)=L_{acc,0}\left(\frac{t}{t_0}\right)^{25/18},$$ (18) where $$L_{acc,0}=\mathrm{\Lambda }\left(\frac{\mu }{0.5}\right)^{4/3}\left(\frac{\kappa }{0.4}\right)^{1/2}\left(\frac{M_{BH}}{M_{}}\right)^{2/3}\rho _{0,He}^{5/6}t_{0,He}^{20/9},$$ (19) and $`\mathrm{\Lambda }1.27\times 10^{40}`$ergs$`\text{s}^1`$ (for $`\rho _{0,He}`$ in gm$`\text{cm}^3`$ and $`t_0`$ in seconds) arises from the radiative efficiency of equation (17). It is this power-law decay rate in the bolometric luminosity which characterizes the presence of a black hole in the aftermath of a supernova explosion. It is important to notice that equations (18) and (19) cannot apply at arbitrarily early times even if the initial conditions satisfy $`t_{acc,0}t_0`$. First, some build up time is naturally required for the accretion rate to reach a maximum and start decaying. This time should be comparable to the shorter of the two time scales $`t_0`$ and $`t_{acc,0}`$ <sup>4</sup><sup>4</sup>4In the case of a pure dust, the build-up time can be estimated by comparing the early time accretion rate, which builds up as $`\dot{M}t`$, (Colpi et al. 1996 (eq. )) to the late-time one. We find that the two become comparable at a time of $`1.2t_0`$.. Furthermore, if during the early evolutionary stages the accretion rate can be so large that the accretion luminosity approaches the Eddington limit, radiation pressure will modify the flow even if pressure forces were initially negligible. On the basis of equation (18), we can define a critical time $`t_{crit}`$ when the luminosity due to dust-like accretion equals the Eddington limit $`t_{crit}14.5\left({\displaystyle \frac{\mu }{0.5}}\right)^{24/25}\left({\displaystyle \frac{\kappa }{0.4}}\right)^{9/25}\left({\displaystyle \frac{M_{BH}}{M_{}}}\right)^{6/25}\times `$ (20) $`\left({\displaystyle \frac{\rho _{0,He}}{10^4\text{gm}\text{cm}^3}}\right)^{3/5}\left({\displaystyle \frac{t_{0,He}}{1\text{hr}}}\right)^{8/5}\text{hrs}.`$ According to equation (17), the critical rate at which $`L_{acc}L_{Edd}`$ is $`\dot{M}_{crit}\mathrm{\hspace{0.33em}1}`$$`M_{}\text{yr}^1`$ for black holes of several solar masses. If the luminosity reaches the Eddington limit, radiation pressure cannot be ignored in comparison to the gravitational pull of the central object, and so material near the marginally bound radius (where a significant fraction of the initially bound mass resides) may receive a sufficient impulse to become unbound. Hence, we expect that if the build-up time of the flow satisfies $`\mathrm{min}(t_0,t_{acc,0})t_{crit}`$, radiation forces will modify the accretion history and limit the accretion rate so that $`\dot{M}(tt_{crit})\dot{M}_{crit}`$ by effectively readjusting the values of $`\rho _{0,He}`$ and $`t_0`$ in the bound region. The total amount of bound material must be reduced as well. If, on the other hand, $`\mathrm{min}(t_0,t_{acc,0})t_{crit}`$ we can expect that radiation pressure will be of lesser significance throughout the accretion history and that, at late times, the accretion rate will settle on the dust solution of equation (16) with the original values of $`\rho _{0,He}`$ and $`t_0`$ (if $`t_{a,0}/t_01`$). ### 2.6. Black Hole Emergence We define the black hole “emergence time”, $`t_{BH}`$ , in a supernova light curve to be the time when the accretion luminosity is comparable to all the other sources of luminosity combined, i.e., $`L_{acc}(t_{BH})=\frac{1}{2}L_{tot}(t_{BH})`$. Even a small abundance of radioactive isotopes is sufficient to impose a bolometric luminosity that will initially dominate accretion luminosity, since the latter is inherently limited by the Eddington rate. However, continuous spherical accretion must eventually result in emergence, since the accretion luminosity (eq. ) decreases as a power law, while radioactive heating decreases at least exponentially, and even more rapidly if $`\gamma `$ray trapping is incomplete (eq. ). Quantitatively, we expect that for any specific radioactive isotope, at a time of several times its life-time, $`\tau _X`$, the heating rate is declining much more rapidly than the accretion luminosity. If the order of magnitude of the accretion luminosity and initial abundance of the dominant radioactive isotope, $`M_X`$ are known, the time of emergence of the accretion luminosity is determined mostly by the value of $`\tau _X`$, with some finer tuning due to the exact values of $`L_{acc}(t)`$, $`M_X`$, and the time dependence of $`\gamma `$ray trapping. Note that the efficiency of $`\gamma `$trapping decreases with time as $`\mathrm{e}^{(1/t^2)}`$, so the dependence of the time of emergence on $`M_X`$ and $`L_{acc}`$ is even weaker. Correspondingly, the luminosity at emergence is mostly determined by $`L_{acc}(\text{several}\tau _X)L_{acc,0}\tau _X^{25/18}`$. In accordance with § 2.5, it is reasonable to assume that after a time not longer than a few $`\times \mathrm{min}(t_0,t_{acc,0})`$ the accretion luminosity is modulated to $`L_{acc}L_{Edd}`$. Hence, for $`\mathrm{min}(t_0,t_{acc,0})`$ of the order of a few days, it is evident that for “typical” core collapse supernova, like SN1987A, with an inferred $`M_{{}_{}{}^{56}\mathrm{Co}}0.1M_{}`$ and assumed $`M_{{}_{}{}^{44}\mathrm{Ti}}10^4M_{}`$ (Woosley & Weaver, 1995), the relevant isotope for emergence is $`{}_{}{}^{44}\mathrm{Ti}`$, and the time scale for emergence is $`t_{BH}\text{several}\tau _{{}_{}{}^{44}\mathrm{Ti}}`$hundreds of years, at which time the accretion luminosity will drop to a few $`10^{32}10^{33}`$ergs$`\text{s}^1`$. Note that since about $`\frac{1}{3}`$ of the decay energy of $`{}_{}{}^{44}\mathrm{Ti}`$is emitted in the form of positrons, $`\gamma `$ray transparency will not affect these estimates significantly. On the other hand, in a supernova where the abundance of radioactive elements is reduced by a factor of $`35`$, as is expected in SN1997D (see below), $`{}_{}{}^{56}\mathrm{Co}`$and $`{}_{}{}^{57}\mathrm{Co}`$are also relevant for examining emergence. The time of emergence is reduced to hundreds of days, and the luminosity at emergence will be a few $`10^{35}10^{36}`$ergs$`\text{s}^1`$. Of course, if no radioactive elements are present, as is expected if the mass cutoff in the supernova is very far out, emergence can occur as soon as the recombination peak clears the envelope, at $`50100`$days, with $`L_{acc}10^{37}`$ergs$`\text{s}^1`$. ## 3. Equations and Numerical Method: Summary and Modifications In order to explore the possibility of detecting black hole emergence in realistic supernovae in detail, we perform a numerical study. Time-dependent simulation of the supernova envelope including accretion requires solution of the relativistic radiation hydrodynamic equations coupled to the moment equations for radiation transport and makes use of a detailed model for the ejecta that includes radioactive decay, realistic composition and opacities. The equations of relativistic, radiation hydrodynamics for a self-gravitating matter fluid interacting with radiation and the general relativistic moment equations for the radiation field have been presented in Zampieri et al. (1998a) (to which we refer for details). These equations were solved using a semi-implicit Lagrangian finite difference scheme in which the time step is controlled by the Courant condition and the requirement that the fractional variation of the variables in one time-step be smaller than 10%. A fundamental obstacle in tracking the evolution of envelope expansion and accretion for several years arises from the extremely wide diversity of relevant time scales in the problem. These range from the dynamical time in the accreting region (milliseconds close to the horizon) to days for the expansion time scale, to hundreds of days for the evolution of the light curve. Since the innermost layers of the envelope have the shortest physical time scales, the computational time can be considerably reduced with a careful choice of the location of the inner boundary. During computation, we keep the inner radial boundary of the integration domain $`R_{in}`$ at hundreds to thousands of Schwarzschild radii, far outside the black hole horizon, but still but always well within the accreting region. This measure is acceptable because the gas in the inner accreting region is in local-thermal-equilibrium (LTE) and in near free-fall, thus allowing a reliable set of boundary conditions at $`R_{in}`$ (Zampieri et al., 1998a). However, during evolution the effective optical depth in the accreting region decreases because of the secular decrease in density caused by expansion. Thus, in order to ensure that the gas is always in LTE at $`R_{in}`$, the inner radial boundary must be continuously moved inward during the simulation. We note that with this treatment, the position of the inner boundary usually limits the effects of general relativity to be quite small. However, we did not alter the general-relativistic nature of the code, especially since we cannot determine a priori how far out the inner boundary can be set (in particular at late times, when it is continuously being moved inward). Furthermore, while the accretion luminosity is insensitive to placing the inner boundary away from the black hole horizon for the canonical parameters of SN1997D, in more general scenarios it may be essential to apply the exact general relativistic formalism with an inner boundary in the strong-field domain. Even taking advantage of the increase in the physical time scales (and hence in the time step) by suitably positioning $`R_{in}`$, the computational time needed to evolve the solution up to emergence is excessive. To reach the emergence stage in a reasonable computational time, the code employs a “Multiple Time-step Procedure” (MTP), originally developed by Zampieri et al. (1998a). With this technique, the integration domain is divided into several sub-grids, and each is evolved on its own characteristic time step. Starting from the same initial conditions and integration domain, the simulation performed using this MTP proved to be faster by a factor of $`58`$ than one where no time-step acceleration algorithm was employed. We note that the most sensitive aspect of the MTP is maintaining accurate communication between the subgrids; correspondingly, it can be used safely only when there are no sharp features crossing the envelope (i.e., after recombination). The main focus of the preliminary investigation of Zampieri et al. (1998a) was to study the fundamental features of fallback in a supernova and, in particular, to determine the gross properties of the accretion history and luminosity. To this end, a simplified pure-hydrogen model of the envelope, generally sufficient to produce a typical early light curve of a Type-II supernova, was adopted. Here our goal is to examine black hole emergence in realistic supernovae, and so we expand the original survey of Zampieri et al. (1998a) to take into account for variable chemical composition of the expelled envelopes, namely radial dependent abundances of various elements. In particular, we must account for the presence of radioactive isotopes and the heating of the envelope by their decays. ### 3.1. A Variable Chemical Composition Inclusion of a variable composition is necessary to reproduce realistic supernova light curves, as has been established in studies of SN1987A (Woosley, 1988; Shigeyama & Nomoto, 1990). For the purpose of following the radiation hydrodynamic evolution of the accretion flow onto the black hole, chemical composition and opacity in the accreting region can have important quantitative and qualitative effects. First, there is an inherent quantitative dependence in the estimate of equation (17), which suggests that a pure hydrogen envelope produces an accretion luminosity larger by a factor of a few than a helium-metal-rich one. Furthermore, it is possible that the presence of metals in sufficient quantities may impose a qualitative change in the accretion history due to increased opacity where the material is partially ionized. Additional effects may arise through the the equation of state, but in the case of supersonic accretion we expect these to be negligible. In this work we study a supernova envelope with a mixtures of hydrogen, helium and oxygen, where the relative partial fractions vary along the radial profile of the envelope. Oxygen is expected to be the most abundant heavier element in the ejected envelope, and we adopt an approximation that it represents the entire metal component in the envelope’s final composition (we note that over most of the density and temperature ranges of interest, an addition of iron at solar abundance would change the frequency-integrated opacities by no more than $`50\%`$). For opacity and ionization fractions, we use the TOPS opacities available at the LANL server (Magee et al., 1995), with extrapolation to low temperatures based on Alexander & Ferguson (1994). The data are stored in the form of tables, with spacing of five entries per decade in both temperature and density for a given composition. In the course of computation the Rosseland and Plank opacities and ionization fractions are calculated with linear interpolation in temperature, density and chemical composition. The gas equation of state is approximated as an ideal gas of ions and electrons. ### 3.2. Radioactive Heating For the purpose of computing the contribution of radioactive decays to the light curve, a consistent treatment of radioactive heating is required to follow the trapping efficiency, $`f_\gamma `$, throughout the envelope. If the rate of radioactive heating is large enough it may also affect the hydrodynamic evolution of the envelope material (there exists observational evidence for such an effect by $`{}_{}{}^{56}\mathrm{Ni}`$decay in the early history of SN1987A \- the “Ni-bubble”, Arnett et al. 1989). The role of radioactive heating may be especially important in the case of interest here, since most of the radioactive elements are likely to lie in the inner part of the envelope, which is the reservoir for accretion. The exact distribution will depend on the extent of mixing caused in the process of the explosion (Shigeyama & Nomoto, 1990; Bethe, 1990). Significant heating at early times may accelerate some of the material which was initially bound to the black hole and unbind it, hence imposing a smaller late-time accretion rate. On the other hand, at late times, even a small amount of heating can maintain a finite ionization fraction, which in turn will impose a larger photospheric radius, hence affecting the luminosity. In our numerical study we introduce an effective treatment of radioactive elements by including a local heating rate. We assume that the abundance of radioactive elements is proportional to the oxygen fraction (Pinto & Woosley, 1989), with the relevant radioactive isotopes listed in table 2.4. We use the approximation discussed in § 2.4 that a $`\gamma `$-ray photon is absorbed after its first scattering. The trapping factor of the $`\gamma `$-rays emitted at radius $`r`$ in the envelope can then be estimated as $$f_{X,\gamma }(r)=1exp\left\{(_r^{R_{out}}\kappa _{X,\gamma }\rho (r)𝑑r)\right\},$$ (21) where the quantity in parenthesis is the appropriate $`\gamma `$-optical depth at radius $`r`$. In our simulation we impose a further approximation that all the trapped energy is absorbed locally at the point of emission. We note that this simplified treatment of radioactive decay energy deposition is satisfactory during the first evolutionary stages when the ejecta are still very optically thick to $`\gamma `$-rays. Later on, however, the constant opacity/single scattering assumption gives only an approximate description of the transfer of $`\gamma `$-rays through the envelope. In fact, as the ejecta become marginally thick, photons can be scattered or absorbed after traveling a significant distance, making non-local effects important. An accurate calculation of the energy deposition and the $`\gamma `$-ray luminosity would require the solution of the energy-dependent transport problem for the $`\gamma `$ rays, which we do not attempt. Our effective approach reduces the treatment of radioactive heating to an additional local energy source. By determining the total deposited energy rate $$Q=\underset{X}{}Q_X,$$ (22) through equations (8) and (21), we modify the energy equation (eq. of Zampieri et al. 1998a) by adding the energy input from radioactive decay $$ϵ_{,t}+\stackrel{~}{a}k_P(Bw_0)+p\left(\frac{1}{\rho }\right)_{,t}=Q,$$ (23) where $`P`$, $`\rho `$ and $`ϵ`$, are the pressure, mass density and internal energy per unit mass of the gas, $`\stackrel{~}{a}`$ is related to the $`00`$ coefficient of the spherically symmetric, comoving-frame metric, $`\kappa _p`$ is the Plank mean opacity, $`BaT^4`$ and $`4\pi w_0`$ is the radiation energy density (see Zampieri et al. 1998a for details). All these quantities are measured in the comoving frame. With the aid of this modified numerical code and our analytic estimates we now proceed to study the evolution and light curve of realistic supernovae including fallback onto a nascent black hole. Our emphasis in this work is the peculiar Type II SN1997D, that may provide the first direct observational evidence of black hole formation in supernovae. This source is the only current candidate where an actual observation of the light curve at emergence of the black hole may be (marginally) possible. ## 4. Black Hole Emergence in SN1997D SN1997D was serendipitously discovered on January 14, 1997 by deMello & Benetti (1997) during an observation of the parent galaxy NGC 1536. In their analysis of the bolometric light curve and spectra, Turatto et al. (1998) found that SN1997D was an extremely peculiar Type II supernova. Modeling of the early light curve indicates that the supernova was discovered close to maximum light (hence, presumably, at the recombination peak). Adopting a distance modulus of $`\mu =30.64`$ (a distance of $`14`$Mpc to NGC 1536), implied that the maximum absolute magnitude of SN1997D was $`M_V14.65(L_{max}10^{41}`$ergs$`\text{s}^1`$), fainter by about 2 magnitudes than typical Type II supernovae at maximum (Patat et al., 1994): SN1997D is possibly the most sub-luminous Type II supernova ever observed. ### 4.1. Fundamental Estimates about the character of SN1997D The most unique feature of SN1997D in the context of black hole emergence is the very low mass of $`{}_{}{}^{56}\mathrm{Co}`$deduced from the light curve tail (beginning at $`30`$days after the first observation). The magnitude and shape of the tail are consistent with a very low abundance of $`{}_{}{}^{56}\mathrm{Co}`$, $`M_{{}_{}{}^{56}\mathrm{Co}}2\times 10^3M_{}`$ (Turatto et al., 1998). This important result has been recently confirmed by Benetti et al. (2000) on the basis of a new set of photometric data which extends up to 1.5 years after the explosion. Turatto et al. (1998) found that the light curve and spectra were best reproduced with $`4\times 10^{50}`$ergs explosion of a $`26M_{}`$ main sequence star (see Young et al. 1998 for details). The progenitor structure prior to collapse was an $`8M_{}`$ helium-rich core surrounded by an $`18M_{}`$ hydrogen-rich envelope. The remnant formed in the supernova is expected to have a total mass of about $`3M_{}`$, composed of a $`1.8M_{}`$ collapsed core and about $`1.2M_{}`$ of material which was accreted in early fallback (T. R. Young 1999, private communication). The large amount of fallback is a consequence of the low explosion energy, and provides a natural explanation for the very low observed $`{}_{}{}^{56}\mathrm{Co}`$mass. Since $`3M_{}`$ is almost certainly larger than the maximum mass of neutron stars (even a rapidly rotating one; Cook, Shapiro & Teukolsky 1994), we can expect that a black hole was formed by SN1997D. We note that a different model of a low mass ($`8M_{}`$) progenitor was suggested by Chugai & Utrobin (1999) to explain the early light curve, as well as a low $`{}_{}{}^{56}\mathrm{Co}`$abundance. In such an explosion little fallback would have occurred and the remnant would be most likely a neutron star. The late time light curve shows no evidence of the presence of a neutron star (Benetti et al., 2000), which would have been observable if it were a Crab-like pulsar, or if it was spherically accreting from late-time fallback. Hence a black hole (and therefore a massive progenitor) is more consistent with the existing photometric data. Assuming that the black hole model for SN1997D is correct, its remnant includes a black hole surrounded by a low-velocity ejecta with a small abundance of radioactive elements. If late-time fallback is proceeding in spherical accretion, the black hole could emerge rather early, while it is still potentially detectable. With a preliminary analytic estimate, Zampieri et al. (1998b) found that emergence could occur as early as three years after the explosion and that the accretion luminosity at the time is roughly in the range $`10^{35}5\times 10^{36}`$ergs$`\text{s}^1`$. We have reconsidered black hole emergence in SN1997D to determine if the accretion tail persists in a realistic model of the ejecta that includes a variable composition, realistic opacities and radioactive heating and takes into account the radiation hydrodynamic evolution of the helium mantle starting from the conditions at few tens of hours after the explosion. In the following we attempt to constrain the luminosity at the epoch of emergence, using both analytic estimates and numerical simulations. ### 4.2. Analytic Estimates for SN1997D In an analytic approach, we can use the average quantities of the helium-rich layer to estimate the radioactive heating and accretion luminosities, since this layer is expected to be both the source of material for late time fallback and the main contributor to $`\gamma `$ray trapping. As a reference composition of the bound material we use the best fit model of Young et al. (1998) (see below) which is – by mass – H:0.10, He:0.45, O:0.45. The electron scattering opacity at full ionization is $`\kappa _{es}=0.22`$ and the atomic weight per electron is $`\mu 1.27`$. #### 4.2.1 Initial Density and Expansion Time Scale Note that for initially homologous expansion, the total kinetic energy of the layer can be related to its total mass, $`M`$, and radius, $`R`$, according to $$E_{kin}=\frac{3}{10}MV^2=\frac{3}{10}M\left(\frac{R}{t_0}\right)^2,$$ (24) where $`V`$ is the velocity at the outer radius and $`t_0R/V`$ is the corresponding expansion time scale. By expressing the radius in terms of the mass and initial average density, $`\rho _0`$ we can write $$(\rho _0t_0^3)=\left(\frac{3}{10}\right)^{3/2}\frac{3}{4\pi }M\left(\frac{M}{E_{kin}}\right)^{3/2}.$$ (25) Equation (24) is especially convenient, since the average energy per unit mass deposited in a roughly homogeneous layer is relatively easily determined in simulations (see Woosley 1988 for SN1987A). The fraction of the total explosion energy carried as kinetic energy of the helium-rich layer is typically a few percent. Adopting the global quantities of the model of Turatto et al. (1998), we have in the case of of SN1997D a helium-rich layer with $`M_{He}5M_{}`$, $`E_{kin,He}1.2\times 10^{49}`$ergs, or $`E_{kin,He}/M_{He}1.2\times 10^{15}`$ergs$`\text{gm}^1`$. For this layer $$\rho _0t_0^39.4\times 10^9\text{gm cm}^3\text{s}^3.$$ (26) Note that the relatively low energy of the explosion is clearly reflected in this quantity, which for SN1987A is only $`10^9\text{gm cm}^3\text{s}^3`$ (Chevalier, 1989). The value of $`\rho _0t_0^3`$ is sufficient for estimating $`\gamma `$ray trapping (eq. ), but additional information is required for estimating the accretion luminosity, which scales as $`(\rho _0t_0^3)^{5/6}t_0^{1/4}`$ (eq. ). This difference is an inevitable consequence of the fact that the bound material is not actually expanding homologously, since the gravitational deceleration cannot be neglected. For a realistic supernova envelope there is no time instant when the entire envelope is in coherent homologous motion. As a first approximation, we assume that the entire envelope has a common expansion time, which turns out to be $`t_030`$hrs for an initial radius of $`2\times 10^{13}`$ and a hydrogen envelope of $`18M_{}`$. We also have $`\rho _{He,0}7.45\times 10^6`$gm$`\text{cm}^3`$ for these parameters. The total mass bound to a black hole of $`3M_{}`$ is then $$M_{bnd}=\frac{4\pi }{3}\rho _{He,0}R_{mb,0}^3=\frac{8\pi }{3}GM_{BH}t_0^2\rho _{0,He}=0.136M_{},$$ (27) where we have used equation (15). #### 4.2.2 Accretion Time Scale Homologous expansion implies that the thermal energy is sufficiently smaller than the kinetic energy. In particular, in order to have $`E_{th}/E_{kin}0.1`$ for the helium-rich layer presented above, the temperature in the layer must be $`T_{0,He}8\times 10^5`$$`{}_{}{}^{}\mathrm{K}`$. The corresponding gas and radiation pressure are of similar magnitude, leading to an estimate of the initial sound speed in this layer of $`c_{s,0}(\text{He})10^7`$cm$`\text{s}^1`$. The initial accretion time is thus $`t_{acc,0}120`$hours, significantly longer than the expansion time scale. In this analytic estimate, the accretion flow is approximately dust-like from its onset. As long as we assume that the total thermal energy is the order of $`10\%`$ of the kinetic energy of the helium layer, which itself is well constrained, the temperature in this layer is approximately fixed at the above value. The initial gas pressure scales as $`P_0T\rho _0`$, so the sound speed cannot decrease significantly below $`c_{s,0}(\text{He})10^7`$cm$`\text{s}^1`$(if the pressure is gas dominated $`c_{s,0}^2(\text{He})P_0/\rho _0T`$), and the initial accretion time scale (13) cannot deviate sharply from $`t_{acc,0}100`$hrs, even if the combination of $`\rho _0`$ and $`t_0`$ is allowed to vary. #### 4.2.3 Critical Time A key feature in this analytic estimate is that $`t_{crit}180`$hrs (eq. ) or $`6t_0`$. As discussed in § 2, dust-like accretion should require a time of about $`t_0`$ to settle into the late-time temporal decay. Hence it is likely that radiation forces will impose some modification to the accretion flow. We can expect that when the accretion flow finally resettles on a dust-like motion the total bound mass will be reduced, and the average density and expansion time in the bound region will be modified. The initial values of $`\rho _0`$ and $`t_0`$ can thus provide an upper limit on the late time accretion rate and luminosity through equations (16) and (18). The effect of radiation pressure on the early-time accretion flow when the luminosity is close to the Eddington limit is non-linear, and therefore cannot be ascertained from our analytic treatment. However, by making two simplifying assumptions we can roughly estimate the impact of the Eddington-limited stage on the late-time accretion rate. First, we assume that the accretion flow adjusts so that the accretion luminosity is exactly the Eddington value. Effective gravity is then negligible everywhere. Thus, while the average density $`\stackrel{~}{\rho }_0(t)`$ and expansion time $`\stackrel{~}{t}_0(t)`$ maintain a constant $`\stackrel{~}{\rho }_0(t)\stackrel{~}{t}_0(t)^3=\rho _0t_0^3`$, individually they vary according to $$\stackrel{~}{\rho }_0(t)=\left(\frac{R_0+V_0t}{R_0}\right)^3,\stackrel{~}{t}_0(t)=t_0\left(\frac{R_0+V_0t}{R_0}\right).$$ (28) The combination $`\rho _0t_0^{8/3}`$, which determines the magnitude of the dust-like accretion rate (eq. ) evolves as $$\stackrel{~}{\rho }_0(t)\stackrel{~}{t}_0^{8/3}(t)=\rho _0t_0^{8/3}\left(1+\frac{t}{t_0}\right)^{1/3}.$$ (29) The second assumption is that the transition to a dust-like flow is instantaneous, and that at the time of transition, $`t_{dust}`$, the values $`\stackrel{~}{\rho }_0(t_{dust})`$ and $`\stackrel{~}{t}_0(t_{dust})`$ exactly yield $`\dot{M}(t_{dust})=\dot{M}_{crit}`$ through equation (16). From equation (28) we find that $`t_{dust}`$ must then satisfy the relation: $`41.15\left({\displaystyle \frac{\mu }{0.5}}\right)^{4/3}\left({\displaystyle \frac{\kappa }{0.4}}\right)^{1/2}\left({\displaystyle \frac{M_{BH}}{M_{}}}\right)^{1/3}`$ $`\times \left({\displaystyle \frac{\rho _{0,He}}{10^4\text{gm}\text{cm}^3}}\right)^{5/6}\left({\displaystyle \frac{t_{0,He}}{1\text{hr}}}\right)^{5/6}`$ (30) $`\times (1+{\displaystyle \frac{t_{dust}}{t_0}})^{5/18}\left({\displaystyle \frac{t_{dust}}{t_0}}\right)^{25/18}=1.`$ The modification due to radiation pressure is that $`t_{dust}<t_{crit}`$, so that the late-time accretion luminosity is $`L(t)=L_{Edd}(t/t_{dust})^{25/18}`$ instead of $`L(t)=L_{Edd}(t/t_{crit})^{25/18}`$. We find that $`t_{dust}129`$hrs, so the late time accretion luminosity will be reduced by a factor of $`0.633`$ compared to that predicted by equation (18) for the initial parameters. Note that in the limit $`t_{dust}t_0`$ equation (4.2.3) reduces to the form $`t_{dust}^2\rho _0t_0^3`$, so $`t_{dust}`$ is dependent only on the initial specific energy of the bound material. This is a natural consequence of the approximation used here, since ongoing homologous expansion eventually loses memory of the initial size $`R_0`$ (which determines $`t_0`$ and $`\rho _0`$ separately). For the values presented above, $`t_{dust}`$ is larger than $`t_0`$ by only a factor of a few so the above argument is not exact, but if $`t_0`$ is confined to a reasonably small range, the late-time accretion luminosity can still be very well constrained by the specific energy of the helium-rich layer. Indeed, we find that $`t_{dust}`$ varies by less than $`10\%`$ when $`t_0`$ is varied in the range $`2040`$hrs if $`\rho _{0,He}t_0^3`$ is kept fixed. #### 4.2.4 Radioactive Heating and an Estimate of Emergence With all the necessary values for estimating the late-time accretion rate and $`\gamma `$-ray opacities, we now proceed to examine the competition between the accretion and radioactive luminosities. The exact time and luminosity at emergence depend on the abundances of $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$in the envelope. These two isotopes are mostly produced in deeper layers than those where $`{}_{}{}^{56}\mathrm{Co}`$is synthesized (especially $`{}_{}{}^{44}\mathrm{Ti}`$, which is produced through Nuclear-Statistical-Equilibrium (NSE) in the innermost layers of the shocked envelope material, Timmes et al. 1996). We examine their effects with two extreme assumptions – either that they are completely absent, or that their abundances scale with that of $`{}_{}{}^{56}\mathrm{Co}`$. In the latter case, using $`M_{{}_{}{}^{57}\mathrm{Co}}/M_{{}_{}{}^{56}\mathrm{Co}}(\text{SN1987A})=0.0243,M_{{}_{}{}^{44}\mathrm{Ti}}/M_{{}_{}{}^{56}\mathrm{Co}}(\text{SN1987A})=1.33\times 10^3`$ we obtain $`M_{{}_{}{}^{57}\mathrm{Co}}(\text{SN1997D})5\times 10^5M_{},M_{{}_{}{}^{44}\mathrm{Ti}}(\text{SN1997D})2.7\times 10^6M_{}`$. Assuming that the opacity to $`\gamma `$rays is independent of the accretion history (as is reasonable, since the initially bound region is a small fraction of the total helium-rich layer), we can treat the radioactive heating history as fixed once the initial combination $`\rho _{0,He}t_0^3`$ is given (eq. ). The resulting luminosities due to accretion and radioactive heating for our reference parameters of $`t_0=30`$hrs and $`\rho _{0,He}=7.45\times 10^6`$gm$`\text{cm}^3`$are shown in Fig. 4.2.4. If the abundances of $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$are negligible, the total luminosity due to radioactive heating is equal to the heating rate through decays of $`{}_{}{}^{56}\mathrm{Co}`$. In this case, the accretion luminosity does emerge with $`L_{acc}(t_{BH})=L_{{}_{}{}^{56}\mathrm{Co}}(t_{BH})3.6\times 10^{35}`$ergs$`\text{s}^1`$at a time of $`t_{BH}1216`$days (April 2000), after which the light curve will settle on a power law decline rather quickly. If $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$are present with abundances rescaled from SN1987A their impact on the luminosity at $`1000`$days is not negligible, and accretion luminosity must be compared to the total radioactive heating (labeled $`L_{totrad}`$ in Fig. 4.2.4). In this case emergence is delayed to $`t_{BH}1466`$days (December 2000), with $`L_{acc}(t_{BH})=L_{totrad}(t_{BH})2.9\times 10^{35}`$ergs$`\text{s}^1`$. Note that in this case the contribution of $`{}_{}{}^{44}\mathrm{Ti}`$is about $`1.4\times 10^{35}`$ergs$`\text{s}^1`$, which decreases only very slowly due to increasing $`\gamma `$ray transparency. Correspondingly, the total luminosity (labeled $`L_{tot}`$) does not follow a perfect power law until $`{}_{}{}^{44}\mathrm{Ti}`$heating becomes negligible, although the accretion luminosity will remain the dominant source of luminosity for several thousands of days after $`t_{BH}`$. The results of similar calculations for the emergence of a $`3M_{}`$ black hole in SN1997D for different values of the initial parameters are presented in table 1. We vary the energy of the $`5M_{}`$ helium-rich layer as 1, 3 (the nominal case) and 5 percent of the total explosion energy of $`4\times 10^{50}`$ergs, while varying $`t_0`$ in the range $`2040`$hrs. We list the derived values of $`t_{crit}`$ and $`t_{dust}`$ for each model and the resulting luminosity and time of emergence. Due to uncertainty regarding the abundances of $`{}_{}{}^{56}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$, we examine the two extreme cases where their abundances scale according to SN1987A, and when they are completely absent. The case of the helium-rich layer having $`5\%`$ of the total energy and with the significant abundance of $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$marks the border-line of achieving emergence after a few years (instead of in thousands of years) as the the combination of $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$heating is slightly larger than the accretion luminosity (by about $`20\%`$) for a very long time. Our results indicate that if a black hole was formed, it is likely to emerge about three years after the explosion. Taking into account also for the uncertainties in composition and black hole mass, we estimate that plausible values of the luminosity at emergence seems to lie in the range $`0.33\times 10^{36}`$ergs$`\text{s}^1`$, and the expected time of emergence is $`11001500`$ days after the explosion, with higher luminosities corresponding to earlier times. We emphasize that the dominant quantity that determines the value of the luminosity at emergence is the kinetic energy of the helium-rich layer, while the time of emergence and the character of the late-time light curve are also sensitive to the possible presence of $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$in non-negligible abundances. ### 4.3. Numerical Results for SN1997D While an approximate analytic treatment and scaling behavior can provide useful estimates of the time and luminosity at emergence, it also underscores the need to perform numerical investigation which can track in detail the early time accretion history and the late-time evolution of the light curve. The numerical simulation is also required to quantitatively assess the role of non-linear radiation-hydrodynamics processes and the effects of variable chemical composition, realistic opacities and radioactive heating. In order to explore in detail the emergence of a black hole in SN1997D, a full radiation-hydrodynamic simulation of fallback of material from the supernova envelope up to the several years after the explosion has been performed, using the numerical code described in § 3. The initial conditions for the ejecta were based on the best-fitted model of Young et al. (1998) for the SN1997D explosion. This model of a $`26M_{}`$ progenitor was fitted on the base of the early time light curve (up to $`100`$days after the explosion), which is practically independent of the inner part of the helium-rich layer. Hence, while a simulation based on this model serves as an essential quantitative estimate, the unavoidable uncertainties regarding the initial velocity, density and temperature profiles of this inner layer obviously limit the applicability of the simulation as a “best fit” also for the accretion history and luminosity. #### 4.3.1 Initial Profile The initial profile describes the supernova envelope at the time of break-out of the shock at the surface, corresponding to about 13 hours after core bounce. The expanding envelope is composed of two main components: an inner helium-rich layer and an outer hydrogen-rich envelope. The properties of the profile are presented in Figs. 4.3.1 and 4.3.1 and the region of initially bound material is shown in more detail in Fig. 4.3.1. By assuming that the abundance of radioactive isotopes scales with the oxygen mass fraction, we can simply estimate the mass fraction of $`{}_{}{}^{56}\mathrm{Ni}`$in every mass shell by requiring that the total $`{}_{}{}^{56}\mathrm{Ni}`$mass is $`2\times 10^3M_{}`$ (equal to the total $`{}_{}{}^{56}\mathrm{Co}`$mass inferred by observations). In the simulation we assume that the abundances of $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$scale with the abundance of $`{}_{}{}^{56}\mathrm{Ni}`$, so their total masses are set according to the aforementioned quantities. In the context of late-time accretion, the most dominant feature of the initial profile is the non-uniformity of the helium-rich layer. A significant fraction of the mass is concentrated in an over-dense region bordering with the hydrogen-rich layer. This region also holds the bulk of the kinetic energy of the helium-rich layer, $`1.7\times 10^{15}`$ergs$`\text{gm}^1`$, somewhat larger than the average value of $`1.2\times 10^{15}`$ergs$`\text{gm}^1`$ mentioned above. The bound region on the other hand, is relatively under-dense and slow. The total initially bound mass is $`0.22M_{}`$. The expansion time scale throughout most of the initially bound material is $`t_05060`$hrs. This region is radiation-pressure dominated, and the initial sound speed is roughly $`c_{s,0}(\text{He})1.3\times 10^7`$cm$`\text{s}^1`$. The corresponding accretion time scale, $`t_{acc,0}50`$hrs is therefore similar to the initial expansion time scale. Unlike the analytic cases discussed above, in this model pressure forces are initially important. The long initial expansion time scale more than compensates for the decreased density in terms of determining the accretion rate, and we have $`t_{crit}300`$hrs. This model is also expected to reach very high accretion rates at early times, which will result in some moderation of the accretion flow when $`\dot{M}_{crit}3.13`$$`M_{}\text{yr}^1`$ is approached. #### 4.3.2 Computational Aspects Tracking the evolution of the envelope over several years (the expected time of emergence of the black hole) is not feasible numerically. Integrating the model until emergence would require a CPU time greater than one month, even when adjusting the position of the inner boundary and using the MTP procedure (see § 3). As a result, we devised a rescaling scheme where the original model is rescaled to a more rapidly evolving one, while maintaining simple relations between the quantities computed for the rescaled model and the original one. This rescaling scheme, presented in detail in the appendix, has allowed for an additional acceleration of a factor of $`5`$ in computational efficiency, thus reducing the total required computational time to an acceptable limit. We must note that several particulars of the initial model used here for SN1997D imposed some specific complications in employing the rescaling scheme. The fact that initially $`t_{acc}t_0`$ and that the flow is expected to be moderated at early times when a critical accretion rate is approached implies that the rescaling scheme described in the appendix cannot be applied at the earliest stages of evolution. Furthermore, at early times the energy generation rate of the radioactive elements, and especially $`{}_{}{}^{56}\mathrm{Ni}`$, is at its largest. The enhancement required for these rates in the course of rescaling (see in the appendix) further compounds the applicability of the rescaling scheme at very early times. Therefore our numerical approach has been to perform the simulation of the model in its original scale until three conditions are met: 1) the time dependent marginally bound radius is significantly smaller (a factor of 5) than the accretion radius; 2) the accretion rate through the inner boundary is smaller than $`0.1\dot{M}_{crit}`$ and 3) that total energy to be emitted in the decays of the remaining $`{}_{}{}^{56}\mathrm{Ni}`$is significantly smaller than the internal energy at that time. We find that these conditions are all satisfied at the physical time of about $`t=25`$days, which is when the simulation was stopped, rescaled by a factor of 5, and restarted. We performed the simulation of the rescaled model with a single time-step integration until the recombination front has settled to a roughly constant Lagrangian position, which occurs after about $`t=150`$days. These first two stages required about 20 CPU hours each. From this point on, the simulation is continued with the Multi Time-step Procedure, where the integration domain was divided into four sub-grids. Performing the simulation up to a physical time of $`t=1500`$days required an additional 120 CPU hours (recall that at later times the time step drops as the inner boundary must be moved inward in order to maintain it in LTE – see § 3). All computations were performed on a SGI 500MHz EV6 machine with a compaq XP1000 processor. #### 4.3.3 Numerical Results The calculated light curve for SN1997D is shown in Fig. 4.3.3. Also shown are the observed data extending up to $`440`$ days after the explosion (Benetti et al., 2000), with which the agreement of the simulated light curve is very good. For comparison, we also show the light curve calculated for an identical model but with the abundance of radioactive elements reduced by a factor of 300, and also the late-time light curve estimated analytically as in Fig. 4.2.4 above. In the simulation we cannot single-out the accretion luminosity directly, by we can evaluate its contribution to the total luminosity by subtracting the contributions of radioactive heating. The simulation does record the total rate of energy deposition in the envelope by the decays of each of the isotopes, $`Q_X(t)`$, and since we can estimate that the deposited heat is emitted as a bolometric luminosity very rapidly after thermalization, i.e., $`L_X(t)Q_X(t)`$, we can attempt to identify the accretion luminosity as $`L_{acc}=L_{tot}(Q_{{}_{}{}^{56}Co}+Q_{{}_{}{}^{57}Co}+Q_{{}_{}{}^{44}Ti})`$. These calculated contributions for the model are shown in Fig. 4.3.3. The transition from a $`{}_{}{}^{56}\mathrm{Co}`$-dominated tail in the luminosity is evident at about 1000 days. From Fig. 4.3.3 we can determine that the accretion luminosity does become the dominant source of the light curve, and the time of emergence (when $`L_{acc}(t_{BH})=\frac{1}{2}L_{tot}(t_{BH}))`$ is $`t_{BH}1050`$days, when $`L_{acc}3.2\times 10^{35}`$. This is roughly $`\frac{1}{5}`$ of the accretion luminosity (at the same physical time) estimated using the initial average values of $`\rho _{0,He}=5\times 10^6`$gm$`\text{cm}^3`$and $`t_0=50`$hrs in equation (18). This difference is the result of moderation by radiation forces as the accretion rate exceeds $`\dot{M}_{crit}`$, and in part also by the fact that initially the envelope’s own pressure is not negligible. An accretion rate of $`\dot{M}>\dot{M}_{crit}`$ is actually reached while the accretion flow is still building, and the radiation pressure it induces causes a significant part of the initially bound material to become unbound. Indeed, the the total accreted mass (extrapolated to $`t\mathrm{}`$) found in the simulation is decreased to only $`0.067M_{}`$, less than a third of the originally bound material. The evolution of the accretion rate is plotted in Fig. 4.3.3, which shows the accretion rate (calculated near the inner boundary) as a function of time. The maximum accretion rate is reached after about 18 hrs, when its value is $`1.5\dot{M}_{crit}`$. At this stage the accretion flow is not quasi-stationary (see below), and significant moderation is imposed. We comment that the details of the accretion rate at these very early times are dependent on the assumed initial conditions. With initial conditions set during collapse, MacFadyen, Woosley, & Heger (1999) find that the earliest accretion reaches a maximum rate as early as a few hundred seconds after bounce. In our numerical simulation the initial conditions are set after the passage of the shock through the envelope and hence we cannot reproduce these earliest stages of the accretion history. More significant in the context of black hole emergence, however, is the character of the flow as it settles on the late-time, dust-accretion track, where there is good agreement between our results and those of MacFadyen et al. (1999). It is apparent that the accretion does eventually settle on a dust-like flow $`\dot{M}t^{5/3}`$, as expected, at $`10`$ days after the explosion. The quantitative effect of the early history of the accretion flow is clearly apparent, since the accretion rate is lower than the value predicted by equation (16) for $`\dot{M}(10\text{days})`$, based on the initial properties of the inner part of the bound region. By extrapolating the dust-like accretion rate back in time, we find that $`\dot{M}=\dot{M}_{crit}`$ at $`100`$hrs, instead of the original estimate of $`t_{crit}300`$hrs. We mark the time of transition between the original model and the rescaled one by a circle in Fig. 4.3.3. After a short transient, the accretion flow resettles on a dust-like accretion flow, with a very good fit (accurate to about $`10\%`$) to the initial unscaled model. This agreement suggests that the rescaled model is indeed capable of reproducing the accretion flow and luminosity. In Fig. 4.3.3 we show the behavior of the computed accretion luminosity as a function of the computed accretion rate $`\dot{M}`$, and compare it to the analytic estimate of the Blondin formula (eq. ). The excellent fit of the simulation to the analytic ratio demonstrates that the radiation-hydrodynamic evolution in the accreting region does indeed follow a sequence of quasistationary states, so the analytic estimates are also justified in the case of a variable chemical composition. The reduced accretion rate is clearly the cause for the accretion luminosity being only about one fifth of its value (had the entire initially bound mass remained available for accretion). It is noteworthy that the black hole emerges somewhat earlier than in the case of Fig. 4.2.4, in spite of the lower accretion luminosity. This is due to a significant complementary effect: the structure of the helium rich layer reduces the effective optical depth to the high energy photons emitted in the radioactive decays. First, the initial column depth of the layer is only about $`\frac{1}{3}`$ of that of a layer with identical mass and size but a constant density. Second, the outer part of this layer has a larger specific kinetic energy and therefore a smaller $`\rho _0t_0^3`$ than the average value used in § 4.2. This combination causes the $`\gamma `$ray column depth of the helium-rich layer to be only 1/5–1/4 of that at an identical time for an analytic estimate, and the trapping efficiency is decreased accordingly. Correspondingly, the contribution of radioactive heating to the bolometric luminosity is reduced to what would have been expected if the entire helium-rich layer were characterized with $`\rho _0=5\times 10^6`$gm$`\text{cm}^3`$ and $`t_0=50`$hrs. The model with reduced radioactive isotopes was investigated in order to assess the quantitative effect of radioactive heating on the dynamics of the flow. This effects can be inferred indirectly from Fig. 4.3.3, and directly by comparing the accretion history in both models. We find that the accretion histories are practically identical, where the late-time accretion rate in the model with reduced radioactive heating is larger by only $`5\%`$ than in the nominal model, implying that even in the nominal model the energy deposited by heating hardly affects the region of bound material. The calculated accretion luminosities also agree to within a few percent, although it is the model with radioactive isotopes that suggests a higher luminosity. While this effect may be partially a numerical artifact, it may also arise from the slightly larger photo-spheric radius found in the original model, presumably caused by radioactive heating. We conclude that when the abundance of radioactive elements in the ejected envelope is as low as it is in the case of SN1997D, it is unlikely to impose a significant effect on the expected accretion rate or luminosity. By contrast, even the small amount of radioactive isotopes present in the SN1997D envelope is sufficient to affect the decline of the recombination peak (see Fig. 4.3.3). This effect was demonstrated clearly in the investigations of the SN1987A light curve (Woosley, 1988; Shigeyama & Nomoto, 1990), and is mainly due to release of heat deposited in the inner part of the envelope by $`{}_{}{}^{56}\mathrm{Ni}`$decays. The negligible effect of radioactive heating on the hydrodynamic evolution allows for a simple extrapolation regarding the expected light curve if the abundance of the relevant isotopes is varied. If the abundances of $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$are very low, emergence can be gauged by comparing the accretion luminosity and the heating by $`{}_{}{}^{56}\mathrm{Co}`$decays, suggesting that emergence occurs at about 970 days. The luminosity at emergence would be slightly higher, with $`L_{acc}3.9\times 10^{35}`$. In this case, the contribution of $`{}_{}{}^{56}\mathrm{Co}`$heating declines rapidly and the late time light curve should indeed fall off as a power-law in time. Note that a perfect power law decline is found, as expected, for the model with reduced abundances of radioactive isotopes. To conclude our investigation of the SN1997D light curve, we examine the history of the luminosity and the accretion flow throughout the envelope. The profiles of the (comoving frame) luminosity, $`L(r)`$, and mass flux, $`|\dot{M}|(r)`$, at various selected times during this evolution are plotted in Figs. 4.3.3a and 4.3.3b respectively. The luminosity profiles demonstrate the evolution of the light curve through its various stages (see Zampieri et al. 1998a for details). During the stage dominated by radioactive heating the photo-sphere lies well above the accreting region. The presence of a separate source of luminosity due to compressional heating in the course of accretion is evident in the inner part of the flow, but only when its magnitude becomes comparable to the heating produced by radioactive decays does it start to dominate the luminosity output. The drop in luminosity near the inner boundary arises because the opacity of the partially ionized helium-rich material becomes larger (much greater than the electron scattering opacity). Closer to the black hole the increased temperature increases the degree of ionization of the helium-rich material, and the comoving luminosity rises again to a local maxima. Note that the mass flux remains self similar through out the evolution, as the accreting region grows in size but the expansion of the envelope reduces the magnitude of the accretion rate. The sharper, inner spike in the mass flow rate is due to the original spike in density of the model at the interface between the helium-rich and hydrogen-rich layers (the smoother, outer peak simply corresponds to a maximum in the combination $`\rho vr^2`$). In the inner region the accretion rate is nearly independent of radius, which also reflects that the accretion flow does indeed follow a sequence of quasi–stationary states. ## 5. Black Hole Emergence in other Supernovae While SN1997D appears to be the only existing candidate for observing black hole emergence, theory suggests that the explosions of the most massive stars in the core-collapse supernovae range would be significantly more favorable for an unequivocal determination of the presence of a black hole. The key feature is, of course, the final yield of heavy elements in the envelope following early fallback. These heavy elements include the radioactive isotopes which mask the accretion luminosity for extended times, so that the lower the abundance of radioactive elements, the shorter the time until emergence and the larger the luminosity at that time. SN1997D may be sufficiently poor in heavy elements so that the accretion luminosity can dominate the heating by $`{}_{}{}^{44}\mathrm{Ti}`$decays, but $`{}_{}{}^{56}\mathrm{Co}`$decays are by far the dominant source of power until $`1000`$days after the explosion. In a survey of the supernovae of massive stars, Woosley & Weaver (1995) concluded that explosions of stars of initial masses of $`3040M_{}`$ are likely to include significant early fallback, thus advecting the newly synthesized heavy elements onto the remnant, a likely black hole. For “nominal” explosion energies of $`1.2\times 10^{51}`$ergs, they found that explosions of stars with masses of $`3040M_{}`$ should easily lead to remnant masses in the range of $`310M_{}`$, while ejecting envelopes practically free of heavy radioactive elements. While they did not consider explosion energies as low as the $`4\times 10^{50}`$ergs inferred for SN1997D, it is natural to assume that lower energy explosions will give rise to enhanced earlier fallback and thus higher remnant masses<sup>5</sup><sup>5</sup>5Woosley & Weaver (1995) did examine explosion energies of $`2\times 10^{51}`$ergs. In general, they found that if explosion energy is large enough to eject a non-negligible amount of radioactive isotopes in stars of $`3040M_{}`$, the remnant mass will be $`2.5M_{}`$, and hence, possibly, a neutron star. In order to gauge the possible signature of black hole formation in a supernova of a very massive progenitor, we computed the light curve of such a supernova using the numerical code described in § 3. As a reference we used a $`35M_{}`$ progenitor based on model S35A of Woosley & Weaver (1995). It has a pre-explosion outer radius of $`8\times 10^{13}`$cm, and is composed of a $`14M_{}`$ helium-rich core and a $`21M_{}`$ hydrogen-rich envelope. The post explosion ejecta profile was constructed so that the total ejected masses are 11.5, 12 and 4 $`M_{}`$ of hydrogen, helium and oxygen, respectively. Most notably, the envelope is assumed to be free of radioactive isotopes. The total kinetic energy of the ejecta is set at $`1.2\times 10^{51}`$ergs, leaving behind a $`7.5M_{}`$ black hole. Assuming that the helium-rich layer carries $`1\%`$ of the total energy and that the entire envelope has a common expansion time scale, our model has $`t_055`$hrs. The initial profile of the bound material thus has $`\rho _{0,He}=4.93\times 10^6`$gm$`\text{cm}^3`$and therefore $`\rho _0t_0^3=3.83\times 10^{10}`$gm$`\text{cm}^3`$$`\text{s}^1`$. This expansion time scale is significantly shorter than the accretion time scale, which for this model is roughly $`10`$days, so that the flow is dust-like at its onset. The critical time is, however, $`280`$hrs, and once again we expect the accretion can begin as dust-like and become then moderated by radiation pressure when the Eddington limit is approached. Through the analytic estimates in § 4.2 we find that $`t_{dust}110`$hrs for this model. The calculated light curve for S35A is shown in Fig. 5. We note that no rescaling was required for these models due to the relatively large time step and short evolutionary time required until emergence; the calculation required about 55 CPU hours. The modest explosion energy combined with the large radius of the progenitor lead to a relatively small initial average temperature of about $`2.5\times 10^5`$$`{}_{}{}^{}\mathrm{K}`$, and recombination occurs around a time of $`t_{rec}35`$days. The light curve during this peak is quite similar to the one observed for SN1994W, a reasonable candidate for an explosion of a very massive star with very little radioactive elements in the ejecta (see below). In the absence of radioactive heating the recombination peak fades rapidly, and the emergence of the black hole occurs at $`t55`$days with a total luminosity at emergence of about $`10^{37}`$ergs$`\text{s}^1`$. The rapid decline of the recombination luminosity causes the accretion luminosity to be nearly the sole contributor to the light curve as early as at $`t60`$days, with a luminosity of $`L(t=60\text{days})7.2\times 10^{36}`$ergs$`\text{s}^1`$. This value is smaller than the analytic result of $`L(t)=L_{Edd}(60\text{days}/t_{dust})^{25/18}`$ by a factor of a few, although the calculated accretion rate does satisfy $`\dot{M}(t)\dot{M}_{crit}(t/t_{dust})^{5/3}`$. This result is due to the fact that at this stage bound-bound and bound-free opacities in the (relatively narrow) partially ionized region in the accreting material are high (much larger than free electron scattering opacity). This imposes a significant contribution to the total optical depth of the accretion flow, increasing its “effective” average opacity. Note that for a given accretion rate, the accretion luminosity is expected to follow a $`L_{acc}\kappa ^{1/2}`$ dependence (using eq. and expressing $`L_{Edd}`$ and $`\dot{M}_{Edd}`$ explicitly). This is a transient effect: as time elapses, expansion of the envelope gradually causes the fully ionized, inner part of the accreting region to dominate the optical depth of the flow, so that the average “effective” opacity is rapidly declining towards the electron scattering value. Consequently, at this stage the luminosity falls off somewhat more slowly than a $`t^{25/18}`$ power-law, eventually settling on such a time dependence. Also shown in Fig. 5 is the same model with the black hole mass arbitrarily increased to $`M_{BH}=15M_{}`$, the light curve evolution is identical until emergence, which occurs at a slightly higher luminosity and therefore at a slightly earlier time ($`3\times 10^{37}`$ergs$`\text{s}^1`$at $`t50`$days). The light curve again evolves thereafter due to accretion $`L(60\text{days})1.33\times 10^{37}`$ergs$`\text{s}^1`$, again falling off somewhat more moderately than a $`t^{25/18}`$ power law. We emphasize that the accretion rate and luminosity in both models are significantly lower than that predicted by equations (16, 19) for the initial parameters of the helium-rich layer. Even though the accretion can begin as dust-like, the moderation of the accretion flow by radiation pressure when the Eddington limit is approached is evident in the absolute magnitude of the accretion luminosity-tail. The model with the larger black hole mass has a higher critical accretion rate, so even though it also induces a smaller $`t_{dust}`$, the ratio of the accretion luminosity is increased by slightly more than just a factor of $`2^{2/3}`$ arising from the ratio of the black hole masses. We conclude, that Type II supernovae with no radioactive isotopes provide the most favorable cases for observing the emergence of a black hole. It is noteworthy that a similar explosion of the bare helium core in Type Ib/Ic supernovae would be far less favorable. Not only will all the explosion energy be deposited in the helium layer (instead of only a few percent), there will also be no inverse shock that slows the inner part of the helium layer and enhances fallback (which is essential for decreasing the abundance of ejected radioactive elements). The very small time step required for simulating such a model prevents us from examining it numerically, but we can place some limits with the analytic derivations presented in § 4.2. For example, consider a $`14M_{}`$ helium star with an initial radius of $`R_010^{11}`$cm exploding with an energy $`1.2\times 10^{51}`$ergs, producing, again, a $`7.5M_{}`$ black hole. We estimate that the post-explosion envelope will have $`\rho _{0,He}t_0^31.85\times 10^7`$ergs$`\text{gm}^1`$, $`\rho _03`$gm$`\text{cm}^3`$and $`t_00.05`$hrs. Even assuming that the flow does begin as dust-like, we find that $`t_{dust}4`$hrs ($`t_{crit}10`$hrs) so that the luminosity at $`t60`$days will be only $`5.3\times 10^{35}`$ergs$`\text{s}^1`$\- a few percent of the values found above. While recombination is likely to occur earlier in such supernova, the actual accretion rate in such an explosion will be strongly modified due to high initial temperatures (and pressure), confirming that even radioactive-free Type Ib/Ic supernovae are unlikely to be good candidates for observing the emergence of black holes. ## 6. Conclusions, Discussion and Observational Prospects In this work we have examined the emergence of luminosity due to spherical accretion onto black holes formed in realistic supernovae. The study is based on a combination of analytic estimates and numerical simulations. Our main focus has been on SN1997D, where the low explosion energy and very low inferred abundance of $`{}_{}{}^{56}\mathrm{Co}`$in its ejecta provide a possible (and currently, the only) candidate where the emergence of a black hole in a supernova may actually be observable. We confirmed the main features of the accretion history and luminosity as derived by Zampieri et al. (1998a). At late times, the accretion flow settles on a dust-like motion, and the accretion rate falls off as a $`t^{5/3}`$ in time (Colpi et al., 1996). Accretion proceeds in a sequence of quasistationary states, so the accretion luminosity achieves the expected magnitude of a hypercritical, spherical accretion flow (Blondin, 1986), accounting for the secular decay in time of the accretion rate. The resulting time dependence of the late-time accretion luminosity is $`L(t)t^{25/18}`$. This relation marks a fundamental feature regarding emergence: since the main competitor to accretion in powering the late time light curve is radioactive heating which decays exponentially with time, emergence of the accretion luminosity is inevitable if spherical accretion does indeed persist. While in typical Type II supernovae the abundance of radioactive isotopes is expected to be large enough to prevent a practical observation of the emergence, more rare explosions where this abundance is significantly reduced and fallback is enhanced provide important exceptions. The case of SN1997D offers a unique opportunity to examine black hole emergence (Zampieri et al., 1998b), in view of its inferred explosion energy of $`4\times 10^{50}`$ergs$`\text{s}^1`$ and very low abundance of $`{}_{}{}^{56}\mathrm{Co}`$, i.e., only $`2\times 10^3M_{}`$ (Turatto et al., 1998; Benetti et al., 2000). We have have examined the expected emergence of a black hole in SN1997D analytically and numerically. In the analytic study we noted that the early-time accretion is almost certain to generate an Eddington-rate luminosity, which will modify the accretion history and decrease the late-time accretion rate. In the numerical investigation we have included the effects of a variable envelope composition. In particular, we incorporated heating due to radioactive decays, including a finite (time-dependent) $`\gamma `$ray optical depth. The wide diversity of time scales involved compelled us to use a rescaling scheme in the numerical simulations (see in the appendix). Our various analyses offer a consistent assessment of the emergence of the black hole in SN1997D. We confirm the preliminary estimate of Zampieri et al. (1998b) that a $`3M_{}`$ should emerge in SN1997D about $`1000`$days after the explosion, in contrast to hundreds of years in the case of “standard” Type II supernovae (like SN1987A) where the abundances of radioactive elements are significantly higher. We show that the dominant parameter in determining the luminosity at emergence is the amount of kinetic energy carried in the post-explosion helium-rich layer (which is the source of late-time accretion). For realistic values of this energy, we estimate that the total luminosity at emergence will lie in the range $`5\times 10^{35}L_{tot}3\times 10^{36}`$ergs$`\text{s}^1`$, emerging at $`10001500`$days after the explosion. In our numerical study we have examined the best fit model of Turatto et al. (1998) for SN1997D. The simulations suggest that, for this particular model, the black hole emerges at about $`1050`$days after the explosion, with $`L_{tot}6.5\times 10^{35}`$ergs$`\text{s}^1`$. The results of the simulation, which is based on a consistent, fully-relativistic, radiation-hydrodynamics code, provides significant support to the analytic approach. Most notably, it demonstrates that, after passing through an early radiation-pressure-limited stage which reduces the accretion rate, the accretion flow does settle on the dust-like solution and that the influence of the radioactive decays on the hydrodynamic evolution is negligible for the low abundances inferred in SN1997D. The time of emergence and the corresponding luminosity depend on the abundances of $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$in the envelope, which are unknown a-priori. However, as long as the accretion luminosity is significantly larger than the threshold set by positron emission in $`{}_{}{}^{44}\mathrm{Ti}`$decays, both emergence time and luminosity will not vary substantially. At the time of emergence $`\gamma `$ray escape from the envelope is non-negligible and increases rather rapidly with time, as the envelope’s optical depth drops. Correspondingly, the bolometric luminosity due to radioactive heating decreases even faster than exponentially with time. Although our results appear to be quite robust within the context of the present study, some uncertainties arise due to effects not considered here. Most critical is the assumption of perfectly spherical accretion: if the bound material can maintain sufficient angular momentum, a centrifugal barrier would drive the flow into a disk structure (Chevalier, 1996). The radiative efficiency of hypercritical disk accretion is uncertain, complicating any a-priori estimate regarding the character of the late-time accretion luminosity. The residual angular momentum is likely to be dependent on the details of the progenitor and the explosion (we note that for SN1987A the observed neutrino burst seemed consistent with nonrotating models (Burrows, 1988), indicating that angular momentum of matter close to the collapsed core did not play a significant role). In principle, late time accretion occurs many viscous time scales after the explosion, and much of the initial angular momentum could be lost, but the minimum specific angular momentum required to force disk accretion is quite small (about $`(GM_{BH}R_{ISCO}c)^{1/2}`$, where $`R_{ISCO}=6GM/c^2`$ is the innermost-stable circular orbit near a black hole). Furthermore, even if at late times angular momentum is negligible, some residual effect may arise if the character of the early-time accretion has been sufficiently altered (Mineshige et al. 1997 found through simulations that sufficient angular momentum strongly affects the magnitude and time dependence of the accretion during the first hours after collapse). We defer further investigation of fallback including angular momentum to future work. Additional effects could arise due to convection, mixing and clumping of envelope material (and radioactive isotopes in particular), and dust extinction, although we expect that at a time of $`3`$years after the explosion, any dust formed will have become transparent to the underlying luminosity. This may not be true if photoionization by the central accreting black hole (not included in our calculation) is important. Our results illuminate the uniqueness of SN1997D. It is the low abundance of radioactive elements, combined with the low explosion energy in general, that allows the black hole to emerge only a few years after the explosion, at which time the absolute magnitude of the accretion luminosity is still quite large (recall that for “standard” explosions we expect emergence luminosities of $`10^{32}10^{33}`$ergs$`\text{s}^1`$). Since gas and radiation are in LTE in the inner accreting region and the photosphere is well localized in radius, the emerging spectrum is expected to be very similar to a blackbody at the photospheric temperature ($`7500`$ K) with the superposition of heavy line elements. We therefore expect that emission will be mainly in the optical band. At a distance of $`14`$Mpc, the corresponding apparent magnitude for the range of emergence luminosities predicted for SN1997D is $`m_V2830`$ and $`m_Rm_I2729`$. The higher luminosity (lower magnitude) end of this range coincides with the detection threshold of the HST STIS camera, indeed making emergence in SN1997D potentially observable. We note that the capabilities of HST are needed both because of the faintness of the source at this time, and the resolution, required to distinguish the supernova over the background of the parent galaxy NGC 1536. A recent estimate (M. Turatto 1999, private communication) suggests that the required exposure time for the HST STIS to resolve the apparent magnitude expected at emergence in SN1997D at a signal to noise (S/N) ratio of 10 would be $`25000`$seconds. The accretion luminosity should be distinguishable from radioactive heating (or from circumstellar interaction, which has not been observed in SN1997D so far, Benetti et al. 2000) due to its unique power law dependence. This would be especially true if the abundances of $`{}_{}{}^{57}\mathrm{Co}`$and $`{}_{}{}^{44}\mathrm{Ti}`$are negligible, since after emergence the luminosity would be due only to accretion. But, even if there is a finite contribution from these two isotopes to the light curve, the presence of a power-law source may still be identifiable if several measurements with reasonable accuracy could be made. The novel consequences of a positive observation are self-evident - the first direct observational evidence that a black hole can be formed by an otherwise “successful” supernova. Although our main focus has been SN1997D, other theoretical Type II supernova models offer more favorable conditions for an actual observation of black hole emergence. The survey of Woosley & Weaver (1995) suggests that progenitors with masses in range $`3040M_{}`$ are quite likely to leave behind a remnant of mass $`310M_{}`$ and a practically radioactive-free envelope, as all the iron group isotopes are engulfed by the early fallback. Our calculations in § 5 show that accretion luminosity will emerge after recombination depletes the initial internal energy of the envelope, hence after a few tens of days. The corresponding luminosity at emergence will be $`10^{37}`$ergs$`\text{s}^1`$, which with the present capabilities of HST would allow for detection at emergence perhaps out to $`2025`$Mpc. The luminosity at emergence is higher than for SN1997D so that the galactic background would be less important, and ground based instruments might be relevant as well. How often can we expect an opportunity to observe a black hole emerging in a supernova? An upper limit is, of course, the rate at which black hole forming supernovae occur. Adopting the recent results of Fryer & Kalogera (1999) that about $`10\%`$ of core collapse supernovae make black holes, we expect a rate of about one event per thousand years per galaxy. There are somewhat less than one thousand galaxies at a distance $`20`$Mpc from the Milky way, and hence there may be approximately one observable event (equal or better than SN1997D) per several years. This estimate is an upper limit, since a significant fraction of high mass progenitors will end their lives as Type Ib/Ic supernovae after losing their hydrogen envelope, and such supernovae are unfavorable for a practical observation at emergence. Furthermore, even “perfect” candidates may not allow for an actual observation, as seems to have been the case in SN1994W. Following its recombination peak, the tail of the SN1994W light curve failed to settle on an exponential decay that would arise due to $`{}_{}{}^{56}\mathrm{Co}`$decay (Sollerman et al., 1998). Interestingly, the last data point led to a limit of $`2\times 10^3`$ on the mass of $`{}_{}{}^{56}\mathrm{Co}`$in the ejected envelope. However, the light curve showed a very steep power-law in time until it dropped below the detection threshold some 220 days after the explosion. The luminosity during this phase was much too high to be due to accretion ($`10^{40}10^{42}`$ergs$`\text{s}^1`$) and Sollerman et al. (1998) found that it was most likely caused by interaction with circumstellar material, with dust formation also playing an important role. We can only speculate that in the absence of circumstellar interaction and dust extinction, the light curve would have declined exponentially after the recombination peak and the accretion luminosity may have emerged while still marginally detectable. The case of SN1994W implies that the rate of potential observable black hole emergence may be significantly smaller than that of black hole forming Type II supernovae. Nonetheless, we urge that any Type II supernovae which shows a diminished $`{}_{}{}^{56}\mathrm{Co}`$abundance should be monitored, and its light curve be tracked for several months after the explosion. The prospects of observing black hole emergence may improve significantly with potential future instruments, such as NGST or a dedicated faint-supernovae project. We are grateful to Tim Young for providing us with the results of early-time simulations for SN1997D and to Chris Fryer, Robert Kirshner, Phillip Pinto and Massimo Turatto for useful discussions. We thank Monica Colpi and Roberto Turolla for carefully reading the manuscript. This work was supported by NSF Grants AST 96-18524 and PHY 99-02833, and NASA Grants NAG 5-7152 and NAG 5-8418 at the University of Illinois at Urbana-Champaign. L. Z. acknowledges finacial support from the Italian Ministry for University and Scientific Research (MURST) under grant cofin–98–2.11.03.07 at the University of Padova. ## Appendix A The Rescaling Scheme: Radiation-Hydrodynamic Renormalization The inclusion of the accreting region imposes an inherent obstacle in any numerical simulation of a supernova light curve. In the innermost accreting region the material is free-falling at a velocity which is a significant fraction of the speed of light, so that the dynamical time scale is less than $`10^3`$sec. On the other hand, the evolutionary time for the light curve is of the order of tens of days to clear the recombination phase, and years to reach black hole emergence if there is a non-negligible amount of radioactive elements in the envelope. This extreme diversity of time scales prevents any single calculation which tracks the entire envelope. As described by Zampieri et al. (1998a), a considerable increase in the shortest physical time scale (and, hence, the Courant integration time step) can be obtained moving the inner boundary, $`R_{in}`$, outward (at hundreds to thousands of Schwarzschild radii, but always well within the accreting region). This turns out to be possible because the gas in the inner accreting region is in LTE and in free-fall, thus allowing a reliable setting of the boundary conditions even if $`R_{in}`$ is much larger than the black hole horizon. When a Lagrangian zone passes through the inner boundary, it is removed from the calculation and the remainder of the envelope quantities are rezoned to maintain a satisfactory resolution. Even with such a relocation of the inner boundary, the feasibility of a numerical computation requires some additional measures to accelerate the computation. To this end, the code incorporates the Multi-Time-Step Procedure, where the integration domain is separated into several subgrids, and each is evolved at its own local time scale; most of the CPU time is thus devoted to evolving the regions with smaller time steps. We note that this procedure requires a communication scheme between the sub-grids based on extrapolation, and can be safely used when there are no sharp features propagating through the envelope. Thus, we use it only after that the recombination front has completed its sweep of the envelope material. The speed-up allowed by the MTP is typically a factor of 5-8. Unfortunately, moving the inner boundary outward and using the MTP is not sufficient when the target physical time is of the order of years<sup>6</sup><sup>6</sup>6We note that using a fully implicit code would not allow a significant gain in computational time because limitations on the time-step imposed by accuracy are comparable to limitations imposed by the Courant condition throughout most of the evolution.. In the case of SN1997D, tracking the evolution of the envelope for $`1000`$days would require roughly 1000 CPU hours on the SGI 500MHz EV6 compaq XP1000 processor which we have used for the simulations. We were hence compelled to develop an additional acceleration algorithm, which is a rescaling scheme that allows us to integrate to the black hole emergence stage within a reasonable computational time. ### A.1. Principles of Rescaling The main objective in devising a rescaled model of a specific supernova evolution is to maintain all the quantitative features of the realistic model, while reducing the required computational time. The code uses an explicit Lagrangian scheme. Since the photon transport is treated by actually solving the transport equation, the reference velocity is the speed of light, and the Courant time step is determined by the width of the thinnest zone. This is always the innermost zone, $`\mathrm{\Delta }R_{in}(t)`$, which at any time $`t`$ is a given fraction of the position of the inner boundary, $`R_{in}(t)`$. The efficiency of the code is essentially determined by comparing the (small) ratio of the time step to a characteristic evolutionary time – for example, the expansion time scale, $`t_0`$ $$\frac{\mathrm{\Delta }R_{in}(t)/c}{t_0}\frac{R_{in}(t)/c}{t_0}.$$ (A1) The essence of a rescaling scheme would then be devising a rescaled model where $$\frac{R_{in}^s(t^s)/c}{t_0^s}=\alpha \frac{R_{in}(t)/c}{t_0},$$ (A2) where a subscript $`s`$ denotes the rescaled quantities. The rescaled model is then more efficient by a factor of $`\alpha >1`$. Successful rescaling must (a) provide a simple relation between its own physical quantities and that of the realistic model and (b) maintain the hierarchy (inequalities) of time and energy scales that exists in the realistic model, so that the relative importance of the different physical process is carried on into the rescaled model. The rescaling scheme we have chosen to employ in our simulations is detailed below. ### A.2. The Rescaling Relations A natural rescaling scheme would be one where all the physical quantities of a mass element ($`r,v,T,\rho \mathrm{}`$) of the realistic model are mapped with simple linear relations to the rescaled model, i.e., $`r^s=\alpha _Rr,v^s=\alpha _Vv`$, etc. In our rescaling scheme we distinguish between the local thermodynamic and kinematic quantities of the mass elements in the envelope. Since the photon opacities are sensitive and complex functions of the thermodynamic quantities , i.e., density and temperature, we find it preferable to leave these quantities unchanged in the course of rescaling $`T^s=T,\rho ^s=\rho `$, which leads to $`c_s^s=c_s,\kappa ^s=\kappa `$. Note that such a choice also implies that the inner boundary of the realistic model should approximately map onto the inner boundary in the rescaled model, since the position of the inner boundary is determined by the requirement of LTE, which arises from the local values of temperature and density. Once this choice has been made, the constraints on the rescaling scheme are quite straight forward. If we require that the ratio of diffusion and recombination times scales to the expansion time scale remain unchanged, i.e, $$\frac{t_{diff}^s}{t_0^s}=\frac{t_{diff}}{t_0},\frac{t_{rec}^s}{t_0^s}=\frac{t_{rec}}{t_0},$$ (A3) the rescaled model must satisfy (see eqs. ): $$R_0^sV_0^s=R_0V_0.$$ (A4) The condition (A4) is sufficient to determine the rescaling strategy when consistently mapping all radii between the realistic and the rescaled models. If all radii are decreased by a factor of $`\alpha `$, all velocities must be increased by the same factor; the efficiency of the simulation (eq. \[A1)\]) is then increased by a factor of $`\alpha ^1/\alpha ^2=\alpha `$. If we further impose invariant luminosity ratios according to $$\frac{L_{diff,0}^s}{L_{acc,0}^s}=\frac{L_{diff,0}}{L_{acc,0}},\frac{L_{rec}^s}{L_{acc,0}^s}=\frac{L_{rec}}{L_{acc,0}},$$ (A5) we arrive (using eqs. \[3,7 and 19\]) at the condition $$M_{BH}^sR_0^s=M_{BH}R_0.$$ (A6) The principle relations of our rescaling scheme are thus $$\left(\frac{R_0^s}{R_0}\right)^1=\left(\frac{V_0^s}{V_0}\right)=\left(\frac{M_{BH}^s}{M_{BH}}\right)=\alpha .$$ (A7) This choice of rescaling allows us to consistently map the marginally bound shell of the envelope, since the marginally bound radius, $`R_{mb}`$ also rescales as $`R_{mb,0}^s=(2M_{BH}^s(t_0^s)^2)^{1/3}=\alpha ^1R_{mb,0}`$. Further relations that immediately follow are $$\left(\frac{t^s}{t}\right)=\alpha ^2,\left(\frac{L_{acc,0}^s}{L_{acc,0}}\right)=\alpha ^1,\left(\frac{M_{end}^s}{M_{env}}\right)=\alpha ^3,\left(\frac{\dot{M}^s}{\dot{M}}\right)=\alpha ^1.$$ (A8) The efficiency of rescaling is explicit in the rescaling of time – in the rescaled model all physical times are reduced by a factor of $`\alpha ^2`$, while the numerical time step is only reduced by a factor of $`R_{in}^s/R_{in}=\alpha `$. ### A.3. Limitations of Rescaling The rescaling scheme presented above is not without limitations; most notably, the acceleration factor, $`\alpha `$ is limited to a value of a few, as discussed below. Nonetheless, even a value of $`\alpha =5`$, which was used in the case of SN1997D reduces the required computational time from weeks to days, hence allowing for a practical numerical investigation. #### A.3.1 The Acceleration Factor Since the rescaling scheme calls for an increase of a factor of $`\alpha `$ of the envelope velocities, clearly $`\alpha `$ cannot be arbitrarily large, since we must maintain a sub-relativistic envelope. Furthermore, since the black hole mass is increased, its Schwarzschild radius increases as well, and naturally it should not be allowed to approach the integration domain throughout the calculation. Both these constraints result in the practical limit on the value of $`\alpha `$ to no more than a few. #### A.3.2 Radioactive Elements The scaling laws of luminosity and mass in equation (A8) create an undesired inconsistency regarding the luminosity from radioactive heating, since for this source $`LM`$. In order to recover the correct rescaled power output from radioactive decays, the energy generation rate from radioactive decays per unit mass in the envelope must therefore be increase by a factor of $`\alpha ^2`$. This could potentially impose a significant deviation in the evolution of the rescaled model from that of the original one, if it artificially enhances the importance of energy deposition from radioactive decays when compared to the other energy scales. However, since we limit ourselves to envelopes where the abundance of radioactive elements is very small, the artificial amplification of the energy production per unit mass of radioactive decays has a negligible effect. It is further noteworthy that the effective transparency of the photons to the $`\gamma `$-ray photons emitted in the decays must be adjusted in order to recover the appropriate scaling of the $`\gamma `$transparency time. The simple form of $`\gamma `$-ray opacity (eq. ) allows to achieve consistency with the change $`\kappa _\gamma ^s=\alpha \kappa _\gamma `$, so that $`t_{trans,\gamma }^s/t_0^s=t_{trans,\gamma }/t_0`$. A similar transformation for the optical photons would be more complex, but can be neglected since the onset of thermal-photon transparency is imposed by recombination (rather than expansion), and is chiefly determined by the thermodynamic variables $`\rho `$ and $`T`$, which are kept invariant. #### A.3.3 The Accretion Time Scale and Radius A second inconsistency that arises in this particular rescaling scheme concerns the initial accretion time and radius, $`t_{acc,0},R_{acc,0}`$. As is evident from equations (13, 15), when the thermodynamic properties of the helium layer are not changed, so that the sound speed is conserved, $`t_{acc},R_{acc}M_{BH}`$. Our rescaling scheme actually increases these quantities by a factor of $`\alpha `$, while other characteristic times and radii are decreased by factors of $`\alpha ^2`$ and $`\alpha `$, respectively. This is an inevitable consequence of our choice of rescaling. This inconsistency is significant, however, only while the accretion flow is Bondi (1952)-like, when $`t_{acc}t_0`$ and $`R_{acc}R_{mb}`$ (Colpi et al., 1996). As the envelope continues to expand, the hierarchy of time scales and radii is eventually reversed and the accretion eventually transforms into a dust-like flow. Since the accretion time does not enter explicitly in any of the estimates for evolution and luminosity (see § 2), it is only this hierarchy which is of real consequence, and hence rescaling can be applied safely to models which are characterized by $`t_{acc}t_0`$, $`R_{acc}R_{mb}`$. #### A.3.4 Kinetic vs. Thermal Energies A similar analysis concerns the ratio of kinetic and thermal energies in the envelope. It is clear that the kinetic energy will scale as $`E_{kin}MV^2E_{kin}^s/E_{kin}=\alpha ^1`$, while the internal energy (neglecting radioactive heating) will scale as $`E_{th}R^3E_{th}^s/E_{th}=\alpha ^3`$, since the temperatures are unchanged. Rescaling thus amplifies the relative importance of kinetic energy over thermal energy in the expanding envelope. However, as is the case for the accretion and expansion time scales, the real issue is the hierarchy of energy scales. In fact, the assumption of homologous expansion is equivalent to assumption that the conversion of thermal energy to kinetic energy is negligible, i.e., that $`E_{kin}E_{th}`$, which is satisfied in the context of our rescaled models. ### A.4. A Test of Rescaling We demonstrate here the reliability of our rescaling scheme by studying a specific example. We assume a supernova envelope somewhat similar to the SN1997D case examined in § 4, with several simplifications that allow for a more rapid computation. The basic features of the model are presented in table 2 and in Figs. A.4-A.4. We note that the initial velocity profile throughout the entire envelope was set so that $`v(r)=r/t_0`$. The initial accretion time-scale is determined by the properties of the region of bound material. A single fiducial radioactive isotope was included, with a mass fraction which is 0.01 of that of oxygen. This isotope was assumed to emit all the decay energy in $`\gamma `$rays. We have carried out two simulations of this model, the first in its original form and once rescaled by a factor $`\alpha =5`$ (since $`t_0t_{acc,0}`$ and $`R_{mb,0}<R_{acc,0}`$ to begin with, the model can be rescaled at $`t=0`$). The light curves of both models are compared in Fig. A.4, where luminosity and time in the rescaled model were rescaled inversely in order to compare them to the results of the original model, i.e., $`L^s(t^s)\alpha L^s(\alpha ^2t^s),`$. We note that the rescaled model required only 60 CPU hours, compared to about 300 CPU hours for the original model to reach an equivalent evolutionary stage. The reliability of rescaling is evident from the very good agreement of the light curve and accretion rate history. In particular, the accretion luminosities at emergence agree to better than $`10\%`$ (and a similar agreement is found in the accretion rates). We note that there is an obvious deviation in the light curve during the recombination phase, which is not unexpected since recombination is a non-linear process (Arnett, 1996): the rescaled model has smaller physical size, and the recombination front crosses it more quickly. This discrepancy is not important for determining the late-time accretion rate and luminosity, since the accreting region is unaffected by the recombination occurring much further out in the expanding envelope.
warning/0004/quant-ph0004054.html
ar5iv
text
# Which Kind of Two-Particle States Can Be Teleported through a Three-Particle Quantum Channel? 11footnote 1Work supported in part by Istituto Nazionale di Fisica Nucleare, Sezione di Trieste, Italy ## 1 Introduction. The quantum teleportation process permits one to transmit unknown quantum states from a sender to a receiver which are spatially separated. The classical idea of teleportation, as portrayed in science fiction novels and movies, involves a complete dematerialization of an object positioned in place A and its reappearence at a distant place B. By contrast, quantum teleportation differs from this fanciful idea since it is only possible to teleport from A to B the state representing one or more particles in A, by transferring it to one or more particles already existing in B. One possible way of performing such a process consists in firstly learning all the properties of the original and unknown state, and then in transmitting them, by means of classical communication channels, to a receiver able to recreate a perfect copy. Nevertheless such a procedure would not work, for it would be necessary to have an infinite ensemble of identically prepared quantum systems in order to completely determine their original state. However one of the most striking features of Quantum Mechanics, i.e. the entanglement, supplies us with a quantum channel of communication being able to transfer an unknown quantum state from a place to another, without knowing it and without violating special relativity constraints (i.e., not instantaneously). Following the original proposal of Bennett et al. , to teleport a single-particle state the sender, traditionally named Alice, and the receiver, Bob, need only to share a particular two-particle maximally entangled state (an EPR singlet state) which acts as a purely quantum channel, and a telephone, the classical channel. The core of the teleportation process resides on a projective local measurement, performed by Alice, in the Bell basis consisting of four orthonormal and maximally entangled two-particle states, involving the unknown one. After sending the result of the measurement by the classical channel ( the crucial step which prevents from sending faster-than light messages ) Bob can reconstruct, applying the unitary and local transformations suggested by Alice, a perfect copy of the original state. At the end the unknown state of Alice is destroyed, respecting in a such way the no-cloning theorem . Bennett et al. applied this type of teleportation to single-particle states suggesting however that it should have been working also for arbitrary $`N`$-particle entangled states. As a matter of fact, it works through the use of $`N`$ two-particle quantum channels. In the present work we explore the possibility of obtaining the same result using only one quantum channel, realized by peculiar kinds of three-particle states, shared by Alice and Bob: the sender possesses one of the particles while the remaining two, belonging to the receiver, will be used as building blocks for the copy of the teleported state. We note that three-particle states have been obtained experimentally and could be used in the present context . As it will appear, only some kinds of two-particle entangled states, and not the most general one, can be teleported by means of three-particle quantum channels and measurements onto two-particle Bell entangled states. In appendix A all the different ways to achieve permitted teleportation are enumerated. ## 2 Teleportation of an arbitrary state Let us suppose that Alice wants to teleport to Bob, spacelike separated from her, an unknown and arbitrary two-particle entangled state : $$\{\begin{array}{cc}|\psi _{12}=\alpha |00_{12}+\beta |10_{12}+\delta |01_{12}+\gamma |11_{12}\hfill & \\ |\alpha |^2+|\beta |^2+|\delta |^2+|\gamma |^2=1\hfill & \end{array}$$ (2.1) We are following the Quantum Information Theory convention of indicating with $`\{|0,|1\}`$ a complete orthonormal basis for a two dimensional, single particle, Hilbert space, made up of eigenstates of operator $`\sigma _z`$ with eigenvalues respectively $`+1`$ and $`1`$. In order to teleport a two-particle state, the simplest single quantum-channel which can be used must be a three-particle one : $$|\varphi _{345}=\frac{1}{\sqrt{N}}[a|000+b|100+c|010+d|001+e|110+f|101+g|011+h|111]_{345}$$ (2.2) with $`a,b\mathrm{}h=\pm 1`$ or $`0`$, and $`N`$ equal to the number of coefficients which are not zeros. The initial state we start from, will be the product of the two states just defined : $$|\mathrm{\Omega }=|\psi _{12}|\varphi _{345}$$ (2.3) where particles labelled $`1,2`$ and $`3`$ belong to Alice ( she can perform local measurements only on them ), while particles labelled $`4`$ and $`5`$ are placed near Bob and they will constitute the bricks to make a copy of the state $`|\psi `$. The core of the teleportation mechanism resides in performing all the direct products between states in $`|\psi _{12}`$ and $`|\varphi _{345}`$ and in expressing subsequently the Alice’s states of particles $`2`$ and $`3`$ in terms of vectors of the Bell basis. This complete basis consists of four orthonormal maximally entangled states, simultaneous eigenvectors of the two commuting operators, $`(\sigma _x)_2(\sigma _x)_3`$ and $`(\sigma _z)_2(\sigma _z)_3`$: $$|\varphi ^+_{23}=\frac{1}{\sqrt{2}}\left(|00+|11\right)_{23}|\psi ^+_{23}=\frac{1}{\sqrt{2}}\left(|01+|10\right)_{23}$$ (2.4) $$|\varphi ^{}_{23}=\frac{1}{\sqrt{2}}\left(|00|11\right)_{23}|\psi ^{}_{23}=\frac{1}{\sqrt{2}}\left(|01|10\right)_{23}$$ Inserting expressions (2.4) into equation (2.3) the resulting state of the system – completely equivalent to the original one since no physical process has taken place – is : $`|\mathrm{\Omega }`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2N}}}|\varphi ^+_{23}\{|0_1[(\alpha a+\delta b)|00+(\alpha c+\delta e)|10+(\alpha d+\delta f)|01+(\alpha g+\delta h)|11]_{45}+`$ $`+|1_1[(\beta a+\gamma b)|00+(\beta c+\gamma e)|10+(\beta d+\gamma f)|01+(\beta g+\gamma h)|11]_{45}\}+`$ $`{\displaystyle \frac{1}{\sqrt{2N}}}|\varphi ^{}_{23}\{|0_1[(\alpha a\delta b)|00+(\alpha c\delta e)|10+(\alpha d\delta f)|01+(\alpha g\delta h)|11]_{45}+`$ $`+|1_1[(\beta a\gamma b)|00+(\beta c\gamma e)|10+(\beta d\gamma f)|01+(\beta g\gamma h)|11]_{45}\}+`$ $`{\displaystyle \frac{1}{\sqrt{2N}}}|\psi ^+_{23}\{|0_1[(\alpha b+\delta a)|00+(\alpha e+\delta c)|10+(\alpha f+\delta d)|01+(\alpha h+\delta g)|11]_{45}+`$ $`+|1_1[(\beta b+\gamma a)|00+(\beta e+\gamma c)|10+(\beta f+\gamma d)|01+(\beta h+\gamma g)|11]_{45}\}+`$ $`{\displaystyle \frac{1}{\sqrt{2N}}}|\psi ^{}_{23}\{|0_1[(\alpha b\delta a)|00+(\alpha e\delta c)|10+(\alpha f\delta d)|01+(\alpha h\delta g)|11]_{45}+`$ $`+|1_1[(\beta b\gamma a)|00+(\beta e\gamma c)|10+(\beta f\gamma d)|01+(\beta h\gamma g)|11]_{45}\}`$ Alice’s strategy to pursue the teleportation process, as already said, will consist in a local projective measurement onto the vectors of the Bell basis of particles $`2`$ and $`3`$ and in a successive measurement on particle $`1`$, to be specified, in order to leave Bob’s particles $`4`$ and $`5`$ in a state looking very much like Alice’s original one. As we are going to show, it turns out that there is no possible choice of the coefficients $`a,b\mathrm{}.h`$, up to now deliberately unspecified, fulfilling the desired task of complete teleportation : it will not be possible for Bob to reconstruct the state $`|\psi `$. First of all it is worthwhile observing that in equation (2) the couples of coefficients $`(\alpha ,\delta )`$ and $`(\beta ,\gamma )`$ are only and always associated to states $`|0_1`$ and $`|1_1`$, respectively. It is so clear that successive measurement on Bell basis for particles $`2`$ and $`3`$ and on canonical basis of particle $`1`$, leave the collapsed state quite dissimilar from the one to teleport, because of the lack of needed coefficients. Suppose for example Alice obtains the state $`|\varphi ^+_{23}`$ and the state $`|0_1`$ in successive measurements: in the remaining state of the system there will be no trace of $`\beta `$ and $`\gamma `$, preventing Bob from reconstructing state $`|\psi `$. In order to overcome this difficulty, Alice must resort to perform an Hadamard transformation on particle $`1`$, so mixing up the coefficients. We note that Alice could obtain the same result by making the substitution : $$|0_1=A|\phi _1+B|\chi _1$$ (2.6) $$|1_1=B^{}|\phi _1+A^{}|\chi _1$$ In fact, in order not to have undesired coefficients in the transmitted state, she should be obliged to choose $`A`$=$`B`$=$`1/\sqrt{2}`$, obtaining an expression having exactly the same distinctive features of equation (2). This means that making the Hadamard transformation or projecting on states $`|\phi _1_1`$ and $`|\chi _1_1`$ has the same effect. Hadamard unitary transformation is defined in terms of Pauli matrices as : $$H=\frac{1}{\sqrt{2}}(\sigma _x+\sigma _z)H^2=1$$ (2.7) It acts as a rotation about the axis $`\stackrel{}{n}=1/\sqrt{2}(\stackrel{}{n_x}+\stackrel{}{n_z})`$: $$\{\begin{array}{cc}H|0=\frac{1}{\sqrt{2}}(|0+|1)\hfill & \\ H|1=\frac{1}{\sqrt{2}}(|0|1)\hfill & \end{array}$$ (2.8) After applying such a transformation, the resulting state of the system with mixed coefficients, is : $`|\stackrel{~}{\mathrm{\Omega }}`$ $`=`$ $`{\displaystyle \frac{1}{2\sqrt{N}}}|\varphi ^+_{23}\{|0_1[((\alpha +\beta )a+(\delta +\gamma )b)|00+((\alpha +\beta )c+(\delta +\gamma )e)|10+`$ $`+((\alpha +\beta )d+(\delta +\gamma )f)|01+((\alpha +\beta )g+(\delta +\gamma )h)|11]_{45}+`$ $`|1_1[((\alpha \beta )a+(\delta \gamma )b)|00+((\alpha \beta )c+(\delta \gamma )e)|10+`$ $`+((\alpha \beta )d+(\delta \gamma )f)|01+((\alpha \beta )g+(\delta \gamma )h)|11]_{45}\}+`$ $`{\displaystyle \frac{1}{2\sqrt{N}}}|\varphi ^{}_{23}\{|0_1[((\alpha +\beta )a(\delta +\gamma )b)|00+((\alpha +\beta )c(\delta +\gamma )e)|10+`$ $`+((\alpha +\beta )d(\delta +\gamma )f)|01+((\alpha +\beta )g(\delta +\gamma )h)|11]_{45}+`$ $`|1_1[((\alpha \beta )a(\delta \gamma )b)|00+((\alpha \beta )c(\delta \gamma )e)|10+`$ $`+((\alpha \beta )d(\delta \gamma )f)|01+((\alpha \beta )g(\delta \gamma )h)|11]_{45}\}+`$ $`{\displaystyle \frac{1}{2\sqrt{N}}}|\psi ^+_{23}\{|0_1[((\alpha +\beta )b+(\delta +\gamma )a)|00+((\alpha +\beta )e+(\delta +\gamma )c)|10+`$ $`+((\alpha +\beta )f+(\delta +\gamma )d)|01+((\alpha +\beta )h+(\delta +\gamma )g)|11]_{45}+`$ $`|1_1[((\alpha \beta )b+(\delta \gamma )a)|00+((\alpha \beta )e+(\delta \gamma )c)|10+`$ $`+((\alpha \beta )f+(\delta \gamma )d)|01+((\alpha \beta )h+(\delta \gamma )g)|11]_{45}\}+`$ $`{\displaystyle \frac{1}{2\sqrt{N}}}|\psi ^{}_{23}\{|0_1[((\alpha +\beta )b(\delta +\gamma )a)|00+((\alpha +\beta )e(\delta +\gamma )c)|10+`$ $`+((\alpha +\beta )f(\delta +\gamma )d)|01+((\alpha +\beta )h(\delta +\gamma )g)|11]_{45}+`$ $`|1_1[((\alpha \beta )b(\delta \gamma )a)|00+((\alpha \beta )e(\delta \gamma )c)|10+`$ $`+((\alpha \beta )f(\delta \gamma )d)|01+((\alpha \beta )h(\delta \gamma )g)|11]_{45}\}`$ Alice now accomplishes Bell measurement on particles $`2`$ and $`3`$, and canonical measurement on particle $`1`$, projecting in a such way the state firstly on Bell basis $`\{|\varphi ^\pm ,|\psi ^\pm \}`$ and secondly onto the canonical basis $`\{|0,|1\}`$: the state of the system will eventually collapse, with equal probability, in one of the eight states of particles $`4`$ and $`5`$ indicated between square brackets. It is clear that there is no possible choice of $`a,b\mathrm{}h`$ suited to reproduce the unknown initial state $`|\psi `$, because the coefficients $`\alpha ,\beta ,\delta ,\gamma `$ – although mixed – remains summed two by two ($`\alpha \pm \beta `$ and $`\delta \pm \gamma `$). Then, it is not possible to achieve complete teleportation of an arbitrary and unknown two-particle entangled state ( i.e. with $`\alpha ,\beta ,\gamma ,\delta 0`$) using the present scheme. One can then ask whether a measurement on orthonormal states $`|\varphi _i_{23}`$, with $`i=1\mathrm{}4`$, different from the Bell states, could make one to achieve the desired goal. A cumbersome but trivial calculation shows that there is no gain on projecting onto general states and in all cases teleportation cannot be obtained. In particular, since the proof is valid for every choice of non zero coefficients $`\alpha ,\beta ,\gamma ,\delta `$, it is not even possible to teleport a two-particle factorized state using the present procedure. In fact, each state of the type $`(m|0_1+n|1_1)(r|0_2+s|1_2)`$ can be written in the form (2.1); conversely, it can be trivially shown that every state of the form (2.1), for which the condition $`\alpha \delta =\gamma \beta `$ is satisfied, can be expressed as a factorized state. This result is not surprising. In fact, Alice has at her disposal only one particle, i.e. one e-bit, which is not sufficient to teleport a general state of two particles. However, in all these cases, one can obtain successful teleportation by simply repeating the original standard teleportation procedure using a sequence of two (two-particle) channels, rather than a single (three-particle) quantum channel as considered in this article. ## 3 Teleportation of peculiar states Since the present mechanism for teleportation cannot work for an arbitrary two-particle entangled state, let us try to focus our attention on some two dimensional subspaces of the whole Hilbert space of the system of the two particles ( which is four dimensional ). Let us try for example with the state $`\alpha |00+\gamma |11`$. Alice is now able to successfully perform the teleportation process by choosing a suitable quantum channel and then following the steps already considered in the previous section: 1. Alice prepares the state $`|\mathrm{\Omega }=(\alpha |00_{12}+\gamma |11_{12})|\varphi _{345}`$ where $`|\varphi _{345}`$ is the three-particle quantum channel, yet to be specified. 2. Alice acts with Hadamard unitary trasformation on states of particle $`1`$, in order to mix up in an appropriate way the coefficients $`\alpha `$ and $`\gamma `$. 3. Alice performs a Bell measurement on particles $`2`$ and $`3`$. 4. Alice performs a measurement onto states $`|0_1`$ and $`|1_1`$. The eight equally probable results are easily obtained by putting $`\beta `$=$`\delta `$=$`0`$ in equation (2) : * $`|0_1|\varphi ^+_{23}(\alpha a+\gamma b)|00_{45}+(\alpha c+\gamma e)|10_{45}+(\alpha d+\gamma f)|01_{45}+(\alpha g+\gamma h)|11_{45}`$ * $`|1_1|\varphi ^+_{23}(\alpha a\gamma b)|00_{45}+(\alpha c\gamma e)|10_{45}+(\alpha d\gamma f)|01_{45}+(\alpha g\gamma h)|11_{45}`$ * $`|0_1|\varphi ^{}_{23}(\alpha a\gamma b)|00_{45}+(\alpha c\gamma e)|10_{45}+(\alpha d\gamma f)|01_{45}+(\alpha g\gamma h)|11_{45}`$ * $`|1_1|\varphi ^{}_{23}(\alpha a+\gamma b)|00_{45}+(\alpha c+\gamma e)|10_{45}+(\alpha d+\gamma f)|01_{45}+(\alpha g+\gamma h)|11_{45}`$ * $`|0_1|\psi ^+_{23}(\alpha b+\gamma a)|00_{45}+(\alpha e+\gamma c)|10_{45}+(\alpha f+\gamma d)|01_{45}+(\alpha h+\gamma g)|11_{45}`$ * $`|1_1|\psi ^+_{23}(\alpha b\gamma a)|00_{45}+(\alpha e\gamma c)|10_{45}+(\alpha f\gamma d)|01_{45}+(\alpha h\gamma g)|11_{45}`$ * $`|0_1|\psi ^{}_{23}(\alpha b\gamma a)|00_{45}+(\alpha e\gamma c)|10_{45}+(\alpha f\gamma d)|01_{45}+(\alpha h\gamma g)|11_{45}`$ * $`|1_1|\psi ^{}_{23}(\alpha b+\gamma a)|00_{45}+(\alpha e+\gamma c)|10_{45}+(\alpha f+\gamma d)|01_{45}+(\alpha h+\gamma g)|11_{45}`$ There are now eight possible ways of choosing coefficients $`a,b\mathrm{}h`$ – so there are eight quantum channels shared between Alice and Bob – and, correspondingly, there are eight different sets of instructions to send via classical communications to Bob, in order to complete the teleportation process (see Appendix A). After having received that, Bob can successfully reconstruct the original state by applying local unitary transformations on his particles. Les us focus for example on the choice $`a=h=1,b=c=d=e=f=g=0`$, corresponding to the quantum channel $`|\varphi =1/\sqrt{2}(|000+|111)`$. We list below the set of instructions for Bob according to the results of Alice’s measurements : | Alice’s measurements | Bob’s states | Bob’s instructions | | --- | --- | --- | | $`|0_1|\varphi ^+_{23}`$ | $`\alpha |00_{45}+\gamma |11_{45}`$ | do nothing | | $`|1_1|\varphi ^+_{23}`$ | $`\alpha |00_{45}\gamma |11_{45}`$ | apply $`(\sigma _z)_4I_5`$ | | $`|0_1|\varphi ^{}_{23}`$ | $`\alpha |00_{45}\gamma |11_{45}`$ | apply $`I_4(\sigma _z)_5`$ | | $`|1_1|\varphi ^{}_{23}`$ | $`\alpha |00_{45}+\gamma |11_{45}`$ | do nothing | | $`|0_1|\psi ^+_{23}`$ | $`\gamma |00_{45}+\alpha |11_{45}`$ | apply $`(\sigma _x)_4(\sigma _x)_5`$ | | $`|1_1|\psi ^+_{23}`$ | $`\gamma |00_{45}+\alpha |11_{45}`$ | apply $`(\sigma _z\sigma _x)_4(\sigma _x)_5`$ | | $`|0_1|\psi ^{}_{23}`$ | $`\gamma |00_{45}+\alpha |11_{45}`$ | apply $`(\sigma _x)_4(\sigma _z\sigma _x)_5`$ | | $`|1_1|\psi ^{}_{23}`$ | $`\gamma |00_{45}+\alpha |11_{45}`$ | apply $`(\sigma _x)_4(\sigma _x)_5`$ | The remaining seven possible quantum channels and seven sets of instructions are listed in Appendix A. As it is shown in the same appendix, there are eight possible ways also for teleporting the state $`\beta |10+\delta |01`$, but only four possible ways for teleporting factorizable (non entangled) states like $`\beta |10+\gamma |11`$ and $`\alpha |00+\delta |01`$. It is also possible to show that the machinery doesn’t work on right-factorizable states like $`\alpha |00+\beta |10`$ and $`\delta |01+\gamma |11`$ (see appendix B). Before concluding this section it is worthwhile summarizing the obtained results in a table : | Two-particle states | Can be teleported? | | --- | --- | | $`\alpha |00_{12}+\beta |10_{12}+\delta |01_{12}+\gamma |11_{12}`$ | No | | $`\alpha |00+\gamma |11`$ | Yes, in eight different ways | | $`\beta |10+\delta |01`$ | Yes, in eight different ways | | $`\beta |10+\gamma |11`$ | Yes, in four different ways | | $`\alpha |00+\delta |01`$ | Yes, in four different ways | | $`\alpha |00+\beta |10`$ | No | | $`\delta |01+\gamma |11`$ | No | ## 4 Conclusions In this work it has been shown that the most general and unknown two-particle entangled state ( i.e. the state $`|\psi =\alpha |00+\beta |10+\delta |01+\gamma |11`$ with $`\alpha ,\beta ,\gamma ,\delta 0`$) cannot be teleported using only one (three-particle) channel and Bell measurements. We have nevertheless shown that some two-particle entangled states, belonging to two dimensional subspaces of the whole Hilbert space, can be successfully teleported from Alice to Bob using suitable and different three-particle quantum channels, with the aim of Hadamard transformation, Bell measurements and classical communication. We have listed which are the states and the sets of unitary transformations to be performed by Bob in order to recreate a perfect copy of the original state, without violating special relativity constraints ( classical communication prevents in fact from sending faster-than-light messages ) and no-cloning theorem ( the original state possessed by Alice is destroyed by Bell measurement). ## Appendix A Let us ask which are the states transmitted from Alice to Bob permitting him to reconstruct the original one. The operations Bob may use are : 1. to do nothing, if the teleported state is already the original one; 2. to make an exchange $`|0|1`$ on particle $`4`$ or on particle $`5`$, or on both; 3. to make a transformation of the $`CNOT`$ type; 4. to use a product of $`|0|1`$ exchange and $`CNOT`$ transformation. The unitary operator $`CNOT`$, which acts on two-particle states by reversing the second entry if the first is $`1`$, is defined as : $$CNOT|00=|00CNOT|01=|01$$ (4.1) $$CNOT|10=|11CNOT|11=|10$$ Then, we can find out the states sent to Bob permitting teleportation by simply applying all the possible inverse operations to the original state. The possibilities, together with the operations which must be done by Bob, are listed below. | $`|00_{45}`$ | $`|10_{45}`$ | $`|01_{45}`$ | $`|11_{45}`$ | Bob’s instructions | | --- | --- | --- | --- | --- | | $`\alpha `$ | $`\beta `$ | $`\delta `$ | $`\gamma `$ | do nothing | | $`\beta `$ | $`\alpha `$ | $`\gamma `$ | $`\delta `$ | apply $`(\sigma _x)_4`$ | | $`\delta `$ | $`\gamma `$ | $`\alpha `$ | $`\beta `$ | apply $`(\sigma _x)_5`$ | | $`\gamma `$ | $`\delta `$ | $`\beta `$ | $`\alpha `$ | apply $`(\sigma _x)_4(\sigma _x)_5`$ | | $`\alpha `$ | $`\gamma `$ | $`\delta `$ | $`\beta `$ | apply $`CNOT`$ | | $`\beta `$ | $`\delta `$ | $`\gamma `$ | $`\alpha `$ | $`(\sigma _x)_4CNOT`$ | | $`\delta `$ | $`\beta `$ | $`\alpha `$ | $`\gamma `$ | $`(\sigma _x)_5CNOT`$ | | $`\gamma `$ | $`\alpha `$ | $`\beta `$ | $`\delta `$ | apply $`(\sigma _x)_4(\sigma _x)_5CNOT`$ | It turns out that there are eight possible channels for transmitting states $`\alpha |00+\gamma |11`$ and $`\beta |10+\delta |01`$. In fact, putting $`(\beta =0,\delta =0)`$ or $`(\alpha =0,\gamma =0)`$ respectively, we obtain eight different transmitted states. In the case of states $`\alpha |00+\delta |01`$ and $`\beta |10+\gamma |11`$, the annihilation of coefficients $`\beta ,\gamma `$ and $`\alpha ,\delta `$ reduces to four the different states from which Bob can restore the original ones. We are now going to list other seven possible quantum channels $`|\varphi _{345}`$ and relative sets of instructions, with whom Alice and Bob can accomplish successful teleportation of the particular state $`\alpha |00+\gamma |11`$. The various channels, being characterized by different choices of the coefficients $`a,b\mathrm{}h`$ to be inserted in equation (2), are indicated in the following schemes together with the results of Alice’s measurements, the collapsed state of particles $`4`$ and $`5`$ and the unitary transformation which Bob must perform in order to complete teleportation process. $$\mathrm{𝐐𝐮𝐚𝐧𝐭𝐮𝐦}\mathrm{𝐂𝐡𝐚𝐧𝐧𝐞𝐥𝐬}$$ 1. Quantum Channel $`|010_{345}+|101_{345}`$ $`c=f=1,a=b=d=e=g=h=0`$ | Alice’s measurements | Bob’s states | Bob’s instructions | | --- | --- | --- | | $`|0_1|\varphi ^+_{23}`$ | $`\alpha |10_{45}+\gamma |01_{45}`$ | apply $`(\sigma _x)_4I_5`$ | | $`|1_1|\varphi ^+_{23}`$ | $`\alpha |10_{45}\gamma |01_{45}`$ | apply $`(\sigma _x)_4(\sigma _z)_5`$ | | $`|0_1|\varphi ^{}_{23}`$ | $`\alpha |10_{45}\gamma |01_{45}`$ | apply $`(\sigma _x)_4(\sigma _z)_5`$ | | $`|1_1|\varphi ^{}_{23}`$ | $`\alpha |10_{45}+\gamma |01_{45}`$ | apply $`(\sigma _x)_4I_5`$ | | $`|0_1|\psi ^+_{23}`$ | $`\gamma |10_{45}+\alpha |01_{45}`$ | apply $`I_4(\sigma _x)_5`$ | | $`|1_1|\psi ^+_{23}`$ | $`\gamma |10_{45}+\alpha |01_{45}`$ | apply $`(\sigma _z)_4(\sigma _x)_5`$ | | $`|0_1|\psi ^{}_{23}`$ | $`\gamma |10_{45}+\alpha |01_{45}`$ | apply $`(\sigma _z)_4(\sigma _x)_5`$ | | $`|1_1|\psi ^{}_{23}`$ | $`\gamma |10_{45}+\alpha |01_{45}`$ | apply $`I_4(\sigma _x)_5`$ | 2. Quantum Channel $`|100_{345}+|011_{345}`$ $`b=g=1,a=c=d=e=f=h=0`$ | Alice’s measurements | Bob’s states | Bob’s instructions | | --- | --- | --- | | $`|0_1|\varphi ^+_{23}`$ | $`\gamma |00_{45}+\alpha |11_{45}`$ | apply $`(\sigma _x)_4(\sigma _x)_5`$ | | $`|1_1|\varphi ^+_{23}`$ | $`\gamma |00_{45}+\alpha |11_{45}`$ | apply $`(\sigma _z\sigma _x)_4(\sigma _x)_5`$ | | $`|0_1|\varphi ^{}_{23}`$ | $`\gamma |00_{45}+\alpha |11_{45}`$ | apply $`(\sigma _z\sigma _x)_4(\sigma _x)_5`$ | | $`|1_1|\varphi ^{}_{23}`$ | $`\gamma |00_{45}+\alpha |11_{45}`$ | apply $`(\sigma _x)_4(\sigma _x)_5`$ | | $`|0_1|\psi ^+_{23}`$ | $`\alpha |00_{45}+\gamma |11_{45}`$ | do nothing | | $`|1_1|\psi ^+_{23}`$ | $`\alpha |00_{45}\gamma |11_{45}`$ | apply $`I_4(\sigma _z)_5`$ | | $`|0_1|\psi ^{}_{23}`$ | $`\alpha |00_{45}\gamma |11_{45}`$ | apply $`I_4(\sigma _z)_5`$ | | $`|1_1|\psi ^{}_{23}`$ | $`\alpha |00_{45}+\gamma |11_{45}`$ | do nothing | 3. Quantum Channel $`|110_{345}+|001_{345}`$ $`d=e=1,a=b=c=f=g=h=0`$ | Alice’s measurements | Bob’s states | Bob’s instructions | | --- | --- | --- | | $`|0_1|\varphi ^+_{23}`$ | $`\gamma |10_{45}+\alpha |01_{45}`$ | apply $`I_4(\sigma _x)_5`$ | | $`|1_1|\varphi ^+_{23}`$ | $`\gamma |10_{45}+\alpha |01_{45}`$ | apply $`(\sigma _z)_4(\sigma _x)_5`$ | | $`|0_1|\varphi ^{}_{23}`$ | $`\gamma |10_{45}+\alpha |01_{45}`$ | apply $`(\sigma _z)_4(\sigma _x)_5`$ | | $`|1_1|\varphi ^{}_{23}`$ | $`\gamma |10_{45}+\alpha |01_{45}`$ | apply $`I_4(\sigma _x)_5`$ | | $`|0_1|\psi ^+_{23}`$ | $`\alpha |10_{45}+\gamma |01_{45}`$ | apply $`(\sigma _x)_4I_5`$ | | $`|1_1|\psi ^+_{23}`$ | $`\alpha |10_{45}\gamma |01_{45}`$ | apply $`(\sigma _x)_4(\sigma _z)_5`$ | | $`|0_1|\psi ^{}_{23}`$ | $`\alpha |10_{45}\gamma |01_{45}`$ | apply $`(\sigma _x)_4(\sigma _z)_5`$ | | $`|1_1|\psi ^{}_{23}`$ | $`\alpha |10_{45}+\gamma |01_{45}`$ | apply $`(\sigma _x)_4I_5`$ | 4. Quantum Channel $`|000_{345}+|110_{345}`$ $`a=e=1,b=c=d=f=g=h=0`$ | Alice’s measurements | Bob’s states | Bob’s instructions | | --- | --- | --- | | $`|0_1|\varphi ^+_{23}`$ | $`\alpha |00_{45}+\gamma |10_{45}`$ | apply $`CNOT`$ | | $`|1_1|\varphi ^+_{23}`$ | $`\alpha |00_{45}\gamma |10_{45}`$ | apply $`(\sigma _z)_5CNOT`$ | | $`|0_1|\varphi ^{}_{23}`$ | $`\alpha |00_{45}\gamma |10_{45}`$ | apply $`(\sigma _z)_5CNOT`$ | | $`|0_1|\varphi ^{}_{23}`$ | $`\alpha |00_{45}+\gamma |10_{45}`$ | apply $`CNOT`$ | | $`|0_1|\psi ^+_{23}`$ | $`\gamma |00_{45}+\alpha |10_{45}`$ | apply $`(\sigma _x)_4(\sigma _x)_5CNOT`$ | | $`|1_1|\psi ^+_{23}`$ | $`\gamma |00_{45}+\alpha |10_{45}`$ | apply $`(\sigma _z\sigma _x)_4(\sigma _x)_5CNOT`$ | | $`|0_1|\psi ^{}_{23}`$ | $`\gamma |00_{45}+\alpha |10_{45}`$ | apply $`(\sigma _z\sigma _x)_4(\sigma _x)_5CNOT`$ | | $`|1_1|\psi ^{}_{23}`$ | $`\gamma |00_{45}+\alpha |10_{45}`$ | apply $`(\sigma _x)_4(\sigma _x)_5CNOT`$ | 5. Quantum Channel $`|100_{345}+|010_{345}`$ $`b=c=1,a=d=e=f=g=h=0`$ | Alice’s measurements | Bob’s states | Bob’s instructions | | --- | --- | --- | | $`|0_1|\varphi ^+_{23}`$ | $`\gamma |00_{45}+\alpha |10_{45}`$ | apply $`(\sigma _x)_4(\sigma _x)_5CNOT`$ | | $`|1_1|\varphi ^+_{23}`$ | $`\gamma |00_{45}+\alpha |10_{45}`$ | apply $`(\sigma _z\sigma _x)_4(\sigma _x)_5CNOT`$ | | $`|0_1|\varphi ^{}_{23}`$ | $`\gamma |00_{45}+\alpha |10_{45}`$ | apply $`(\sigma _z\sigma _x)_4(\sigma _x)_5CNOT`$ | | $`|1_1|\varphi ^{}_{23}`$ | $`\gamma |00_{45}+\alpha |10_{45}`$ | apply $`(\sigma _x)_4(\sigma _x)_5CNOT`$ | | $`|0_1|\psi ^+_{23}`$ | $`\alpha |00_{45}+\gamma |10_{45}`$ | apply $`CNOT`$ | | $`|1_1|\psi ^+_{23}`$ | $`\alpha |00_{45}\gamma |10_{45}`$ | apply $`(\sigma _z)_5CNOT`$ | | $`|0_1|\psi ^{}_{23}`$ | $`\alpha |00_{45}\gamma |10_{45}`$ | apply $`(\sigma _z)_5CNOT`$ | | $`|0_1|\psi ^{}_{23}`$ | $`\alpha |00_{45}+\gamma |10_{45}`$ | apply $`CNOT`$ | 6. Quantum Channel $`|101_{345}+|011_{345}`$ $`f=g=1,a=b=c=d=e=h=0`$ | Alice’s measurements | Bob’s states | Bob’s instructions | | --- | --- | --- | | $`|0_1|\varphi ^+_{23}`$ | $`\gamma |01_{45}+\alpha |11_{45}`$ | apply $`(\sigma _x)_4CNOT`$ | | $`|1_1|\varphi ^+_{23}`$ | $`\gamma |01_{45}+\alpha |11_{45}`$ | apply $`(\sigma _z)_5(\sigma _x)_4CNOT`$ | | $`|0_1|\varphi ^{}_{23}`$ | $`\gamma |01_{45}+\alpha |11_{45}`$ | apply $`(\sigma _z)_5(\sigma _x)_4CNOT`$ | | $`|1_1|\varphi ^{}_{23}`$ | $`\gamma |01_{45}+\alpha |11_{45}`$ | apply $`(\sigma _x)_4CNOT`$ | | $`|0_1|\psi ^+_{23}`$ | $`\alpha |01_{45}+\gamma |11_{45}`$ | apply $`(\sigma _x)_5CNOT`$ | | $`|1_1|\psi ^+_{23}`$ | $`\alpha |01_{45}\gamma |11_{45}`$ | apply $`(\sigma _z)_4(\sigma _x)_5CNOT`$ | | $`|0_1|\psi ^{}_{23}`$ | $`\alpha |01_{45}\gamma |11_{45}`$ | apply $`(\sigma _z)_4(\sigma _x)_5CNOT`$ | | $`|0_1|\psi ^{}_{23}`$ | $`\alpha |01_{45}+\gamma |11_{45}`$ | apply $`(\sigma _x)_5CNOT`$ | 7. Quantum Channel $`|001_{345}+|111_{345}`$ $`d=h=1,a=b=c=e=f=g=0`$ | Alice’s measurements | Bob’s states | Bob’s instructions | | --- | --- | --- | | $`|0_1|\varphi ^+_{23}`$ | $`\alpha |01_{45}+\gamma |11_{45}`$ | apply $`(\sigma _x)_5CNOT`$ | | $`|1_1|\varphi ^+_{23}`$ | $`\alpha |01_{45}\gamma |11_{45}`$ | apply $`(\sigma _z)_4(\sigma _x)_5CNOT`$ | | $`|0_1|\varphi ^{}_{23}`$ | $`\alpha |01_{45}\gamma |11_{45}`$ | apply $`(\sigma _z)_4(\sigma _x)_5CNOT`$ | | $`|0_1|\varphi ^{}_{23}`$ | $`\alpha |01_{45}+\gamma |11_{45}`$ | apply $`(\sigma _x)_5CNOT`$ | | $`|0_1|\psi ^+_{23}`$ | $`\gamma |01_{45}+\alpha |11_{45}`$ | apply $`(\sigma _x)_4CNOT`$ | | $`|1_1|\psi ^+_{23}`$ | $`\gamma |01_{45}+\alpha |11_{45}`$ | apply $`(\sigma _z)_5(\sigma _x)_4CNOT`$ | | $`|0_1|\psi ^{}_{23}`$ | $`\gamma |01_{45}+\alpha |11_{45}`$ | apply $`(\sigma _z)_5(\sigma _x)_4CNOT`$ | | $`|1_1|\psi ^{}_{23}`$ | $`\gamma |01_{45}+\alpha |11_{45}`$ | apply $`(\sigma _x)_4CNOT`$ | ## Appendix B For the sake of simplicity we will only enumerate the permitted three-particle quantum channels that Alice and Bob may use to teleport the following two-particle states : | Two-particle states | Quantum channels | | --- | --- | | $`[\beta |10+\delta |01]_{12}`$ | $`(|010+|101)_{345}`$ | | | $`(|000+|111)_{345}`$ | | | $`(|100+|011)_{345}`$ | | | $`(|110+|001)_{345}`$ | | | $`(|001+|111)_{345}`$ | | | $`(|101+|011)_{345}`$ | | | $`(|000+|110)_{345}`$ | | | $`(|100+|010)_{345}`$ | | $`[\alpha |00+\delta |01]_{12}`$ | $`(|000+|110)_{345}`$ | | | $`(|100+|010)_{345}`$ | | | $`(|001+|111)_{345}`$ | | | $`(|101+|011)_{345}`$ | | Two-particle states | Quantum channels | | --- | --- | | $`[\beta |10+\gamma |11]_{12}`$ | $`(|000+|110)_{345}`$ | | | $`(|100+|010)_{345}`$ | | | $`(|001+|111)_{345}`$ | | | $`(|101+|011)_{345}`$ | Two-particle factorized states in which the unknown part is in channel $`1`$ ( i.e., states $`\alpha |00_{12}+\beta |10_{12}(\alpha |0_1+\beta |1_1)|0_2`$ and $`\delta |01_{12}+\gamma |11_{12}(\delta |0_1+\gamma |1_1)|1_2`$ cannot be transmitted using the present method. Let us consider for example the state $`\alpha |00_{12}+\beta |10_{12}`$; this means to choose $`\gamma =0`$ and $`\delta =0`$ in equation (2), which becomes : $`|\stackrel{~}{\mathrm{\Omega }}`$ $`=`$ $`{\displaystyle \frac{1}{2\sqrt{N}}}[(\alpha +\beta )|0_1+(\alpha \beta )|1_1][|\varphi ^+_{23}+|\varphi ^{}_{23}+|\psi ^+_{23}+|\psi ^{}_{23}]`$ (4.2) $`[a|00_{45}+c|10_{45}+d|01_{45}+g|11_{45}]`$ and the teleporting method cannot be applied. The unknown state must be put in channel $`2`$, which is involved in the Bell measurement. However, as already noted, in such a case one can resort to the original standard teleportation procedure.
warning/0004/hep-th0004015.html
ar5iv
text
# 1 Introduction ## 1 Introduction During the last few years, it has been realized that Type IIA (IIB) string theory contains non-BPS D-branes of odd (even) spatial dimensions. An analysis of the open string states living on such a brane reveals the existence of a tachyon which comes from the NS sector. It has been argued by Sen that the tachyon potential has a minimum and that the minimum corresponds to the usual vacuum of closed string theory without any D-brane. This implies that the negative energy density contribution from the tachyon potential must exactly cancel the tension of the non-BPS D-brane. To analyze the tachyon potential in order to test this conjecture, ordinary first quantized string theory cannot be used since the zero momentum tachyon state is not on-shell. We need an off-shell formalism of string theory. Such a formalism is given by string field theory. Recently it has been realized that calculating the tachyon potential in string field theory by including fields and terms in the string field action up to a certain level gives a good approximation to the tachyon potential. In , the zeroth order and the first non-trivial correction to the tachyon potential on a non-BPS D-brane of Type II string theory was computed using open string field theory action formulated in . It was found that the potential has a minimum which cancels 60% of the D-brane tension at the zeroth level and 85% of the D-brane tension at the next non-trivial level (level 3/2). In this paper, we extend the work of to include fields up to level 2. Similar computation was done in . This paper is organized as follows: In section 2, we review the formalism of to set up the calculation. In section 3, we write down the expansion of the string field including fields up to level 2. We also give some details of the calculation — we give one simple example of calculation of a term in the action at each level. In the appendix, we list the conformal and BRST transformations of operators that we need in the calculation. ## 2 Superstring field theory on a Non-BPS D-brane Superstring field theory on a non-BPS D-brane was described in detail in . Here, we will just review the basic features to set up the calculation and establish notation. We will first review the essential features of string field theory on a BPS D-brane which only has a GSO(+) sector and then introduce Chan Paton matrices and GSO($``$) sector to discuss the non-BPS D-brane. We follow the notation and conventions of throughout the paper. #### BPS D-brane In , a string field configuration in the GSO$`(+)`$ NS sector corresponds to a Grassmann even open string vertex operator $`\mathrm{\Phi }`$ of ghost and picture number 0 in the combined conformal field theory of a $`c=15`$ superconformal matter system, and the $`b,c,\beta ,\gamma `$ ghost system with $`c=15`$. Here, $`b`$, $`c`$ are the reparameterization ghosts and $`\beta `$, $`\gamma `$ are their superconformal partners. The $`\beta `$, $`\gamma `$ ghost system can be bosonized and can be replaced by ghost fields $`\xi ,\eta ,\varphi `$, related to $`\beta ,\gamma `$ through the relations $$\beta =\xi e^\varphi ,\gamma =\eta e^\varphi .$$ The ghost number ($`n_g`$) and picture number ($`n_p`$) assignments and the conformal weights ($`h`$) are as follows : $`b:n_g=1,n_p=0,h=2c:n_g=1,n_p=0,h=1,`$ $`e^{q\varphi }:n_g=0,n_p=q,h=(q+{\displaystyle \frac{q^2}{2}}),`$ $`\xi :n_g=1,n_p=1,h=0,\eta :n_g=1,n_p=1,h=1.`$ The SL(2,R) invariant vacuum has zero ghost and picture number. We shall denote by $`_iA_i`$ the correlation function of a set of vertex operators in the combined matter-ghost conformal field theory on the unit disk with open string vertex operators inserted on the boundary of the disk, without including trace over CP factors. These correlation functions are to be computed with the normalization $$\xi (z)cc^2c(w)e^{2\varphi (y)}=2.$$ (1) The BRST operator is given by $$Q_B=Q_0+Q_1+Q_2,$$ where $`Q_0`$ $`=`$ $`={\displaystyle 𝑑zc(T_m+T_{\xi \eta }+T_\varphi )}+ccb,`$ $`Q_1`$ $`=`$ $`{\displaystyle 𝑑z\eta e^\varphi G_m},Q_2={\displaystyle 𝑑z\eta \eta e^{2\varphi }b}.`$ (2) Here $$T_{\xi \eta }=\xi \eta ,T_\varphi =\frac{1}{2}\varphi \varphi ^2\varphi ,$$ $`T_m`$ is the matter stress tensor and $`G_m`$ is the matter superconformal generator. $`G_m`$ is a dimension $`3/2`$ primary field and satisfies: $$G_m(z)G_m(w)\frac{10}{(zw)^3}+\frac{2T_m}{(zw)}.$$ The normalization of $`\varphi `$, $`\xi `$, $`\eta `$, $`b`$ and $`c`$ are as follows: $$\xi (z)\eta (w)\frac{1}{zw},b(z)c(w)\frac{1}{zw},\varphi (z)\varphi (w)\frac{1}{(zw)^2}.$$ We denote by $`\eta _0=𝑑z\eta (z)`$ the zero mode of the field $`\eta `$ acting on the Hilbert space of matter ghost CFT. The string field theory action is given by $$S=\frac{1}{2g^2}(e^\mathrm{\Phi }Q_Be^\mathrm{\Phi })(e^\mathrm{\Phi }\eta _0e^\mathrm{\Phi })_0^1𝑑t(e^{t\mathrm{\Phi }}_te^{t\mathrm{\Phi }})\{(e^{t\mathrm{\Phi }}Q_Be^{t\mathrm{\Phi }}),(e^{t\mathrm{\Phi }}\eta _0e^{t\mathrm{\Phi }})\}.$$ (3) This action is defined by expanding all exponentials in a formal Taylor series. $`A_1,\mathrm{}A_n`$ is defined as: $$A_1\mathrm{}A_n=f_1^{(n)}A_1(0)\mathrm{}f_n^{(n)}A_n(0).$$ (4) Here, $`fA`$ for any function $`f(z)`$, denotes the conformal transform of $`A`$ by $`f`$, and $$f_k^{(n)}(z)=e^{\frac{2\pi i(k1)}{n}}\left(\frac{1+iz}{1iz}\right)^{2/n}\text{for}n1.$$ (5) In particular if $`\phi `$ denotes a primary field of weight $`h`$, then $$f\phi (0)=(f^{}(0))^h\phi (f(0)).$$ Since we have, in general, non-integer weight vertex operators, we should be more careful in defining $`fA`$ for such vertex operators. Noting that $$f_k^{(N)}(0)=\frac{4i}{N}e^{2\pi i\frac{k1}{N}}\frac{4}{N}e^{2\pi i(\frac{k1}{N}+\frac{1}{4})},$$ we adopt the following definition of $`f_k^{(N)}\phi (0)`$ for a primary vertex operator $`\phi (x)`$ of conformal weight $`h`$: $$f_k^{(N)}\phi (0)=\left|\left(\frac{4}{N}\right)^h\right|e^{2\pi ih(\frac{k1}{N}+\frac{1}{4})}\phi (f_k^{(N)}(0)).$$ (6) We also recall identities which will be used in the calculation later: $$\{Q_B,\eta _0\}=0,Q_B^2=\eta _0^2=0,$$ $$Q_B(\mathrm{\Phi }_1\mathrm{\Phi }_2)=(Q_B\mathrm{\Phi }_1)\mathrm{\Phi }_2+\mathrm{\Phi }_1(Q_B\mathrm{\Phi }_2),\eta _0(\mathrm{\Phi }_1\mathrm{\Phi }_2)=(\eta _0\mathrm{\Phi }_1)\mathrm{\Phi }_2+\mathrm{\Phi }_1(\eta _0\mathrm{\Phi }_2),$$ (7) $$Q_B(\mathrm{})=\eta _0(\mathrm{})=0.$$ Note that in the identities of the second line there are no minus signs necessary as $`Q_B`$ or $`\eta _0`$ go through the string field because the string field is Grassmann even (since we are in the GSO($`+`$) sector only). #### Non-BPS D-brane The algebraic structure described in the previous section works for the GSO($`+`$) (Grassman even) sector living on a BPS D-brane. However, on a non-BPS D-brane, the open string states live in both the GSO($``$) and GSO($`+`$) sector. The GSO($``$) states are Grassman odd and to incorporate them in the algebraic structure of the previous section, internal $`2\times 2`$ Chan Paton matrices were introduced in . These are added both to the vertex operators and $`Q_B`$ and $`\eta _0`$ as described in detail below. The $`2\times 2`$ identity matrix $`I`$ is attached on the usual GSO$`(+)`$ sector and the Pauli matrix $`\sigma _1`$ to the GSO$`()`$ sector. The complete string field is thus written as $$\widehat{\mathrm{\Phi }}=\mathrm{\Phi }_+I+\mathrm{\Phi }_{}\sigma _1,$$ where the subscripts denote the $`()^F`$ eigenvalue of the vertex operator. In addition, $`Q_B`$ and $`\eta _0`$ are tensored with $`\sigma _3`$: $$\widehat{Q}_B=Q_B\sigma _3,\widehat{\eta }_0=\eta _0\sigma _3.$$ We define $$\widehat{A}_1\mathrm{}\widehat{A}_n=\mathrm{Tr}f_1^{(n)}\widehat{A}_1(0)\mathrm{}f_n^{(n)}\widehat{A}_n(0),$$ (8) where the trace is over the internal Chan Paton matrices. As in , fields or operators with internal CP matrices included are denoted by symbols with a hat on them, and fields or operators without internal CP matrices included are denoted by symbols without a hat. In addition, we have the analogs of (7) holding: $$\{\widehat{Q}_B,\widehat{\eta }_0\}=0,\widehat{Q}_{B}^{}{}_{}{}^{2}=\widehat{\eta }_{0}^{}{}_{}{}^{2}=0,$$ $$\widehat{Q}_B(\widehat{\mathrm{\Phi }}_1\widehat{\mathrm{\Phi }}_2)=(\widehat{Q}_B\widehat{\mathrm{\Phi }}_1)\widehat{\mathrm{\Phi }}_2+\widehat{\mathrm{\Phi }}_1(\widehat{Q}_B\widehat{\mathrm{\Phi }}_2),\widehat{\eta }_0(\widehat{\mathrm{\Phi }}_1\widehat{\mathrm{\Phi }}_2)=(\widehat{\eta }_0\widehat{\mathrm{\Phi }}_1)\widehat{\mathrm{\Phi }}_2+\widehat{\mathrm{\Phi }}_1(\widehat{\eta }_0\widehat{\mathrm{\Phi }}_2),$$ (9) $$\widehat{Q}_B(\mathrm{})=\widehat{\eta }_0(\mathrm{})=0.$$ The reason no extra signs appear in the middle equation is that when the string field is Grassmann odd the sign arising by moving $`Q_B`$ across the vertex operator is canceled by having to move $`\sigma _3`$ across $`\sigma _1`$. Given that the relations satisfied by the hatted objects are the same as those of the unhatted ones, the string field action for the non-BPS D-brane takes the same structural form as that in (3) and is given by : $$S=\frac{1}{4g^2}(e^{\widehat{\mathrm{\Phi }}}\widehat{Q}_Be^{\widehat{\mathrm{\Phi }}})(e^{\widehat{\mathrm{\Phi }}}\widehat{\eta }_0e^{\widehat{\mathrm{\Phi }}})_0^1𝑑t(e^{t\widehat{\mathrm{\Phi }}}_te^{t\widehat{\mathrm{\Phi }}})\{(e^{t\widehat{\mathrm{\Phi }}}\widehat{Q}_Be^{t\widehat{\mathrm{\Phi }}}),(e^{t\widehat{\mathrm{\Phi }}}\widehat{\eta }_0e^{t\widehat{\mathrm{\Phi }}})\},$$ (10) Here the overall normalization is divided by a factor of two in order to compensate for the trace operation on the internal matrices. ## 3 Tachyon Potential and Level Approximation To study the phenomenon of tachyon condensation on the non-BPS D-brane, we will restrict our string field to be in $`_1`$ , the subset of vertex operators of ghost and picture number zero, created from the matter stress tensor ($`T_m(z)`$), its superconformal partner ($`G_m(z)`$) and the ghost fields $`b,c,\xi ,\eta ,\varphi `$. The level of a string field component multiplying a vertex operator of conformal weight $`h`$ is defined to be $`(h+\frac{1}{2})`$, so that the tachyon field multiplying the vertex operator $`\xi ce^\varphi \sigma _1`$, has level zero. The level of a given term in the string field action is defined to be the sum of the levels of the individual fields appearing in that term, and define a level $`2n`$ approximation to the action to be the one obtained by including fields up to level $`n`$ and terms in the action up to level $`2n`$. Thus for example, a level 4 approximation to the action will involve string field components up to level 2. This is the approximation we shall be using to compute the action. The string field action has a gauge invariance which can be used to choose a gauge in which $$b_0\widehat{\mathrm{\Phi }}=0,\xi _0\widehat{\mathrm{\Phi }}=0,$$ (11) which is valid at least at the linearized level. All relevant states in the “small” Hilbert space can be obtained by acting with ghost number zero combinations of oscillators $`\{b,c,\beta ,\gamma ,L^m,G^m\}`$ on $`|\stackrel{~}{\mathrm{\Omega }}`$. The $`b_0\widehat{\mathrm{\Phi }}=0`$ gauge condition allows us to ignore states with a $`c_0`$ oscillator in them. The states one finds up to $`L_0`$ eigenvalue $`\frac{3}{2}`$ are given in Table 1. For ease of notation we have not included the CP factor. The string field $`\mathrm{\Phi }`$ which uses the “large” Hilbert space, is obtained by acting the states of the table with $`\xi _0=𝑑z\frac{\xi (z)}{z}`$. As shown in , the string field theory action, restricted to $`_1`$ has a $`Z_2`$ twist symmetry under which string field components associated with a vertex operator of dimension $`h`$ carry charge $`(1)^{h+1}`$ for even $`2h`$, and $`(1)^{h+\frac{1}{2}}`$ for odd $`2h`$. Thus it is possible to consistently restrict the string field $`\widehat{\mathrm{\Phi }}`$ to be twist even. #### States and Vertex Operators The states that appear in the string field up to level two are $$|\stackrel{~}{\mathrm{\Omega }},\{c_1\beta _{\frac{1}{2}},b_1\gamma _{\frac{1}{2}},G_{\frac{3}{2}}\}|\stackrel{~}{\mathrm{\Omega }},\{b_1c_1,\beta _{\frac{1}{2}}\gamma _{\frac{3}{2}},\beta _{\frac{3}{2}}\gamma _{\frac{1}{2}},(\beta _{\frac{1}{2}}\gamma _{\frac{1}{2}})^2,L_2\}|\stackrel{~}{\mathrm{\Omega }},$$ where $`|\stackrel{~}{\mathrm{\Omega }}`$ is the tachyon state. The contribution to the tachyon potential of the tachyon and the three level $`\frac{3}{2}`$ states was calculated in . We are interested in the contribution of level two states to the tachyon potential. The vertex operators corresponding to these level two states are linear combination of the following five vertex operators $$^2ce^\varphi ,T_mce^\varphi ,T_{\xi \eta }ce^\varphi ,T_\varphi ce^\varphi ,^2\varphi ce^\varphi .$$ We can pass to the string field $`\widehat{\mathrm{\Phi }}`$ by acting the above vertex operators with $`\xi _0`$. Denoting the tachyon vertex operator by $`\widehat{T}`$, the three vertex operators at level $`\frac{3}{2}`$ by $`\widehat{A}`$, $`\widehat{E}`$ and $`\widehat{F}`$ and the five vertex operators at level 2 by $`\widehat{V}_\alpha `$ ($`\alpha =1\mathrm{}5`$) $`\widehat{T}`$ $`=`$ $`\xi ce^\varphi \sigma _1,\widehat{A}=\xi \xi c^2ce^{2\varphi }I,\widehat{E}=\xi \eta I,\widehat{F}=\xi G_mce^\varphi I`$ $`\widehat{V}_1`$ $`=`$ $`\xi T_mce^\varphi \sigma _1,\widehat{V}_2=\xi ^2ce^\varphi \sigma _1,\widehat{V}_3=\xi T_{\xi \eta }ce^\varphi \sigma _1,`$ $`\widehat{V}_4`$ $`=`$ $`\xi T_\varphi ce^\varphi \sigma _1,\widehat{V}_5=\xi ^2\varphi ce^\varphi \sigma _1.`$ Therefore, the general twist even string field up to level $`\frac{3}{2}`$, satisfying the gauge condition (11), has the following form: $$\widehat{\mathrm{\Phi }}=t\widehat{T}+a\widehat{A}+e\widehat{E}+f\widehat{F}+v_1\widehat{V}_1+v_2\widehat{V}_2+v_3\widehat{V}_3++v_4\widehat{V}_4+v_5\widehat{V}_5.$$ (13) #### Expansion of the String Field Action We shall now substitute (13) into the action (10) and keep terms to all orders in $`t`$, but only up to quadratic order in $`a`$, $`e`$, $`f`$ and $`v_\alpha `$. Although the string field action contains vertices up to arbitrarily high order, at a given level, the action has a finite number of terms (as shown in ). The string field action, for $`\widehat{\mathrm{\Phi }}`$ restricted to $`_1`$ and twist even fields is given by $`S`$ $`={\displaystyle \frac{1}{2g^2}}{\displaystyle \frac{1}{2}}(\widehat{Q}_B\widehat{\mathrm{\Phi }})(\widehat{\eta }_0\widehat{\mathrm{\Phi }})+{\displaystyle \frac{1}{3}}(\widehat{Q}_B\widehat{\mathrm{\Phi }})\widehat{\mathrm{\Phi }}(\widehat{\eta }_0\widehat{\mathrm{\Phi }})+{\displaystyle \frac{1}{12}}(\widehat{Q}_B\widehat{\mathrm{\Phi }})(\widehat{\mathrm{\Phi }}^2(\widehat{\eta }_0\widehat{\mathrm{\Phi }})\widehat{\mathrm{\Phi }}(\widehat{\eta }_0\widehat{\mathrm{\Phi }})\widehat{\mathrm{\Phi }})`$ (16) $`+{\displaystyle \frac{1}{60}}(\widehat{Q}_B\widehat{\mathrm{\Phi }})\left(\widehat{\mathrm{\Phi }}^3(\widehat{\eta }_0\widehat{\mathrm{\Phi }})3\widehat{\mathrm{\Phi }}^2(\widehat{\eta }_0\widehat{\mathrm{\Phi }})\widehat{\mathrm{\Phi }}\right)`$ $`+{\displaystyle \frac{1}{360}}(\widehat{Q}_B\widehat{\mathrm{\Phi }})(\widehat{\mathrm{\Phi }}^4(\widehat{\eta }_0\widehat{\mathrm{\Phi }})4\widehat{\mathrm{\Phi }}^3(\widehat{\eta }_0\widehat{\mathrm{\Phi }})\widehat{\mathrm{\Phi }}+3\widehat{\mathrm{\Phi }}^2(\widehat{\eta }_0\widehat{\mathrm{\Phi }})\widehat{\mathrm{\Phi }}^2).`$ ### Calculation of the Tachyon Potential The calculation of the tachyon potential is simple but tedious. Using the expansion of $`\widehat{\mathrm{\Phi }}`$ to level 4 fields (13) in the string field action (16), we collect various terms in the fields $`t,a,e,f`$ and $`v_\alpha `$. The coefficient of each term is a sum of “big” correlators $`\mathrm{}`$. We can then use the definition of $`\mathrm{}`$<sup>1</sup><sup>1</sup>1We will need to conformal and BRST transformation properties of the various operators appearing inside $`\mathrm{}`$. These are given in the appendix. in (4), and trace over the $`\sigma `$ matrices to express each “big” correlator in terms of ordinary correlators in the matter ghost CFT. These can then be calculated using the normalization in (1). We will give a simple example at each level of the action to illustrate the procedure and give some useful tricks which simplify the calculation. #### Level 2 terms The only non zero level 2 terms in the action involving $`v_\alpha `$’s are of the form $`t^3v_\alpha `$ <sup>2</sup><sup>2</sup>2$`tv_\alpha `$ terms are also level 2 but the coefficients of these terms vanish. . The relevant term in the string field action which gives rise to this term is $$S_{quartic}=\frac{1}{2g^2}\frac{1}{12}(\widehat{Q}_B\widehat{\mathrm{\Phi }})\left(\widehat{\mathrm{\Phi }}^2(\widehat{\eta }_0\widehat{\mathrm{\Phi }})\widehat{\mathrm{\Phi }}(\widehat{\eta }_0\widehat{\mathrm{\Phi }})\widehat{\mathrm{\Phi }}\right).$$ We can write $`\widehat{\mathrm{\Phi }}=\mathrm{\Phi }_+I+\mathrm{\Phi }_{}\sigma _1`$, $`\widehat{Q}_B=Q_B\sigma _3`$ and $`\eta _0=\eta _0\sigma _3`$. Tracing over the $`\sigma `$ matrices, we obtain $`12g^2S_{quartic}`$ $`=`$ $`(Q_B\mathrm{\Phi }_{})\mathrm{\Phi }_{}^2(\eta _0\mathrm{\Phi }_{})+(Q_B\mathrm{\Phi }_{})\mathrm{\Phi }_{}(\eta _0\mathrm{\Phi }_{})\mathrm{\Phi }_{}+\mathrm{}`$ (17) where $`\mathrm{}`$ represents other terms in $`S_{quartic}`$ which are irrelevant for the calculation of the coefficient of $`t^3v_\alpha `$ (they are also irrelevant for $`t^2v_\alpha v_\beta `$ which are level 4 terms). Here $$\mathrm{\Phi }_{}=tT+v_1V_1+v_2V_2+v_3V_3+v_4V_4+v_5V_5.$$ (18) Substituting (18) in (17) and collecting the $`t^3v_\alpha `$ terms together, we find $`12g^2S=\mathrm{}`$ $`+`$ $`t^3v_\alpha (\left(Q_BT\right)TT\left(\eta _0V_\alpha \right)+\left(Q_BT\right)TV_\alpha \left(\eta _0T\right)+\left(Q_BT\right)V_\alpha T\left(\eta _0T\right)`$ (19) $`+`$ $`\left(Q_BV_\alpha \right)TT\left(\eta _0T\right)+\left(Q_BT\right)T\left(\eta _0V_\alpha \right)T+\left(Q_BT\right)T\left(\eta _0T\right)V_\alpha `$ $`+`$ $`\left(Q_BT\right)V_\alpha \left(\eta _0T\right)T+\left(Q_BV_\alpha \right)T\left(\eta _0T\right)T)\mathrm{}`$ To calculate the $`t^3v_\alpha `$ term, we only need to calculate one correlator at general points: $`(Q_2T)TV_\alpha (\eta _0T)`$. All other correlators which appear in (19) can be related to this one by using $`Q_B(\mathrm{})=0`$ and $`\eta _0(\mathrm{})=0`$ as we now show. We define <sup>3</sup><sup>3</sup>3We will drop the $`n`$ index from $`f_k^{(n)}`$ for clarity. In what follows in the calculation of level 2 terms, it is assumed that $`n=4`$. $`X(I,J,K,L)`$ $``$ $`\left(Q_2T\left(w_I\right)\right)T\left(w_J\right)V_\alpha \left(w_K\right)\left(\eta _0T\left(w_L\right)\right)`$ (20) $``$ $`{\displaystyle \frac{\left(f_K^{}\right)^{\frac{3}{2}}}{\left(f_I^{}f_J^{}f_L^{}\right)^{\frac{1}{2}}}}\left(Q_2T\left(w_I\right)\right)T\left(w_J\right)\left(V_\alpha \left(w_K\right)+𝒫_{V_\alpha }^KU_\alpha \left(w_K\right)+_{V_\alpha }^KT\left(w_K\right)\right)\left(\eta _0T\left(w_L\right)\right),`$ where $`𝒫_{V_\alpha }^K`$ and $`_{V_\alpha }^K`$ are defined in the appendix, $`U_\alpha `$ is the operator multiplying $`𝒫_{V_\alpha }^K`$ in the conformal transformation of $`V_\alpha `$ ($`U_1=0,U_2=\xi ce^\varphi ,U_3=\xi ce^\varphi ,U_4=U_5=\xi \varphi ce^\varphi `$). Notice that $`X(1,2,3,4)=(Q_BT)TV_\alpha (\eta _0T)`$ where we have defined $`w_1=1,w_2=i,w_3=1,w_4=i,`$ and $`i=\sqrt{1}`$. Using $`\eta _0((Q_BT)TV_\alpha T)=0`$ we obtain<sup>4</sup><sup>4</sup>4We have used $`\eta _0Q_2T=Q_2\eta _0T=0`$. $`\left(Q_2T\left(w_I\right)\right)T\left(w_J\right)\left(\eta _0V\left(w_K\right)\right)T\left(w_L\right)`$ $`=`$ $`\left(Q_2T\left(w_I\right)\right)\left(\eta _0T\left(w_J\right)\right)V_\alpha \left(w_K\right)T\left(w_L\right)`$ $`+\left(Q_2T\left(w_I\right)\right)T\left(w_J\right)V_\alpha \left(w_K\right)\left(\eta _0T\left(w_L\right)\right),`$ $`=`$ $`X(I,L,K,J)+X(I,J,K,L).`$ Similarly, using $`Q_2\left(V_\alpha TT(\eta _0T)\right)=0`$ we obtain $$\left(Q_2V_\alpha \left(w_I\right)\right)T\left(w_J\right)T\left(w_K\right)\left(\eta _0T\left(w_L\right)\right)=X(K,J,I,L)X(J,K,I,L).$$ Hence we can write (19) in terms of $`X(I,J,K,L)`$ <sup>5</sup><sup>5</sup>5$`S_{t^3v_\alpha }`$ is the coefficient of the $`t^3v_\alpha `$ term in the action.: $`12g^2S_{t^3v_\alpha }`$ $`=`$ $`X(1,2,3,4)X(1,3,4,2)+X(1,2,4,3)+X(1,2,3,4)X(1,3,2,4)X(1,4,3,2)`$ (21) $`X(1,4,2,3)+X(1,2,4,3)+X(3,2,1,4)X(2,3,1,4)+X(4,2,1,3)`$ $`X(2,4,1,3).`$ We will give some explicit results to calculate $`X(I,J,K,L)`$ for the case $`t^3v_2`$. The various CFT correlators appearing in (20) for $`\alpha =2`$ are: $`\left(Q_2T\left(w_I\right)\right)T\left(w_J\right)V_2\left(w_K\right)\left(\eta _0T\left(w_L\right)\right)`$ $`=`$ $`2{\displaystyle \frac{w_{IL}}{w_{KL}}},`$ $`\left(Q_2T\left(w_I\right)\right)T\left(w_J\right)\xi ce^\varphi \left(w_K\right)\left(\eta _0T\left(w_L\right)\right)`$ $`=`$ $`{\displaystyle \frac{w_{IL}}{w_{KL}}}\left(w_{JK}w_{KL}\right),`$ $`\left(Q_2T\left(w_I\right)\right)T\left(w_J\right)T\left(w_K\right)\left(\eta _0T\left(w_L\right)\right)`$ $`=`$ $`w_{JK}w_{IL}.`$ Here $`w_{IJ}=w_Iw_J`$ etc. Using these expressions in (20) and substituting in (21), we find $$g^2S=\mathrm{}\frac{17}{6}t^3v_2\mathrm{}$$ #### Level 4 At level 4, the non-zero terms are of the form $`v_\alpha v_\beta `$ and $`t^2v_\alpha v_\beta `$. As an example, we will discuss the $`v_4v_5`$ term. $`2g^2S_{v_\alpha v_\beta }`$ $`=`$ $`\frac{1}{2}\left((\widehat{Q}_B\widehat{V}_\alpha )(\widehat{\eta }_0\widehat{V}_\beta )+(\widehat{Q}_B\widehat{V}_\beta )(\widehat{\eta }_0\widehat{V}_\alpha )\right),`$ (22) $`=`$ $`\frac{1}{2}\mathrm{Tr}(\sigma _3\sigma _1\sigma _3\sigma _1)\left((Q_BV_\alpha )(\eta _0V_\beta )+(Q_BV_\beta )(\eta _0V_\alpha )\right),`$ $`=`$ $`\left((Q_BV_\alpha )(\eta _0V_\beta )+(Q_BV_\beta )(\eta _0V_\alpha )\right).`$ $`(Q_BV_\beta )(\eta _0V_\alpha )`$ can be related to $`(Q_BV_\alpha )(\eta _0V_\beta )`$ as follows<sup>6</sup><sup>6</sup>6We have used $`Q_B\left(V_\alpha (\eta _0V_\beta )\right)=(Q_BV_\alpha )(\eta _0V_\beta )V_\alpha (Q_B\eta _0V_\beta )`$ which is different from (9). $`V_\alpha `$, being in the GSO($``$) sector is Grassman odd and we had multiplied it by $`\sigma _1`$ in section which resulted in (9). Here, we have traced over the $`\sigma `$ matrices. $$0=Q_B\left(V_\beta (\eta _0V_\alpha )\right)=(Q_BV_\beta )(\eta _0V_\alpha )V_\beta (Q_B\eta _0V_\alpha )=(Q_BV_\beta )(\eta _0V_\alpha )+V_\beta (\eta _0Q_BV_\alpha ),$$ $$0=\eta _0\left(V_\beta (Q_BV_\alpha )\right)=(\eta _0V_\beta )(Q_BV_\alpha )V_\beta (\eta _0Q_BV_\alpha ),$$ which implies $`2g^2S_{v_\alpha v_\beta }`$ $`=`$ $`(Q_BV_\alpha )(\eta _0V_\beta )(\eta _0V_\beta )(Q_BV_\alpha ),`$ $`=`$ $`Y(1,2)Y(2,1).`$ $`Y(I,J)`$ $``$ $`\left(Q_BV_\alpha \left(w_I\right)\right)\left(\eta _0V_\beta \left(w_J\right)\right),`$ (23) $``$ $`\left(f_i^{^{}}\right)^{3/2}\left(f_j^{^{}}\right)^{3/2}Q_0(V_\alpha \left(w_I\right)+𝒫_{V_\alpha }^IU_\alpha \left(w_I\right)+_{V_\alpha ^I}T\left(w_I\right))`$ $`\eta _0(V_\beta \left(w_J\right)+𝒫_{V_\beta }^JU_\beta \left(w_J\right)+_{V_\alpha ^J}T\left(w_J\right)).`$ $`𝒫_{V_\alpha }^I`$ and $`_{V_\alpha }^I`$ are defined in the appendix <sup>7</sup><sup>7</sup>7We have suppressed indices corresponding to $`n`$, the number of vertex operators inside $`\mathrm{}`$. For this subsection, $`n=2`$.. Here, we have used $`\varphi `$ momentum conservation to replace $`Q_B`$ by $`Q_0`$ since this is the only part of $`Q_B`$ which contributes in the correlators. | $`\left(Q_0V_4(w_I)\right)\left(\eta _0V_5(w_J)\right)=\frac{15}{4}\frac{1}{(w_Iw_I)^3},`$ | $`\left(Q_0\xi c\varphi e^\varphi (w_I)\right)\left(\eta _0V_5(w_J)\right)=\frac{1}{2}\frac{1}{(w_Iw_J)^2},`$ | | --- | --- | | $`\left(Q_0T(w_I)\right)\left(\eta _0V_5(w_J)\right)=\frac{1}{2}\frac{1}{w_Iw_J},`$ | $`\left(Q_0V_4(w_I)\right)\left(\eta _0\xi c\varphi e^\varphi (w_J)\right)=\frac{1}{4}\frac{1}{(w_Iw_J)^2},`$ | | $`\left(Q_0V_4(w_I)\right)\left(\eta _0T(w_J)\right)=\frac{1}{4}\frac{1}{w_Iw_J},`$ | $`\left(Q_0\xi c\varphi e^\varphi (w_I)\right)\left(\eta _0\xi c\varphi e^\varphi (w_J)\right)=0,`$ | | $`\left(Q_0\xi c\varphi e^\varphi (w_I)\right)\left(\eta _0T(w_J)\right)=\frac{1}{2},`$ | $`\left(Q_0T(w_I)\right)\left(\eta _0\xi c\varphi e^\varphi (w_J)\right)=\frac{1}{2},`$ | | $`\left(Q_0T(w_I)\right)\left(\eta _0T(w_J)\right)=\frac{1}{2}(w_Iw_J).`$ | | Setting $`w_1=1`$ and $`w_2=1`$, we find $$2g^2S_{v_4v_5}=Y(1,2)Y(2,1)=3+3=6.$$ ### Level $`\frac{7}{2}`$ At this level we have terms of the form $`tav_\alpha `$, $`tev_\alpha `$, $`tfv_\alpha `$ and $`t^3ev_\alpha `$. The coefficients of the terms of the form $`t^3av_\alpha `$ and $`t^3fv_\alpha `$ vanish because of $`\varphi `$ momentum conservation. Here we will explain the calculation of the coefficient of the cubic term $`tav_1`$. The relevant term in the string field action which gives rise to cubic terms is $$S_{cubic}=\frac{1}{2g^2}\frac{1}{3}(\widehat{Q}_B\widehat{\mathrm{\Phi }})\widehat{\mathrm{\Phi }}(\widehat{\eta }_0\widehat{\mathrm{\Phi }}).$$ In terms of GSO(+) ($`\mathrm{\Phi }_+`$) and GSO($``$) ($`\mathrm{\Phi }_{}`$) terms in the string field action, the cubic part of the action after tracing over the $`\sigma `$ matrices becomes $$S_{cubic}=\frac{1}{3g^2}\left\{(Q_B\mathrm{\Phi }_+)\mathrm{\Phi }_{}(\eta _0\mathrm{\Phi }_{})+(Q_B\mathrm{\Phi }_{})\mathrm{\Phi }_+(\eta _0\mathrm{\Phi }_{})\right\}+\mathrm{}$$ (24) where $`(\mathrm{})`$ represent terms which are not relevant for the calculation of the coefficient of $`tav_1`$. The expansion of $`\mathrm{\Phi }_+`$ and $`\mathrm{\Phi }_{}`$ relevant here is $$\mathrm{\Phi }_+=aA,\mathrm{\Phi }_{}=tT+v_1V_1.$$ Substituting the above expansion in (24) we obtain, $`S|_{tav_1}=\frac{1}{3g^2}\{(Q_BT)A(\eta _0V_1)+(Q_BA)T(\eta _0V_1)(Q_BV_1)T(\eta _0A)`$ $`(Q_BT)V_1(\eta _0A)+(Q_BA)V_1(\eta _0T)+(Q_BV_1)A(\eta _0T)\}`$ To calculate the coefficient of $`tav_1`$ term we actually only need to calculate one correlator. This is partly due to the fact that in the correlators given above involving $`V_1`$ we can replace $`V_1`$ by a multiple of $`T`$. $`\left(Q_BV_1\left(w_I\right)\right)A\left(w_J\right)\left(\eta _0T\left(w_K\right)\right)`$ $`=`$ $`\left(f_I^{}\right)^{{\scriptscriptstyle \frac{3}{2}}}\left(f_J^{}\right)\left(f_K^{}\right)^{{\scriptscriptstyle \frac{1}{2}}}Q_B(V_1\left(w_I\right)+_{V_1}^IT\left(w_I\right))`$ $`(A\left(w_J\right)+𝒫_A^JU_A\left(w_J\right))\eta _0\left(T\left(w_K\right)\right),`$ $`=`$ $`\left(f_I^{}\right)^2^I\left(Q_BT\left(w_I\right)\right)A\left(w_J\right)\left(\eta _0T\left(w_K\right)\right),`$ $``$ $`\left(f_I^{}\right)^2^IC(I,J,K),`$ where in the second line we have used the fact that since neither $`A`$ nor $`T`$ have fields from the matter sector therefore the term with $`V_1`$ in the correlator vanishes. Similarly <sup>8</sup><sup>8</sup>8In this subsection we will suppress some of the indices on $`_{V_1}^{I,3}^I`$. $`\left(Q_BT(w_I)\right)A(w_J)\left(\eta _0V_1(w_K)\right)`$ $`=`$ $`(f_K^{})^2^KC(I,J,K),`$ where the factor $`^K`$ comes from the conformal transformation of $`V_1`$ and $`(f_K^{})^2`$ is needed to compensate for the difference in the conformal dimensions of $`T`$ and $`V_1`$. Using the fact that $`Q_2\eta _0(T)=Q_2\eta _0(V_1)=0`$ and $`Q_B`$ acts like a derivation it follows that, $`\left(Q_BA(w_I)\right)V_1(w_J)\left(\eta _0T(w_K)\right)`$ $`=`$ $`(Q_BV_1(w_J)A(w_I)\left(\eta _0T(w_K)\right),`$ $`\left(Q_BT(w_I)\right)V_1(w_J)\left(\eta _0A(w_K)\right)`$ $`=`$ $`\left(Q_BT(w_I)\right)A(w_K)\left(\eta _0V_1(w_J)\right),`$ $`\left(Q_BA(w_I)\right)T(w_J)\left(\eta _0V_1(w_K)\right)`$ $`=`$ $`\left(Q_BT(w_J)\right)A(w_I)\left(\eta _0V_1(w_K)\right).`$ Thus using the properties of $`Q_B`$ and $`\eta _0`$ we have reduced the calculation to one correlator $`g^2S|_{tav_1}=`$ $``$ $`\frac{1}{3}\{(f_3^{})^2^3C(1,2,3)(f_3^{})^2^3C(2,1,3)(f_1^{})^2^1C(1,3,2)`$ $``$ $`(f_2^{})^2^2C(1,3,2)(f_2^{})^2^2C(2,1,3)+(f_1^{})^2^1C(1,2,3)\}.`$ $`C(I,J,K)`$ $`=`$ $`\left(Q_BT(w_I)\right)A(w_J)\left(\eta _0T(w_K)\right)`$ , $`=`$ $`(f_I^{})^{\frac{1}{2}}(f_J^{})(f_K^{})^{\frac{1}{2}}\left(Q_BT(w_I)\right)[A(w_J)+𝒫_A^J\xi \xi cce^{2\varphi }(w_J)]\left(\eta _0T(w_K)\right),`$ $`=`$ $`(f_I^{})^{\frac{1}{2}}(f_J^{})(f_K^{})^{\frac{1}{2}}(2{\displaystyle \frac{w_{IK}}{w_{JK}}}𝒫_A^Jw_{IK}).`$ $$g^2S|_{tav_1}=\frac{50}{18}.$$ ### Tachyon Potential We now give the result of our calculation <sup>9</sup><sup>9</sup>9Exact coefficients of the terms $`t^3ev_\alpha `$ are sum of 40 correlators and have very long expressions in terms of radicals, therefore only their decimal expansion is given.. $`g^2S_0`$ $`=`$ $`\frac{1}{4}t^2\frac{1}{2}t^4,`$ $`g^2S_{{\scriptscriptstyle \frac{3}{2}}}`$ $`=`$ $`at^2\frac{1}{4}et^2\frac{5}{96}\sqrt{50+22\sqrt{5}}et^4,`$ $`g^2S_2`$ $`=`$ $`\frac{15}{4}t^3v_1\frac{17}{6}t^3v_2\frac{1}{4}t^3v_3+\frac{35}{12}t^3v_4\frac{11}{6}t^3v_5,`$ $`g^2S_3`$ $`=`$ $`2ae+5f^2\left(\frac{1}{\sqrt{2}}\frac{1}{24}\right)e^2t^2+\frac{5}{18}e^2t^4+\frac{5}{4}(4\sqrt{2}1)f^2t^2`$ $`+`$ $`\frac{1}{12}(\mathrm{\hspace{0.17em}3}+40\sqrt{2})aet^2\frac{5}{12}(\mathrm{\hspace{0.17em}10}\sqrt{2}1)eft^2,`$ $`g^2S_{{\scriptscriptstyle \frac{7}{2}}}`$ $`=`$ $`\frac{25}{9}tav_1+\frac{44}{27}tav_2+\frac{44}{27}tav_3\frac{329}{27}tav_4+\frac{64}{9}tav_5`$ $``$ $`\frac{25}{36}tev_1+\frac{35}{27}tev_2\frac{17}{18}tev_3\frac{41}{108}tev_4+\frac{8}{27}tev_5`$ $``$ $`\frac{80}{9}tfv_1\frac{80}{27}tfv_3+\frac{160}{9}tfv_4\frac{320}{27}tfv_5`$ $`+`$ $`4.35732t^3ev_14.77364t^3ev_2+.605194t^3ev_3+4.351715t^3ev_4`$ $``$ $`1.94348t^3ev_5,`$ $`g^2S_4`$ $`=`$ $`\frac{45}{8}v_{1}^{}{}_{}{}^{2}+3v_{2}^{}{}_{}{}^{2}+\frac{3}{4}v_{3}^{}{}_{}{}^{2}\frac{15}{8}v_{4}^{}{}_{}{}^{2}\frac{3}{2}v_{5}^{}{}_{}{}^{2}+3v_4v_5`$ $``$ $`\frac{945}{64}t^2v_{1}^{}{}_{}{}^{2}\frac{35}{6}t^2v_{2}^{}{}_{}{}^{2}+\frac{5}{4}t^2v_{3}^{}{}_{}{}^{2}\frac{3511}{192}t^2v_{4}^{}{}_{}{}^{2}\frac{161}{24}t^2v_{5}^{}{}_{}{}^{2}`$ $`+`$ $`\frac{255}{16}t^2v_1v_2+\frac{45}{32}t^2v_1v_3\frac{525}{32}t^2v_1v_4+\frac{165}{16}t^2v_1v_5+\frac{19}{48}t^2v_2v_3`$ $`+`$ $`\frac{991}{48}t^2v_2v_4\frac{143}{12}t^2v_2v_5\frac{65}{32}t^2v_3v_4+\frac{7}{8}t^2v_3v_5+\frac{985}{48}t^2v_4v_5.`$ The potential is given by $$V(t,a,e,f,v_i)=S(t,a,e,f,v_i)=2\pi ^2Mg^2S(t,a,e,f,v_i),$$ where $`M=\frac{1}{2\pi ^2g^2}`$. The potential has a minimum for the following values of $`t,a,e,f,v_i`$. $`t=\pm 0.603455,a=0.059291,e=0.030980,f=0.015746,v_1=\pm 0.035688,`$ $`v_2=\pm 0.0571190,v_3=\pm 0.0153737,v_4=\pm 0.031538,v_5=0.006012.`$ The value of the potential at the minimum is $$V=0.90454M.$$ The expected answer for the value of the potential at the minimum is $`M`$. Thus we see that the level four approximation produces $`90.5\%`$ of the expected value. We can integrate out the fields $`a,e,f,`$ and $`v_\alpha `$ to obtain an effective potential for the tachyon which is given by $$V_{effective}=2\pi ^2M\frac{P(t)}{Q(t)},$$ where $`P(t)`$ $`=`$ $`\mathrm{0.8168083768\hspace{0.17em}10}^5t^2+0.0002047077013t^4+0.002821258794t^6`$ $`+`$ $`0.02594499700t^8+0.1738511409t^{10}+0.8795191385t^{12}`$ $`+`$ $`3.409051376t^{14}+10.11847495t^{16}+22.58030878t^{18}`$ $`+`$ $`35.88635644t^{20}+33.50595320t^{22}4.683061411t^{24}`$ $``$ $`77.91238815t^{26}149.1412411t^{28}168.7145393t^{30}`$ $``$ $`125.0599362t^{32}57.59279282t^{34}12.56479811t^{36},`$ and $`Q(t)=0.00003267233507+0.0008678393076t^2+0.01254363691t^4`$ $`+`$ $`0.1215742201t^6+0.8648433845t^8+4.706706334t^{10}`$ $`+`$ $`20.03806794t^{12}+67.54526560t^{14}+181.3752555t^{16}`$ $`+`$ $`388.8655996t^{18}+664.8883542t^{20}+902.1103038t^{22}`$ $`+`$ $`959.2815429t^{24}+778.5205080t^{26}+459.3896684t^{28}`$ $`+`$ $`177.3949239t^{30}+33.42209885t^{32}.`$ ## Acknowledgements We would like to thank Barton Zwiebach and Leonardo Rastelli for valuable discussions and P. De Smet and J. Raemaekers for useful correspondence. This research was supported in part by the US Department of Energy under contract #DE-FC02-94ER40818. ## 4 Appendix Conformal Transformations: $`fT(0)`$ $`=`$ $`(f^{}(0))^{\frac{1}{2}}T(w)`$ $`fA(0)`$ $`=`$ $`f^{}(0)\{A(w)\frac{f^{\prime \prime }(0)}{(f^{}(0))^2}cc\xi \xi e^{2\varphi }(w)\}`$ $`fE(0)`$ $`=`$ $`f^{}(0)\{E(w)\frac{f^{\prime \prime }(0)}{2(f^{}(0))^2}\}`$ $`fF(0)`$ $`=`$ $`f^{}(0)F(w)`$ $`fV_1(0)`$ $`=`$ $`(f^{}(0))^{\frac{3}{2}}\{V_1(w)\frac{15}{12}\{z,f\}\xi ce^\varphi (w)\}`$ $`fV_2(0)`$ $`=`$ $`(f^{}(0))^{\frac{3}{2}}\{V_2(w)\frac{f^{\prime \prime }(0)}{(f^{}(0))^2}\xi ce^\varphi (w)+(2\frac{(f^{\prime \prime }(0))^2}{(f^{}(0))^4}\frac{f^{\prime \prime \prime }(0)}{(f^{}(0))^3})\xi ce^\varphi (w)\}`$ $`fV_3(0)`$ $`=`$ $`(f^{}(0))^{\frac{3}{2}}\{V_3(w)+\frac{1}{6}\{z,f\}\xi ce^\varphi (w)+\frac{f^{\prime \prime }(0)}{2(f^{}(0))^2}\xi ce^\varphi \}`$ (25) $`fV_4(0)`$ $`=`$ $`(f^{}(0))^{\frac{3}{2}}\{V_4(w)\frac{1}{2}\frac{f^{\prime \prime }(0)}{(f^{}(0))^2}\xi c\varphi e^\varphi +(\frac{1}{12}\frac{f^{\prime \prime \prime }(0)}{(f^{}(0))^3}\frac{2}{3}\{z,f\})\xi ce^\varphi (w)\}`$ $`fV_5(0)`$ $`=`$ $`(f^{}(0))^{\frac{3}{2}}\{V_5(w)+\frac{f^{\prime \prime }(0)}{(f^{}(0))^2}\xi c\varphi e^\varphi +(\frac{1}{2}\{z,f\}\frac{1}{6}\frac{f^{\prime \prime \prime }(0)}{(f^{}(0))^3})\xi ce^\varphi \}`$ $`𝒫_A^I`$ $`=`$ $`\frac{f_I^{\prime \prime }(0)}{(f_I^{}(0))^2},𝒫_E^I=\frac{f_I^{\prime \prime }(0)}{2(f_I^{}(0))^2},𝒫_F^I=0`$ (26) $`𝒫_{V_1}^I`$ $`=`$ $`0,𝒫_{V_2}^I=\frac{f_I^{\prime \prime }(0)}{(f_I^{}(0))^2},𝒫_{V_3}^I=\frac{f_I^{\prime \prime }(0)}{2(f_I^{}(0))^2},𝒫_{V_4}^I=\frac{f_I^{\prime \prime }(0)}{2(f_I^{}(0))^2},𝒫_{V_5}^I=\frac{f_I^{\prime \prime }(0)}{(f_I^{}(0))^2},`$ (27) $`_{V_1}^I`$ $`=`$ $`\frac{15}{12}\{z,f\},_{V_2}^I=2\frac{(f^{\prime \prime }(0))^2}{(f^{}(0))^4}\frac{f^{\prime \prime \prime }(0)}{(f^{}(0))^3},_{V_3}^I=\frac{1}{6}\{z,f\},`$ (28) $`_{V_4}^I`$ $`=`$ $`\frac{1}{12}\frac{f^{\prime \prime \prime }(0)}{(f^{}(0))^3}\frac{2}{3}\{z,f\},_{V_5}^I=\frac{1}{6}\frac{f^{\prime \prime \prime }(0)}{(f^{}(0))^3}+\frac{1}{2}\{z,f\},`$ (29) BRST transformations: $`Q_B={\displaystyle 𝑑zj_B(z)}={\displaystyle 𝑑z\left\{c(T_m+T_{\xi \eta }+T_\varphi )+ccb+\eta e^\varphi G_m\eta \eta e^{2\varphi }b\right\}},`$ $`Q_0`$ $``$ $`{\displaystyle }dz\{c(T_m+T_{\xi \eta }+T_\varphi )+ccb,`$ $`Q_1`$ $``$ $`{\displaystyle 𝑑z\eta e^\varphi G_m},Q_2{\displaystyle 𝑑z\eta \eta e^{2\varphi }b}.`$ $$U_2=\xi ce^\varphi ,U_3=\xi ce^\varphi ,U_4=U_5=\xi c\varphi e^\varphi .$$ $`Q_0`$: $`Q_0(T)`$ $`=`$ $`\frac{1}{2}\xi cce^\varphi ,`$ $`Q_0(A)`$ $`=`$ $`2cc^2c\xi \xi e^{2\varphi },`$ $`Q_B(E)`$ $`=`$ $`c\xi \eta c\eta \xi c\eta \xi \frac{1}{2}^2c+\eta e^\varphi G_m+2\eta \eta e^{2\varphi }b,`$ $`Q_B(F)`$ $`=`$ $`c\xi cG_me^\varphi ,`$ $`Q_0(V_1)`$ $`=`$ $`\frac{c_m}{2}\frac{1}{3!}^3c\xi ce^\varphi +\frac{3}{2}c\xi cT_me^\varphi ,`$ $`Q_0(V_2)`$ $`=`$ $`\xi c^2ce^\varphi \frac{3}{2}\xi c^2ce^\varphi \xi c^3ce^\varphi +c\xi ^2ce^\varphi ,`$ $`Q_0(V_3)`$ $`=`$ $`\frac{c_{\xi \eta }}{2}\frac{1}{3!}^3c\xi ce^\varphi +\frac{3}{2}cc\xi T_{\xi \eta }e^\varphi +\frac{1}{2}c^2c\xi e^\varphi ,`$ $`Q_0(V_4)`$ $`=`$ $`\frac{9}{12}^3c\xi ce^\varphi +\frac{3}{2}c\xi T_\varphi ce^\varphi +\frac{1}{2}\xi ^2cc\varphi e^\varphi ,`$ $`Q_0(V_5)`$ $`=`$ $`\frac{2}{3}^3c\xi c+\xi c^2c\varphi e^\varphi +\frac{3}{2}\xi cc^2\varphi e^\varphi ,`$ $`Q_0(U_A)`$ $`=`$ $`0,`$ $`Q_0(U_2)`$ $`=`$ $`(c\xi ce^\varphi ),`$ $`Q_0(U_3)`$ $`=`$ $`\frac{1}{2}\xi cce^\varphi ,`$ $`Q_0(U_4)`$ $`=`$ $`\frac{1}{2}cc\xi \varphi e^\varphi \frac{1}{2}c^2c\xi e^\varphi .`$ $$c_m=15,c_{\xi \eta }=2.$$ $`Q_1`$: $`Q_1(F)`$ $`=`$ $`10\eta \xi c10\eta \xi c(\varphi )2T_mc\varphi 5(^2\varphi +(\varphi )^2)c,`$ $`Q_1(V_1)`$ $`=`$ $`\frac{3}{2}\xi \eta cG_m\frac{3}{2}cG_m\varphi \frac{1}{2}cG_m,`$ $`Q_1(V_2)`$ $`=`$ $`0,`$ $`Q_1(V_3)`$ $`=`$ $`\xi \eta cG_m,`$ $`Q_1(V_4)`$ $`=`$ $`\frac{3}{2}G_m\eta \xi c\frac{5}{2}G_m\varphi c\frac{3}{2}G_mc,`$ $`Q_1(V_5)`$ $`=`$ $`G_m\eta \xi c+G_mc+G_m\varphi c,`$ $`Q_1(U_2)`$ $`=`$ $`0,`$ $`Q_1(U_3)`$ $`=`$ $`G_mc,`$ $`Q_1(U_4)`$ $`=`$ $`G_mc.`$ $`Q_2`$: $`Q_2(T)`$ $`=`$ $`\eta e^\varphi ,`$ $`Q_2(V_1)`$ $`=`$ $`\eta e^\varphi T_m,`$ $`Q_2(V_2)`$ $`=`$ $`2\eta \eta \xi e^\varphi 3^2\eta e^\varphi 8\eta \varphi e^\varphi 2\eta ^2\varphi e^\varphi 4\eta (\varphi )^2e^\varphi ,`$ $`Q_2(V_3)`$ $`=`$ $`\eta bce^\varphi +2\eta bc(\varphi )e^\varphi +3\eta \xi \eta e^\varphi +\eta ^2\varphi e^\varphi +2\eta (\varphi )^2e^\varphi ,`$ $`Q_2(V_4)`$ $`=`$ $`4\eta \eta \xi e^\varphi +8\eta bce^\varphi +4\eta bce^\varphi +10\eta bc\varphi e^\varphi ,`$ $`+`$ $`6^2\eta e^\varphi +20\eta \varphi e^\varphi +5\eta ^2\varphi e^\varphi +\frac{25}{2}\eta (\varphi )^2e^\varphi ,`$ $`Q_2(V_5)`$ $`=`$ $`2\eta \eta \xi e^\varphi 4\eta bce^\varphi 2\eta bce^\varphi 4\eta bc\varphi e^\varphi 3^2\eta e^\varphi 8\eta \varphi e^\varphi ,`$ $``$ $`3\eta ^2\varphi e^\varphi 4\eta (\varphi )^2e^\varphi ,`$ $`Q_2(U_3)`$ $`=`$ $`2\eta bce^\varphi 3\eta e^\varphi 4\eta \varphi e^\varphi ,`$ $`Q_2(U_2)`$ $`=`$ $`2(\eta e^\varphi ),`$ $`Q_2(U_4)`$ $`=`$ $`2\eta bce^\varphi 5\eta \varphi e^\varphi 4\eta e^\varphi .`$
warning/0004/cond-mat0004090.html
ar5iv
text
# Possible Localized Modes in the Uniform Quantum Heisenberg Chains of Sr2CuO3. \[ ## Abstract A model of mobile bond-defects is tentatively proposed to analyze the “anomalies” observed on the NMR spectrum of the quantum Heisenberg chains of Sr<sub>2</sub>CuO<sub>3</sub>. A bond-defect is a local change in the exchange coupling. It results in a local alternating magnetization (LAM), which, when the defect moves, creates a flipping proces of the local field seen by each nuclear spin. At low temperature, when the overlap of the LAM becomes large, the defects form a periodic structure, which extends over almost all the chains. In that regime, the density of bond-defects decreases linearly with T. \] In low-dimensional quantum antiferromagnets, the effects of impurities or defects in “gapped” and “ungapped” systems may display drastic differences but also common features. A good example of the latter case is provided by the uniform quantum ($`s=1/2`$) Heisenberg chains (UQHC). Hereafter, we consider the effects due to defects in such systems. A “defect” is defined as a local change in the magnetic bond coupling. As the translational symmetry of the spin system is broken, a local alternating magnetization (LAM) develops around the defects in the direction of the applied field: this is one of the common features mentioned above. Such a LAM is probed accurately by NMR. Our analysis relies on results obtained on the compound Sr<sub>2</sub>CuO<sub>3</sub>. In such a “real system”, the elastic coupling between the spins and the lattice cannot be ignored. It can result in a lattice distortion giving rise to the so-called spin-Peierls transition (at temperature $`T_{sp}`$). Dynamical effects on LAM structures can also been expected from such magneto-elastic couplings. In the dimerized (D) phase of a spin-Peierls system (SPS), i.e., for $`T<T_{sp}`$, the LAM induced by chain-end effects become mobile in the pure one-dimensional case. In the “high-field” phase of a SPS, a periodic LAM structure occurs, which results from the field-induced incommensurate lattice modulation. In that case, dynamical effects are due to the quantum fluctuations of the phason modes describing the vibrations of the modulated lattice structure. In the uniform phase of a SPS, i.e. for $`T>T_{sp}`$, the spin-lattice coupling is also of crucial importance. It can appreciably change the magnetic susceptibility of the system well above $`T_{sp}`$. In Sr<sub>2</sub>CuO<sub>3</sub>, no spin-Peierls transition is observed as a 3-dimensional magnetic ordering takes place at $`T_N5`$ $`K`$. Above $`T_N`$, this compound provides a remarkable example of UQHC. The exchange coupling is large ($`J2200`$ K). The logarithmic corrections characterizing the quantum ($`T<<J`$) behavior of UQHC have been observed. LAM associated with the edges of finite chains have also been observed in Sr<sub>2</sub>CuO<sub>3</sub>. Due to that LAM, the NMR lineshape is changed in a very specific way, giving rise to “features”, which have been well identified. This is the case of features “A” displayed in Figs. 1 and 2. Other features, however, are visible on the NMR line (features “B” and “C” in Figs. 1 and 2), which have not been explained yet. We propose an analysis of these additional features in terms of LAM associated with mobile “bond-defects”. The NMR spectra to be analyzed - a few examples are displayed in Figs. 1 and 2 - have been obtained on an Ar-annealed single crystal. That annealing procedure minimizes in that compound the possible source of impurities induced by interstitial excess oxygen. After such a treatment, the concentration of the residual spin-1/2 impurities becomes as low as $`1.3`$ $`10^4`$ . Some of the NMR spectra displayed in Figs. 1 and 2 have been previously reported (Figs. 1a, b and Fig. 2a) in together with the experimental conditions. As clearly established in, features “A” originate from the chain edges, while we show here that features “B” is well explained by the presence of mobile bond-defects. Features “C” would result from the periodic structure they form at low $`T`$. Finally, possible origins of these defects are suggested in relation with the lattice properties. First we consider features “A”. LAM associated with the edges of finite UQHC have been described by Eggert and Affleck. The LAM amplitude increases with distance from the chain end. At $`T0`$ however, thermal fluctuations act as a cutoff, and at long distance, the LAM amplitude decreases exponentially with a characteristic scale given by the correlation length of the system: $`\xi =J/2T`$. An analytical expression for the local susceptibility has been derived, which agrees well with Monte-Carlo calculations. Accordingly, for an hyperfine coupling $`A_\beta `$ (expressed in GHz and $`\beta `$ = a, b or c referring to the field direction with respect to the crystal axes - chain axis b), the local field seen by a nuclear spin at site n (counted from a chain end) can be written $`h_n=(1)^nB\text{ }A_\beta \text{ }H\text{ }\{(n+\varphi )/\sqrt{\xi \mathrm{sinh}(2(n+\varphi )/\xi }(1)`$ with $`B0.020`$ GHz<sup>-1</sup> and $`h_n`$ and $`H`$ expressed in Tesla. In (1), $`\varphi `$ allows for a small shift of the LAM along the chain.$`\varphi `$ changes the position $`n_{max}`$(see Fig. 3a) of the maximum field value, but for $`\varphi `$ $`n_{max}`$, it does not affect this value:$`h_{max}1.0810^2A_\beta H\sqrt{\xi }`$ ($`A_a0.14`$ GHz and $`A_c0.50`$ GHz). In the NMR spectrum, the field splitting $`\mathrm{\Delta }H`$ (see Fig. 1) where features “A” occur is given by $`\mathrm{\Delta }H=2h_{max}`$. In that respect, as explained in , a very good agreement is obtained between theory and experiment. The shape of features “A”, however, depends appreciably on $`\varphi `$ (see dot line $`\varphi =0`$ in Fig. 1b). Using $`\varphi `$ as an adjustable parameter, a good agreement with the experimental line shape can be achieved for all the collected data (see dots in Figs. 1 and 2). With that procedure, one obtains that $`\varphi `$ is field independent. It shows, however, a $`T`$ dependence: $`\varphi =(1200\pm 200)/T`$, which, remarkably, agrees with $`\varphi \xi `$. As shown in Fig. 3a (solid line), for that particular value of $`\varphi `$, the LAM displays no maximum, but a rather “flat” initial behavior. Despite the rough description used here, one is led to conclude that the initial increase of the LAM amplitude is not observed experimentally, in contradiction with the theoretical prediction. The elastic spin-lattice coupling may play here an important role. The model of a sudden cutoff of the exchange coupling (J = 0) just at the chain edges would not applied to such “real” systems. Instead, as in the D phase of a SPS, on a length scale $``$ $`\xi `$, a distribution of the J-bonds would take place, giving rise to the observed “smoothing” of the initial LAM amplitude. Second, we consider features “B”. We assume that they result from a local change in the exchange coupling ($`J^{}J`$). As studied by Eggert (for $`J^{}<J`$), LAM similar to the case $`J^{}=0`$ develop around such bond-defects. We suppose that the associated LAM can be described by an expression similar to (1), but with a multiplying factor $`\alpha `$ ($`\alpha `$ $`<`$1) to account for the expected smaller amplitude. As before, a parameter $`\varphi ^{}`$ describes a possible “smoothing” effect around the defect position. Finally, we allow these defects to move along the chains. This dynamical behavior is of crucial importance for the NMR lineshape as it results in a “flipping” process of the hyperfine local field $`h_n^{}`$ seen by the nuclear spins. Accordingly, the NMR signal of each nuclear spin is composed of 1 or 2 distinct lines depending on the flipping rate G with respect to the local hyperfine field $`h_n^{}`$. That “motional narrowing” is described as follows. For a nuclear spin at site n in the chain, the NMR spectrum is written $`I_n=\left[D_0+K^{}\left(\nu \right)\right]/\left\{\left[\nu +K^{\prime \prime }\left(\nu \right)\right]^2+D_0^2+K^{}\left(\nu \right)^2\right\}\text{ }\left(2\right)`$ with $`K^{}\left(\nu \right)=Re\left\{K\left(\nu \right)\right\}`$, $`K^{\prime \prime }\left(\nu \right)=Im\left\{K\left(\nu \right)\right\}`$ and $`\nu =\nu ^0\gamma H/2\pi `$, where $`\nu ^0`$ is the experimental frequency and $`D_0`$ is the “natural” NMR width. $`K(\nu )`$ describes the effect of the dynamical local field as $`K(\nu )=(\gamma h_n^{}/2\pi )^2k(\nu )`$. Here, $`k(\nu )`$ is the Laplace representation of $`k(t)`$ which, as a function of time $`t`$, describes the flipping process due to the motion of the defects. A ballistic motion results in a linear exponential, $`k(t)=exp(t/G)`$, while a diffusive behavior gives $`k(t)=exp(\sqrt{t/G})`$. Such a single moving defect in a long chain may create additional “features” in the total NMR line, i.e., $`I=\mathrm{\Sigma }_nI_n`$. That model allows us to reproduce very well the observed features “B”. An example is given in Fig. 3b, where a comparison with the data obtained for $`H`$ // a, at $`\nu ^0=8710^3`$ GHz and $`T=30`$ K is presented. The left side corresponds to a diffusive behavior with the parameters: $`\varphi ^{}=20`$, $`\alpha =0.6`$, $`G=2.810^4`$ GHz and $`D_0=1.510^5`$ GHz. The ballistic model (right side) gives usually a slightly poorer agreement (here with $`\alpha =0.45`$, $`G=1.510^4`$ GHz). That procedure applied to the different experimental conditions yields an interesting result concerning $`\varphi ^{}`$. This parameter is independent on $`H`$, but it depends on $`T`$ according to $`\varphi ^{}=(600\pm 100)/T\xi /2`$. As for $`\varphi `$, this smoothing effect can be assumed to result from a local induced lattice distortion around the bond-defects. In presence of several such bond-defects, the above description applies as long as the average distance between the defects remains large compared to $`\xi `$, i.e., a model of independent defects. At low $`T`$, however, interactions between defects are expected: a magnetic interaction should result from the overlap of the LAM structures and an elastic interaction from the lattice distortion associated with the defects. As a balance between these two interactions, we assume now that the defects form a periodic structure. If it is the case, the shape of the resulting LAM will depend on the number $`l`$ of spins between two neighboring bond-defects. From the dashed line in Fig. 3a, approximate descriptions of such a LAM are represented in Fig. 3c, for $`l`$ odd and even, respectively. As expected, for $`l`$ even, a node occurs at the middle position. For $`l`$ odd, the LAM amplitude remains finite at all positions. In the former case, the resulting NMR line displays a single narrow peak (Fig. 3d, right side). In the latter case, however, as the local field seen by the nuclear spins never cancels, a double structure develops at the centre of the NMR line (left side). For $`H`$ // a , in particular (see Fig. 1), such a central double structure is clearly observed below $`30`$ K (and well reproduced by our model). For $`H`$ // c, the NMR line observed at $`T=30`$ K (Fig. 2a) is more complex. In addition to the overlap of the three components of the quadrupolar splitting, it can be analyzed as composed of two contributions as illustrated in Fig. 2b. For one contribution (the dotted line), each component displays a narrow central peak (similar to the case described in Fig. 3d, right, for $`l`$ even) and for the other, it displays a central double structure (as in Fig. 3d, left, for $`l`$ odd). At lower $`T`$, however, only the double peak structure (features “C”) remains (see Fig. 2c). The contribution from segments with $`l`$ even is seen to decrease rapidly with $`T`$, and one is led to conclude that a periodic structure with $`l`$ odd becomes energetically preferable. A complete description of the NMR line at low $`T`$ can now be proposed. A chain of $`N`$ spins is considered, which contains a number $`\eta `$ of bond-defects forming a periodic LAM structure with $`l`$ odd. The dynamical behavior, i.e., the function $`K(\nu )`$ in Eq. 2, is now to be considered as describing the dynamics (or the thermal fluctuations) of that whole periodic LAM structure. To account for the additional static local fields induced by the chain edges, expressions for $`h_n`$ and $`h_{Nn}`$ \[given by Eq. (1)\] are added to the applied field $`H`$ in Eq. 2. The total NMR line, $`I=\mathrm{\Sigma }_n`$ $`I_n`$, is then calculated by varying the parameters $`\alpha `$, $`G`$, $`l`$ and $`D_0`$. The parameters $`\alpha `$ and $`G`$ are used to adjust the width of feature “B”. The shape of feature “C” is then determined by varying $`l`$ and $`D_0`$, keeping in mind that the narrowing effect, i.e., the parameter $`G`$, plays here a crucial role on the occurrence or not of such a small splitting at the centre of the NMR line. Finally the number $`\eta `$ of bond-defects (with the condition $`\eta l<N`$) is used to adjust the intensity of features “A” relatively to that of the main line. With that procedure, one obtains the open dots in Figs. 1 and 2. The agreement with the experiments is in general excellent. Particular confidence in this approach is provided by the high sensitivity of the final agreement to only a few parameters, essentially $`l`$ and $`G`$ (compare the different lines in Fig. 3d, left). At low temperature, when features “C” are clearly visible, one always obtains the following determination: $`\eta lN`$ ($`1800\pm 200`$). This means that the periodic LAM structure develops practically over all the chain length. In that case, an evaluation of the density of the bond-defects is simply given by $`\rho l^1`$. For $`H`$ // a, one obtains a rather accurate determination of the parameters $`\alpha `$ and $`G`$. They do not depend appreciably on $`T`$ and $`H`$: $`\alpha =0.60\pm 0.04`$, $`G=(2.8\pm 0.2)`$ $`10^4`$ GHz. For $`H`$ // c, the determination is much more uncertain ($`\alpha 0.46`$ and $`G4.2`$ $`10^4`$ GHz) as a small contribution with $`l`$ even is ignored in our description. The parameter $`D_0`$ remains also constant ($`D_01.5`$ $`10^5`$ GHz for $`H`$ // a, and $`D_05.6`$ $`10^5`$ GHz for $`H`$ // c), except at the lowest temperature ($`20`$ K) where larger values (by a factor $`2`$) provide a better agreement. At low $`T`$, however, the description represented in Fig. 3c becomes insufficient when the overlap of the LAM becomes very large. The main variation is finally observed on the density of bond-defects $`\rho l^1`$, which decreases almost linearly with $`T`$. This result is well established for $`H`$ // a (Fig. 2d). The values obtained for $`H`$ // c are also shown in that figure, though they are more uncertain for the reason explained above. A first question raised by the present description concerns the source of the bond-defects ($`J^{}<J)`$. Are they related to the interstitial excess oxygen which characterizes that compound? There are two reasons to rule out that possibility. First, at the lowest temperature, $`T=20`$ K, the density of bond-defects ($`\rho 2.8`$ $`10^3`$) remains more than ten times larger than the concentration of the residual spin-1/2 impurities ($`1.3`$ $`10^4`$) after annealing. Second, the observed T dependence of $`\rho `$ tells us that the number of bond-defects is not a constant but it decreases with T. Remarkably, $`\rho `$ is observed to decrease linearly with T, as does the inverse of the spin correlation length $`\xi ^12T/J`$. That last remark strongly supports our assumption that the bond-defects are an intrinsic property of the “real” spin chains in Sr<sub>2</sub>CuO<sub>3</sub>. In the spin-Peierls phenomenology - and a priori above $`T_N`$, Sr<sub>2</sub>CuO<sub>3</sub> is a SPS in its U phase - the description of the lattice is usually considered in its linear approximation. The lattice modes are described by simple phonon branches (characteristic frequency $`\nu _p`$) and the spin-lattice coupling is considered in its adiabatic limit ($`\nu _p<J`$). A more realistic description, however, should take into account the non-linear properties of the lattice. They may give rise to additional “localized” (and rather mobile) excitations (lattice solitons, for instance). In the case of uniform chains, one kind of such non-linear excitations are of particular interest: the so-called “localized modes” as defined in . They correspond to local oscillations of atoms, at high frequency and with large amplitude (breather like). Their presence in the lattice of an antiferromagnetic chain would result in bond-defects as those discussed above. Alternatively, if the SP phenomenology is considered in its non-adiabatic limit ($`\nu _p>J`$), non-linearities might be also expected from the spin-lattice coupling itself. In that case, as discussed recently, the SP transition, instead of being driven by a phonon softening, would be associated with a “critical” central peak, which usually signals the presence of non-linear excitations. Well above $`T_{sp}`$, could this effective non-linearity explain the presence of localized excitations? The present NMR results may provide the first evidence of magneto-elastic localized modes in UQHC. A recent numerical investigation of such bond-defects in UQHC has already established the following results: i) the expected magnetic interactions between two (static) bond-defects are attractive, ii) a LAM structure with $`l`$ odd is energetically more favorable, iii) the LAM amplitude of an l-even structure decreases rapidly at low temperature. All these results agree with our analysis of the NMR line. Together with the possible repulsive interactions due to the lattice, they reinforce our proposition that a periodic LAM structure takes place at low $`T`$ in UQHC. At that point, it is worth reminding that a double-peak structure, similar to our features “C”, has been observed in the SPS CuGeO<sub>3</sub> well above $`T_{sp}`$, i.e., in its U phase. Our conclusion is that the presence of a LAM structure is strongly supported by the NMR results. We propose a model - bond-defects and a periodic LAM structure - which explains both features “B” and “C” in an unique consistent model. That model deserves certainly to be comforted experimentally and theoretically. For instance, the quasi-static periodic AF structure of the LAM could be observed by neutron diffraction. Any description of such a “real” UQHC should consider explicitly the subtle relations between the spins and the lattice. One of us (JPB) thanks G. Uhrig and T. Ziman for illuminating discussions, and S. Miyashita and the authors of Ref. for communicating their results. Y. Ajiro is acknowledged for his invitation at Kyushu University, and the Japan Society for the Promotion of Science for its financial support.
warning/0004/hep-th0004188.html
ar5iv
text
# 1 Introduction ## 1 Introduction In recent years an interesting idea has been advocated to regard our world as a domain wall embedded in higher dimensional spacetime . Most of the particles in the standard model should be realized as modes localized on the wall. Phenomenological implications of the idea have been extensively studied from many aspects. Another fascinating possibility has also been proposed to consider walls in the bulk spacetime which has negative cosmological constant . The model can give large mass hierarchy or can give massless graviton localized on the wall. Subsequently a great deal of research activity has been performed to study and extend the proposal . Since walls typically have co-dimension one, it is desirable to consider intersections and/or junctions of walls in order to obtain our four dimensional world from a spacetime with much higher dimensions. The model with the bulk cosmological constant has been extended to produce an intersection of walls . Supersymmetry has been useful to achieve stability of solitonic solutions such as domain walls. Domain walls in supersymmetric theories can saturate the Bogomol’nyi bound . Such a domain wall preserves half of the original supersymmetry and is called a $`1/2`$ BPS state . It has also been noted that these BPS states possess a topological charge which becomes a central charge $`Z`$ of the supersymmetry algebra . Thanks to the topological charge, these BPS states are guaranteed to be stable under arbitrary local fluctuations. Various properties of domain walls in $`𝒩=1`$ supersymmetric field theories in four dimensions have been extensively studied , . In particular the modes on the domain wall background have been worked out and are found to contain fermions and/or bosons localized on the wall in many cases , . Recently domain wall junctions have attracted much attention as another interesting possibility for BPS states . Domain walls occur in interpolating two discrete degenerate vacua in separate region of space. If three or more different discrete vacua occur in separate region of space, segments of domain walls separate each pair of the neighboring vacua. If the two spatial dimensions of all of these domain walls have one dimension in common, these domain walls meet at a one-dimensional junction. The solitonic configuration for the junction can preserve a quarter of supersymmetry and is called a $`1/4`$ BPS state. There has been progress to study general properties of such domain wall junctions. For instance a new topological charge $`Y`$ is found to appear for such a $`1/4`$ BPS state . If we start from $`𝒩=1`$ four dimensional supersymmetric field theories, the domain wall junction preserves only one supercharge. Consequently the resulting theory was expected to be a $`(1,0)`$ supersymmetric theory in $`1+1`$ dimensions which offers an intriguing possibility of chiral fermions. Moreover, there have been a number of numerical simulations which indicate the existence of the domain wall junction solutions . In spite of all these efforts, it has been difficult to obtain an explicit solution and to prove the existence of a BPS domain wall junction. Recently we have succeeded to work out an exact solution for the BPS domain wall junction for the first time . The exact solution allows a thorough study of the properties of the BPS domain wall junction. Consequently several misconceptions can be pointed out and rectified. One such point is the sign and meaning of the new central charge $`Y`$ which arises when walls form a junction. Our exact solution showed that the central charge $`Y`$ contributes negatively to the mass of the domain wall junction configuration. Therefore we should not consider the central charge $`Y`$ alone as the mass of the junction. Various other aspects of the domain wall junctions are also studied recently . The purpose of the present paper is to give a more detailed study of the properties of the BPS domain wall junction in $`𝒩=1`$ supersymmetric field theories. We study the modes on the background of the domain wall junction, especially the Nambu-Goldstone modes. We will use our exact solution as a concrete example and will extract the generic properties of the BPS domain wall junctions. We define mode equations and demonstrate explicitly that fermion and boson with the same mass have to come in pairs except massless modes. Massless modes can appear singly without accompanying fields with opposite statistics. We also show that unitary representations of the surviving $`(1,0)`$ supersymmetry are classified into doublets for massive modes and singlets for massless modes. We work out explicitly massless Nambu-Goldstone modes associated with the broken supersymmetry and translational invariance. We find that the Nambu-Goldstone fermions exhibit an interesting chiral structure in accordance with the surviving $`(1,0)`$ supersymmetry algebra. However, we also find that any linear combinations of the Nambu-Goldstone modes associated with the junctions become a linear combination of zero modes on at least one of the domain walls asymptotically along these walls. Since their wave functions are extended along these walls without damping, they are not localized states on the junction. Therefore they are not normalizable, contrary to a previous expectation . This indicates that the resulting theory cannot be regarded as a genuine $`1+1`$ dimensional field theory with discrete particle spectrum even at zero energy. Although the remaining supersymmetry is just $`(1,0)`$ which is characteristic to $`1+1`$ dimensions, we have to keep in mind that the domain wall junction configuration is actually living in one more dimensions similarly to the domain wall itself. Zero modes on the junction turn out to have properties quite similar to those on the domain wall. The non-normalizability of Nambu-Goldstone modes on the junction configuration is not an accident in this particular model. We observe that the origin of this property can be traced back to the fact that the supersymmetry is broken by the coexistence of nonparallel walls. Therefore the fact that the Nambu-Goldstone modes on the BPS domain wall junction are not normalizable is a generic feature of supersymmetric field theories in the bulk flat space. One should note that our conclusion need not apply to the case with negative cosmological constant in the bulk. In the presence of a bulk negative cosmological constant in six dimensions, five dimensional walls can intersect in Anti de Sitter space. If one demands a flat space at the four dimensional intersection, one has an Anti de Sitter space not only in the bulk but also even on the walls . Since Anti de Sitter space does not have translational invariance, the wave function of the zero mode does not become constant along the wall asymptotically, contrary to our situation. If one approaches the intersection along the wall, one meets precisely the same situation as the wall in the five dimensional Anti de Sitter space. For instance graviton zero mode is exponentially suppressed away from the intersection along the wall direction to produce a normalizable wave function. Therefore the Anti de Sitter geometry along the wall plays an essential role to achieve the localization of the wave function on the intersection in models with cosmological constant. In sect. 2, we introduce BPS equations and the exact solution for the domain wall junction and discuss representations of the surviving $`(1,0)`$ supersymmetry algebra. In sect. 3, we present mode equations which define the fluctuations on the background of domain wall junction. We work out the Nambu-Goldstone mode explicitly and show that they are not normalizable. Physical origin of the nonnormalizability is clarified and the general validity of this phenomenon is argued. In sect. 4, the relation between the choice of BPS equations and the boundary condition is discussed for a general Wess-Zumino model. In sect. 5, the central charge density and the energy density are examined and an interesting behavior is observed. The fermionic contributions to the central charges and mode equations in a convenient gamma matrix representation are given in appendices. ## 2 BPS equations and the $`(1,0)`$ supersymmetry algebra ### 2.1 Two $`1/4`$ BPS states and two BPS equations It is known that if the translational invariance is broken as is the case for domain walls and/or junctions, the $`𝒩=1`$ superalgebra in general receives contributions from central charges , , . The anti-commutator between two left-handed supercharges has central charges $`Z_k`$, $`k=1,2,3`$ $$\{Q_\alpha ,Q_\beta \}=2i(\sigma ^k\overline{\sigma }^0)_\alpha {}_{}{}^{\gamma }ϵ_{\gamma \beta }^{}Z_k.$$ (2.1) Here and the following we use two-component spinors following the convention of ref. except that the four dimensional indices are denoted by Greek letters $`\mu ,\nu =0,1,2,3`$ instead of roman letters $`m,n`$. The anti-commutator between left- and right-handed supercharges receives a contribution from central charges $`Y_k,k=1,2,3`$ besides the energy-momentum four-vector $`P^\mu ,\mu =0,\mathrm{},3`$ of the system $$\{Q_\alpha ,\overline{Q}_{\dot{\alpha }}\}=2(\sigma _{\alpha \dot{\alpha }}^\mu P_\mu +\sigma _{\alpha \dot{\alpha }}^kY_k).$$ (2.2) One may call $`Z_k`$ and $`Y_k`$ as $`(1,0)`$ and $`(1/2,1/2)`$ central charges in accordance with the transformation properties under the Lorentz group. Central charges, $`Z_k`$ and $`Y_k`$, come from the total divergence, and they are non-vanishing when there are nontrivial differences in asymptotic behavior of fields in different region of spatial infinity as is the case of domain walls and junctions . Therefore these charges are topological in the sense that they are determined completely by the boundary conditions at infinity. For instance, we can take a general Wess-Zumino model with an arbitrary number of chiral superfields $`\mathrm{\Phi }^i`$, an arbitrary superpotential $`𝒲`$ and an arbitrary Kähler potential $`K(\mathrm{\Phi }^i,\mathrm{\Phi }^j)`$ $$=d^2\theta d^2\overline{\theta }K(\mathrm{\Phi }^i,\mathrm{\Phi }^j)+\left[d^2\theta 𝒲(\mathrm{\Phi }^i)+\text{h.c.}\right],$$ (2.3) and compute the anticommutators (2.1), (2.2) to find the central charges. The contributions to these central charges from bosonic components of chiral superfields are given <sup>*</sup><sup>*</sup>*The central charge $`Y_k`$ also receives contributions from fermionic components of chiral superfields which is given in appendix A. by $$Z_k=2d^3x_k𝒲^{}(A^{}),$$ (2.4) $$Y_k=iϵ^{knm}d^3xK_{ij^{}}_n(A^j_mA^i),ϵ^{123}=1,$$ (2.5) where $`A^i`$ is the scalar component of the $`i`$-th chiral superfield $`\mathrm{\Phi }^i`$ and $`K_{ij^{}}=^2K(A^{},A)/A^iA^j`$ is the Kähler metric. BPS domain wall is a $`1/2`$ BPS state and BPS domain wall junction is a $`1/4`$ BPS state . To find the BPS equations satisfied by these BPS states, we consider a hermitian linear combination of operators $`Q`$ and $`\overline{Q}`$ with an arbitrary complex two-vector $`\beta ^\alpha `$ and its complex conjugate $`\overline{\beta }^{\dot{\alpha }}=(\beta ^\alpha )^{}`$ as coefficients $$K=\beta ^\alpha Q_\alpha +\overline{\beta }^{\dot{\alpha }}\overline{Q}_{\dot{\alpha }}.$$ (2.6) We treat $`\beta ^\alpha `$ as c-numbers rather than the Grassmann numbers. Since $`K`$ is hermitian, the expectation value of the square of $`K`$ over any state is non-negative definite, $`S|K^2|S0`$. The field configuration of static junction must be at least two-dimensional. If we assume, for simplicity, that it depends on $`x^1`$, $`x^2`$ then we obtain $`Z_3=Y_1=Y_2=0`$ from Eqs.(2.4) and (2.5), and the inequality implies in this case $`H`$ $``$ $`{\displaystyle \frac{1}{|\beta ^1|^2+|\beta ^2|^2}}\{(|\beta ^1|^2|\beta ^2|^2)Y_3+\mathrm{Re}\left[(\beta ^1)^2Z_2iZ_1\right]`$ (2.7) $`+\mathrm{Re}\left[(\beta ^2)^2Z_2+iZ_1\right]\},`$ for any $`\beta ^\alpha `$ and for any state. The equality holds if and only if the linear combination of supercharges, $`K`$, is preserved by the state $`|S`$ $$K|S=0.$$ (2.8) In this case, the state $`|S`$ saturates the energy bound and is called a BPS state. We find that there are two candidates for the saturation of the energy bound ; $`H=H_\mathrm{I}|iZ_1Z_2|Y_3,\text{when}\overline{\beta }^{\dot{1}}=\beta ^1{\displaystyle \frac{iZ_1+Z_2}{|iZ_1+Z_2|}},\beta ^2=\overline{\beta }^{\dot{2}}=0,`$ (2.9) $`H=H_{\mathrm{II}}|iZ_1Z_2|+Y_3,\text{when}\beta ^1=\overline{\beta }^{\dot{1}}=0,\overline{\beta }^{\dot{2}}=\beta ^2{\displaystyle \frac{iZ_1+Z_2}{|iZ_1+Z_2|}}.`$ (2.10) In the case of $`H_\mathrm{I}H_{\mathrm{II}}`$, the BPS bound becomes $`H\mathrm{max}\{H_\mathrm{I},H_{\mathrm{II}}\}.`$ If $`H_\mathrm{I}>H_{\mathrm{II}}`$, then supersymmetry can only be preserved at $`H=H_\mathrm{I}`$ and the only one combination of supercharges is conserved $$\left(Q_1+\frac{iZ_1+Z_2}{|iZ_1+Z_2|}\overline{Q}_{\dot{1}}\right)|H=H_\mathrm{I}=0.$$ (2.11) If $`H_{\mathrm{II}}>H_\mathrm{I}`$, then supersymmetry can only be preserved at $`H=H_{\mathrm{II}}`$ and the only one combination of supercharges is conserved $$\left(Q_2+\frac{iZ_1+Z_2}{|iZ_1+Z_2|}\overline{Q}_{\dot{2}}\right)|H=H_{\mathrm{II}}=0.$$ (2.12) In the case of $`H_\mathrm{I}=H_{\mathrm{II}}`$, two candidates of BPS bounds coincide and BPS state conserves both of two supercharges, (2.11) and (2.12); this is a $`1/2`$ BPS state. For the general Wess-Zumino model in Eq.(2.3), the condition of supercharge conservation (2.11) for $`H=H_\mathrm{I}`$ applied to chiral superfield $`\mathrm{\Phi }^i=(A^i,\psi ^i,F^i)`$ gives after eliminating the auxiliary field $`F^i`$ $$2\frac{A^i}{\overline{z}}=\mathrm{\Omega }_+F^i=\mathrm{\Omega }_+K^{1ij^{}}\frac{𝒲^{}}{A^j},\mathrm{\Omega }_+i\frac{iZ_1^{}+Z_2^{}}{|iZ_1^{}+Z_2^{}|},$$ (2.13) where complex coordinates $`z=x^1+ix^2,\overline{z}=x^1ix^2`$, and the inverse of the Kähler metric $`K^{1ij^{}}`$ are introduced. We can also consider gauge interactions. For simplicity we take only the $`U(1)`$ gauge interaction. Then the derivative $`A^i/\overline{z}`$ in the above Eq.(2.13) should be replaced by the gauge covariant derivative $`𝒟_{\overline{z}}A^i`$ $$2𝒟_{\overline{z}}A^i=\mathrm{\Omega }_+K^{1ij^{}}\frac{𝒲^{}}{A^j},𝒟_{\overline{z}}=\frac{1}{2}(𝒟_1+i𝒟_2),𝒟_\mu A^i=\left(\frac{}{x^\mu }+i\frac{e_i}{2}v_\mu \right)A^i.$$ (2.14) Moreover the same BPS condition (2.11) applied to vector superfield in the Wess-Zumino gauge $`V=(v_\mu ,\lambda ,D)`$ gives after eliminating the auxiliary field $`D`$ $$v_{12}=D=\frac{1}{2}\underset{j}{}A^je_jA^j,v_{03}=0,v_{01}=v_{31},v_{23}=v_{02},$$ (2.15) where $`v_{\mu \nu }_\mu v_\nu _\nu v_\mu `$ and $`e_j`$ is the charge of the field $`A^j`$. Here we assume for simplicity the minimal kinetic term both for the chiral superfield $`K_{ij^{}}=\delta _{ij^{}}`$ and for the vector superfield. Similarly the condition of supercharge conservation (2.12) for $`H=H_{\mathrm{II}}`$ applied to chiral superfield in the Wess-Zumino model gives after eliminating the auxiliary field $$2\frac{A^i}{z}=\mathrm{\Omega }_{}F^i=\mathrm{\Omega }_{}K^{1ij^{}}\frac{𝒲^{}}{A^j},\mathrm{\Omega }_{}i\frac{iZ_1^{}Z_2^{}}{|iZ_1^{}Z_2^{}|}.$$ (2.16) If $`U(1)`$ gauge interaction is present, the derivative $`A^i/z`$ should be replaced by the covariant derivative $`𝒟_zA^i=\frac{1}{2}(𝒟_1i𝒟_2)A^i`$. In this case the BPS condition applied to $`U(1)`$ vector superfield in the Wess-Zumino gauge becomes in the case of minimal kinetic terms $$v_{12}=D=\frac{1}{2}\underset{j}{}A^je_jA^j,v_{03}=0,v_{01}=v_{31},v_{23}=v_{02}.$$ (2.17) In sect.4, we shall present a simple way to find the correspondence between the choice of boundary conditions and the choice of BPS equations (2.13) and (2.15) or (2.16) and (2.17). ### 2.2 The exact solution of BPS domain wall junction In a previous article , we have found an exact solution of BPS domain wall junction in a model motivated by the $`𝒩=2`$ supersymmetric $`SU(2)`$ gauge theory with one flavor broken to $`𝒩=1`$ by the mass of the adjoint chiral superfield. This model has the following chiral superfields with the charge assignment for the $`U(1)\times U(1)^{}`$ gauge group $$\begin{array}{cccccccc}& & \stackrel{~}{}& 𝒟& \stackrel{~}{𝒟}& 𝒬& \stackrel{~}{𝒬}& T\\ U(1)& 0& 0& 1& 1& 1& 1& 0\\ U(1)^{}& 1& 1& 1& 1& 0& 0& 0,\end{array}$$ (2.18) interacting with a superpotential $$𝒲=(T\mathrm{\Lambda })\stackrel{~}{}+(T+\mathrm{\Lambda })𝒟\stackrel{~}{𝒟}+(Tm)𝒬\stackrel{~}{𝒬}h^2T,$$ (2.19) where parameters $`\mathrm{\Lambda }`$ and $`h`$ can be made real positive and a parameter $`m`$ is complex . In this model there are three discrete vacua, $`\mathrm{Vac}.1`$ $`:`$ $`T=\mathrm{\Lambda },=\stackrel{~}{}=h,𝒬=\stackrel{~}{𝒬}=𝒟=\stackrel{~}{𝒟}=0,𝒲_1=h^2\mathrm{\Lambda },`$ $`\mathrm{Vac}.2`$ $`:`$ $`T=m,𝒬=\stackrel{~}{𝒬}=h,=\stackrel{~}{}=𝒟=\stackrel{~}{𝒟}=0,𝒲_2=h^2m,`$ $`\mathrm{Vac}.3`$ $`:`$ $`T=\mathrm{\Lambda },𝒟=\stackrel{~}{𝒟}=h,𝒬=\stackrel{~}{𝒬}==\stackrel{~}{}=0,𝒲_3=h^2\mathrm{\Lambda },`$ (2.20) and when $`m=i\sqrt{3}\mathrm{\Lambda }`$, this model becomes $`Z_3`$ symmetric. Thus three half walls are expected to connect at the junction with relative angles of $`2\pi /3`$. For definiteness, we specify the boundary condition where the wall $`1`$ extends along the negative $`x^2`$ axis separating the vacuum $`1`$ ($`x^1>0`$) and $`3`$ ($`x^1<0`$) as shown in Fig. 1. If we have only the wall $`1`$, we obtain the central charge $`Z_k`$ (vanishing $`Y_k`$) and find the two conserved supercharges from Eqs.(2.11) and (2.12) as $$Q^{(1)}=\frac{1}{\sqrt{2}}(\mathrm{e}^{i\frac{\pi }{4}}Q_2+\mathrm{e}^{i\frac{\pi }{4}}\overline{Q}_{\dot{2}}),Q^{(2)}=\frac{1}{\sqrt{2}}(\mathrm{e}^{i\frac{\pi }{4}}Q_1+\mathrm{e}^{i\frac{\pi }{4}}\overline{Q}_{\dot{1}}).$$ (2.21) The other two walls have also two conserved supercharges at wall 2 $`Q^{(3)}={\displaystyle \frac{1}{\sqrt{2}}}(\mathrm{e}^{i\frac{\pi }{12}}Q_1+\mathrm{e}^{i\frac{\pi }{12}}\overline{Q}_{\dot{1}}),\text{besides}Q^{(1)}={\displaystyle \frac{1}{\sqrt{2}}}(\mathrm{e}^{i\frac{\pi }{4}}Q_2+\mathrm{e}^{i\frac{\pi }{4}}\overline{Q}_{\dot{2}}),`$ at wall 3 $`Q^{(4)}={\displaystyle \frac{1}{\sqrt{2}}}(\mathrm{e}^{i\frac{5\pi }{12}}Q_1+\mathrm{e}^{i\frac{5\pi }{12}}\overline{Q}_{\dot{1}}),\text{besides}Q^{(1)}={\displaystyle \frac{1}{\sqrt{2}}}(\mathrm{e}^{i\frac{\pi }{4}}Q_2+\mathrm{e}^{i\frac{\pi }{4}}\overline{Q}_{\dot{2}}).`$ (2.22) When these three half walls coexist, we can have only one common conserved supercharge $`Q^{(1)}=(\mathrm{e}^{i\frac{\pi }{4}}Q_2+\mathrm{e}^{i\frac{\pi }{4}}\overline{Q}_{\dot{2}})/\sqrt{2}`$. In fact we find that the domain wall junction configuration conserves precisely this single combination of supercharges, even though it has also another central charge $`Y_k`$ contributing. Correspondingly we obtain the BPS equations (2.16) and (2.17) for $`H=H_{\mathrm{II}}`$ with $`\mathrm{\Omega }_{}=1`$. The BPS equations (2.17) for the vector superfield can be trivially satisfied by $`v_\mu =0`$ and $`D=0`$. The BPS equations (2.16) for chiral superfields become in this case $$2\frac{A^i}{z}=\frac{𝒲^{}}{A^i},$$ (2.23) assuming the minimal kinetic term. The solution for these BPS equations is given by , $`(z,\overline{z})`$ $`=`$ $`\stackrel{~}{}(z,\overline{z})={\displaystyle \frac{\sqrt{2}\mathrm{\Lambda }s}{s+t+u}},`$ $`𝒟(z,\overline{z})`$ $`=`$ $`\stackrel{~}{𝒟}(z,\overline{z})={\displaystyle \frac{\sqrt{2}\mathrm{\Lambda }t}{s+t+u}},`$ $`𝒬(z,\overline{z})`$ $`=`$ $`\stackrel{~}{𝒬}(z,\overline{z})={\displaystyle \frac{\sqrt{2}\mathrm{\Lambda }u}{s+t+u}},`$ $`T(z,\overline{z})`$ $`=`$ $`{\displaystyle \frac{2\mathrm{\Lambda }}{\sqrt{3}}}{\displaystyle \frac{e^{i\frac{1}{6}\pi }s+e^{i\frac{5}{6}\pi }t+e^{i\frac{1}{2}\pi }u}{s+t+u}}+{\displaystyle \frac{i}{\sqrt{3}}}\mathrm{\Lambda },`$ (2.24) $`s=\mathrm{exp}\left({\displaystyle \frac{2\mathrm{\Lambda }}{\sqrt{3}}}\mathrm{Re}\left(e^{i\frac{1}{6}\pi }z\right)\right),t=\mathrm{exp}\left({\displaystyle \frac{2\mathrm{\Lambda }}{\sqrt{3}}}\mathrm{Re}\left(e^{i\frac{5}{6}\pi }z\right)\right),u=\mathrm{exp}\left({\displaystyle \frac{2\mathrm{\Lambda }}{\sqrt{3}}}\mathrm{Re}\left(e^{i\frac{1}{2}\pi }z\right)\right).`$ (2.25) This model is motivated by the softly broken $`𝒩=2`$ $`SU(2)`$ gauge theory with one flavor. However, we can simplify the model without spoiling the solvability to obtain a Wess-Zumino model consisting of purely chiral superfields by the following procedure. The vector superfields actually serve to constrain chiral superfields to have the identical magnitude pairwise through $`D=0`$ to satisfy the BPS equation (2.17) for vector superfields: $`|\stackrel{~}{}|=||,|\stackrel{~}{𝒟}|=|𝒟|,|\stackrel{~}{𝒬}|=|𝒬|`$. Therefore we can eliminate the vector superfields and reduce the number of chiral superfields by identifying pairwise $`\stackrel{~}{}=,\stackrel{~}{𝒟}=𝒟,\stackrel{~}{𝒬}=𝒬`$. Correspondingly we should take the superpotential as $$𝒲=\frac{1}{2}(T\mathrm{\Lambda })^2+\frac{1}{2}(T+\mathrm{\Lambda })𝒟^2+\frac{1}{2}(Ti\sqrt{3}\mathrm{\Lambda })𝒬^2\frac{h^2}{2}T.$$ (2.26) This Wess-Zumino model has the same solution as ours by changing $`h^2h^2/2,\mathrm{\Lambda }\sqrt{3}\mathrm{\Lambda }/2`$. A similar observation has also been made in ref.. ### 2.3 Unitary representations of $`(1,0)`$ supersymmetry algebra Let us examine states on the background of a domain wall junction from the point of view of surviving symmetry. In the case of the BPS states satisfying the BPS equation (2.16) corresponding to $`H=H_{\mathrm{II}}`$, we have only one surviving supersymmetry charge $`Q^{(1)}`$, two translation generators $`H,P^3`$, and one Lorentz generator $`J^{03}`$, out of the $`𝒩=1`$ four dimensional super Poincaré generators. Since we are interested in excitation modes on the background of the domain wall junction, we define the hamiltonian $`H^{}=HH`$ measured from the energy $`H`$ of the background configuration. By projecting from the supersymmetry algebra (2.1), (2.2) with central charges in four dimensions, we immediately find $$\left(Q^{(1)}\right)^2=H^{}P^3.$$ (2.27) We also obtain the Poincaré algebra in $`1+1`$ dimensions $$[J^{03},Q^{(1)}]=\frac{i}{2}Q^{(1)},[J^{03},H^{}P^3]=i(H^{}P^3),[J^{03},H^{}+P^3]=i(H^{}+P^3).$$ (2.28) Other commutation relations are trivial $$[H^{}P^3,H^{}+P^3]=[H^{}P^3,Q^{(1)}]=[H^{}+P^3,Q^{(1)}]=0.$$ (2.29) This is precisely the $`(1,0)`$ supersymmetry algebra on the domain wall junction as anticipated . To obtain unitary representations, we can diagonalize $`H^{}`$ and $`P^3`$ $$H^{}|E,p^3=E|E,p^3,P^3|E,p^3=p^3|E,p^3,E|p^3|,$$ (2.30) and combine them by means of $`Q^{(1)}`$. If $`Ep^3>0`$, we can construct bosonic state from fermionic state and vice versa by operating $`Q^{(1)}`$ on the state. $$|B=\frac{1}{\sqrt{Ep^3}}Q^{(1)}|F,|F=\frac{1}{\sqrt{Ep^3}}Q^{(1)}|B.$$ (2.31) Therefore we obtain a doublet representation $`(|B,|F)`$. If $`Ep^3=0`$, operating by $`Q^{(1)}`$ on the state gives an unphysical zero norm state $$\left|Q^{(1)}|E,p^3\right|^2=E,p^3|\left(Q^{(1)}\right)^2|E,p^3=E,p^3|H^{}P^3|E,p^3=Ep^3=0.$$ (2.32) Then the massless right-moving state $`|E,p^3=E`$ is a singlet representation. This singlet state can either be boson or fermion. Thus we find that there are only two types of representations of the $`(1,0)`$ supersymmetry algebra, doublet and singlet. We also find that massive modes should appear in pairs of boson and fermion, whereas the massless right-moving mode can appear singly without accompanying a state with opposite statistics. This provides an interesting possibility of a chiral structure for fermions. If another BPS equation (2.13) corresponding to $`H=H_\mathrm{I}`$ is satisfied instead of Eq. (2.16), we have $`(0,1)`$ supersymmetry and the left-moving massless states can appear as singlets. ## 3 Nambu-Goldstone and other modes on the junction ### 3.1 Mode equation on the junction Since the vector superfields have no nontrivial field configurations, Nambu-Goldstone modes have no component of vector superfield. Moreover we can replace our model, if we wish, by another model with purely chiral superfields without spoiling the essential features including the solvability. Consequently we shall neglect vector superfields and consider the general Wess-Zumino model in Eq.(2.3) in the following. For simplicity we assume the minimal kinetic term here $`K_{ij^{}}=\delta _{ij^{}}`$ . Let us consider quantum fluctuations $`A^i,\psi ^i`$ around a classical solution $`A_{\mathrm{cl}}^i`$ which satisfies the BPS equations (2.13) and (2.15) for $`H=H_\mathrm{I}`$ or (2.16) and (2.17) for $`H=H_{\mathrm{II}}`$. $$A^i=A_{\mathrm{cl}}^i+A^i.$$ (3.33) We retain the part of the Lagrangian quadratic in fluctuations and eliminate the auxiliary fields $`F^i`$ to obtain the linearized equation for the scalar fluctuations $$_\mu ^\mu A^i+\frac{^2𝒲}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^k}\frac{^2𝒲^{}}{A_{\mathrm{cl}}^kA_{\mathrm{cl}}^j}A^j+\frac{^3𝒲}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^kA_{\mathrm{cl}}^j}\frac{𝒲^{}}{A_{\mathrm{cl}}^k}A^j=0.$$ (3.34) In order to separate variables in $`x^0,x^3`$ and $`x^1,x^2`$ we have to define mode equations on the background which has a nontrivial dependence in two dimensions, $`x^1,x^2`$. The bosonic modes $`A_n^i(x^1,x^2)`$ can easily be defined in terms of a differential operator $`𝒪_B`$ in $`x^1,x^2`$ space $$𝒪_B{}_{}{}^{i}{}_{j}{}^{}\left[\begin{array}{cc}(_1^2+_2^2)\delta ^i{}_{j}{}^{}+\frac{^2𝒲}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^k}\frac{^2𝒲^{}}{A_{\mathrm{cl}}^kA_{\mathrm{cl}}^j}& \frac{^3𝒲}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^kA_{\mathrm{cl}}^j}\frac{𝒲^{}}{A_{\mathrm{cl}}^k}\\ \frac{^3𝒲^{}}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^kA_{\mathrm{cl}}^j}\frac{𝒲}{A_{\mathrm{cl}}^k}& (_1^2+_2^2)\delta ^i{}_{j}{}^{}+\frac{^2𝒲^{}}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^k}\frac{^2𝒲}{A_{\mathrm{cl}}^kA_{\mathrm{cl}}^j}\end{array}\right]$$ (3.35) $$𝒪_B{}_{}{}^{i}{}_{j}{}^{}\left[\begin{array}{c}A_n^j\\ A_n^j\end{array}\right]=M_n^2\left[\begin{array}{c}A_n^i\\ A_n^i\end{array}\right],$$ (3.36) where the eigenvalue $`M_n^2`$ has to be real from Majorana condition. The quantum fluctuation for scalar can be expanded in terms of these mode functions to obtain a real scalar field equation with the mass $`M_n`$ for the coefficient bosonic field $`a_n(x^0,x^3)`$ $$A^i(x^0,x^1,x^2,x^3)=\underset{n}{}a_n(x^0,x^3)A_n^i(x^1,x^2)$$ (3.37) $$\left(_0^2_3^2+M_n^2\right)a_n(x^0,x^3)=0.$$ (3.38) Similarly the linearized equation for fermions is given by $$i\overline{\sigma }^\mu _\mu \psi ^i\frac{^2𝒲^{}}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^j}\overline{\psi }^j=0$$ (3.39) $$i\sigma ^\mu _\mu \overline{\psi }^i\frac{^2𝒲}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^j}\psi ^j=0.$$ (3.40) To separate variables for fermion equations, it is more convenient to use a gamma matrix representation where direct product structure of $`2\times 2`$ matrices for $`(x^0,x^3)`$ and $`(x^1,x^2)`$ space is manifest. We shall describe one such representation in appendix B. Transforming from such a representation to the Weyl representation which we are using, we can define the fermionic modes $`\psi _{n\alpha }^i,\overline{\psi }_n^{i\dot{\beta }}`$ combining components of left-handed and right-handed spinors by means of the following operators $$𝒪_1{}_{}{}^{i}{}_{j}{}^{}\left[\begin{array}{cc}\frac{^2𝒲^{}}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^j}& i\left(_1+i_2\right)\delta _j^i\\ i\left(_1+i_2\right)\delta _j^i& \frac{^2𝒲}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^j}\end{array}\right]$$ (3.41) $$𝒪_2{}_{}{}^{i}{}_{j}{}^{}\left[\begin{array}{cc}\frac{^2𝒲}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^j}& i\left(_1i_2\right)\delta _j^i\\ i\left(_1i_2\right)\delta _j^i& \frac{^2𝒲^{}}{A_{\mathrm{cl}}^iA_{\mathrm{cl}}^j}\end{array}\right]$$ (3.42) $$𝒪_1{}_{}{}^{i}{}_{j}{}^{}\left[\begin{array}{c}\overline{\psi }_n^{j\dot{1}}\\ \psi _{n2}^j\end{array}\right]=im_n^{(1)}\left[\begin{array}{c}\psi _{n1}^i\\ \overline{\psi }_n^{i\dot{2}}\end{array}\right]$$ (3.43) $$𝒪_2{}_{}{}^{i}{}_{j}{}^{}\left[\begin{array}{c}\psi _{n1}^j\\ \overline{\psi }_n^{j\dot{2}}\end{array}\right]=im_n^{(2)}\left[\begin{array}{c}\overline{\psi }_n^{i\dot{1}}\\ \psi _{n2}^i\end{array}\right],$$ (3.44) where the mass eigenvalues $`m_n^{(1)},m_n^{(2)}`$ are real. Please note a peculiar combination of left- and right-handed spinor components to define eigenfunctions. We can expand $`\psi ^i`$ in terms of these mode functions $$\psi _\alpha ^i(x^0,x^1,x^2,x^3)=\underset{n}{}\left(\begin{array}{c}b_n(x^0,x^3)\psi _{n1}^i(x^1,x^2)\\ c_n(x^0,x^3)\psi _{n2}^i(x^1,x^2)\end{array}\right)$$ (3.45) Since $`\psi (x^0,x^1,x^2,x^3)`$ is a Majorana spinor, the coefficient fermionic fields $`b_n,c_n`$ are real. The linearized equations (3.39) (3.40) for the fermion gives a Dirac equation in $`1+1`$ dimensions for the coefficient fermionic fields $`(c_n,ib_n)`$ with two mass parameters $`m_n^{(1)},m_n^{(2)}`$ $$\left[i\left(\rho _1_0+i\rho _2_3\right)m_n^{(1)}\frac{1+\rho _3}{2}m_n^{(2)}\frac{1\rho _3}{2}\right]\left[\begin{array}{c}c_n(x^0,x^3)\\ ib_n(x^0,x^3)\end{array}\right]=0,$$ (3.46) where we use Pauli matrices $`\rho _a,a=1,2,3`$ to construct the $`2\times 2`$ gamma matrices $`\rho _1,i\rho _2`$ in $`1+1`$ dimensions. Since we have a Majorana spinor in $`1+1`$ dimensions which does not allow chiral rotations, we have two distinct real mass parameters $`m_n^{(1)},m_n^{(2)}`$. To relate the mass eigenvalues of fermions and bosons, let us multiply two differential operators for fermions $`𝒪_2`$ to $`𝒪_1`$. In this ordering, we can use the BPS equation (2.16) corresponding to $`H=H_{\mathrm{II}}`$ to find the differential operator for bosons $`𝒪_B`$ $$𝒪_2^i{}_{k}{}^{}𝒪_{1}^{k}{}_{j}{}^{}=\left[\begin{array}{cc}\mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_{}^{\frac{1}{2}}& 0\\ 0& \mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_{}^{\frac{1}{2}}\end{array}\right]𝒪_B^i{}_{j}{}^{}\left[\begin{array}{cc}\mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_{}^{\frac{1}{2}}& 0\\ 0& \mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_{}^{\frac{1}{2}}\end{array}\right].$$ (3.47) Therefore the BPS equation (2.16) corresponding to $`H=H_{\mathrm{II}}`$ guarantees that the existence of a solution $`\overline{\psi }_n^{i\dot{1}},\psi _{n2}^i`$ of fermionic mode equations implies the existence of a solution of bosonic mode equations with the mass squared $`M_n^2=m_n^{(1)}m_n^{(2)}`$ $$A_n^i=\mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_{}^{\frac{1}{2}}\overline{\psi }_n^{i\dot{1}},A_n^i=\mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_{}^{\frac{1}{2}}\psi _{n2}^i.$$ (3.48) If another BPS equation (2.13) corresponding to $`H=H_\mathrm{I}`$ is valid, operator multiplication with different ordering gives the same bosonic operator whose rows and columns are interchanged $$𝒪_1^i{}_{k}{}^{}𝒪_{2}^{k}{}_{j}{}^{}=\left[\begin{array}{cc}0& \mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_+^{\frac{1}{2}}\\ \mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_+^{\frac{1}{2}}& 0\end{array}\right]𝒪_B^i{}_{j}{}^{}\left[\begin{array}{cc}0& \mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_+^{\frac{1}{2}}\\ \mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_+^{\frac{1}{2}}& 0\end{array}\right].$$ (3.49) Therefore the BPS equation (2.13) corresponding to $`H=H_\mathrm{I}`$ guarantees that the existence of a solution $`\overline{\psi }_n^{i\dot{2}},\psi _{n1}^i`$ of fermionic mode equations implies the existence of a solution of bosonic mode equations with the mass squared $`M_n^2=m_n^{(1)}m_n^{(2)}`$ $$A_n^i=\mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_+^{\frac{1}{2}}\overline{\psi }_n^{i\dot{2}},A_n^i=\mathrm{e}^{i\frac{\pi }{4}}\mathrm{\Omega }_+^{\frac{1}{2}}\psi _{n1}^i.$$ (3.50) Therefore we find that all massive states come in pairs of boson and fermion with the same mass squared $`M_n^2=m_n^{(1)}m_n^{(2)}`$ in accordance with the result of the unitary representation of the $`(1,0)`$ supersymmetry algebra. ### 3.2 Nambu-Goldstone modes Since we are usually most interested in a low energy effective field theory, we wish to study massless modes here. If global continuous symmetries are broken spontaneously, there occur associated massless modes which are called the Nambu-Goldstone modes. To find the wave functions of the Nambu-Goldstone modes, we perform the associated global transformations and evaluate the transformed configuration by substituting the classical field. For supersymmetry we obtain nontrivial wave function by substituting the classical field $`A_{\mathrm{cl}}^i(x^1,x^2)`$ and $`F_{\mathrm{cl}}^i(x^1,x^2)`$ to the transformation of fermions by a Grassmann parameter $`\xi `$, since classical field configuration of fermion vanishes $`\psi _{\mathrm{cl}}^i=0`$ $$\delta _\xi \psi ^i=i\sqrt{2}\sigma ^\mu \overline{\xi }_\mu A_{\mathrm{cl}}^i+\sqrt{2}\xi F_{\mathrm{cl}}^i.$$ (3.51) If the BPS equation (2.16) for the junction background is valid, we obtain $$\delta _\xi \psi ^i=\sqrt{2}\left[(i\sigma ^1\overline{\xi }\mathrm{\Omega }_{}^{}\xi )_1A_{\mathrm{cl}}^i+(i\sigma ^2\overline{\xi }+i\mathrm{\Omega }_{}^{}\xi )_2A_{\mathrm{cl}}^i\right].$$ (3.52) We see that there is one conserved direction in the Grassmann parameter: $$i\sigma ^1\overline{\xi }=\mathrm{\Omega }_{}^{}\xi \text{and}\sigma ^2\overline{\xi }=\mathrm{\Omega }_{}^{}\xi .$$ (3.53) The other three real Grassmann parameters $`\xi `$ correspond to broken supercharges. For our exact solution, for instance, we find it convenient to choose the three broken supercharges as the following real supercharges $$Q_\mathrm{I}=\frac{1}{\sqrt{2}}(e^{i\pi /4}Q_2+e^{i\pi /4}\overline{Q}_{\dot{2}}),Q_{\mathrm{II}}=\frac{1}{\sqrt{2}}(e^{i\pi /4}Q_1+e^{i\pi /4}\overline{Q}_{\dot{1}}),Q_{\mathrm{III}}=\frac{1}{\sqrt{2}}(e^{i\pi /4}Q_1+e^{i\pi /4}\overline{Q}_{\dot{1}}).$$ (3.54) Then the corresponding massless mode functions are given by $`\psi _0^{(\mathrm{I})i}(x^1,x^2)`$ $`=`$ $`\left(\begin{array}{c}4_zA_{\mathrm{cl}}^i(x^1,x^2)e^{i\pi /4}\\ 0\end{array}\right),`$ (3.57) $`\psi _0^{(\mathrm{II})i}(x^1,x^2)`$ $`=`$ $`\left(\begin{array}{c}0\\ 2_1A_{\mathrm{cl}}^i(x^1,x^2)e^{i\pi /4}\end{array}\right),`$ (3.60) $`\psi _0^{(\mathrm{III})i}(x^1,x^2)`$ $`=`$ $`\left(\begin{array}{c}0\\ 2_2A_{\mathrm{cl}}^i(x^1,x^2)e^{i\pi /4}\end{array}\right).`$ (3.63) Since the transformation parameter should correspond to the Nambu-Goldstone field with zero momentum and energy, the three transformation parameters $`\xi `$ should be promoted to three real fermionic fields in $`x^0,x^3`$ space, $`b_0^{(\mathrm{I})}(x^0,x^3),c_0^{(\mathrm{II})}(x^0,x^3)`$, and $`c_0^{(\mathrm{III})}(x^0,x^3)`$, to obtain the Nambu-Goldstone component of the mode expansion $`\psi ^i(x^0,x^1,x^2,x^3)`$ $`=`$ $`b_0^{(\mathrm{I})}(x^0,x^3)\psi _0^{(\mathrm{I})i}(x^1,x^2)+c_0^{(\mathrm{II})}(x^0,x^3)\psi _0^{(\mathrm{II})i}(x^1,x^2)`$ (3.64) $`+c_0^{(\mathrm{III})}(x^0,x^3)\psi _0^{(\mathrm{III})i}(x^1,x^2)+{\displaystyle \underset{n>0}{}}\left(\begin{array}{c}b_n(x^0,x^3)\psi _{n1}^i(x^1,x^2)\\ c_n(x^0,x^3)\psi _{n2}^i(x^1,x^2)\end{array}\right).`$ We have explicitly displayed three massless Nambu-Goldstone fermion components distinguishing from the massive ones ($`n>0`$). The Dirac equation for the coefficient fermionic fields (3.46) shows that $`b_0^{(\mathrm{I})}(x^0x^3)`$ is a right-moving massless mode, and $`c_0^{(\mathrm{II})}(x^0+x^3)`$, and $`c_0^{(\mathrm{III})}(x^0+x^3)`$ are left-moving modes. We plot the absolute values of $`|\psi _0^{(a)i=T}|`$ of the $`i=T`$ component of the wave function of the Nambu-Goldstone fermions $`a=\mathrm{I},\mathrm{II},\mathrm{III}`$ in Fig. 2. We can see that Nambu-Goldstone fermions have wave functions which extend to infinity along three walls. They become identical to fermion zero modes on at least two of the walls asymptotically and hence they are not localized around the center of the junction. We can construct a linear combination of the Nambu-Goldstone fermions to have no support along one out of the three walls. However, no linear combination of these Nambu-Goldstone fermions can be formed which does not have support extended along any of the wall. Therefore these wave functions are not localized and are not normalizable. This fact means that the low energy dynamics of BPS junction cannot be described by a $`1+1`$ dimensional effective field theory with a discrete particle spectrum. Similarly the Nambu-Goldstone bosons corresponding to the broken translation $`P^a,a=1,2`$ are given by $$A_0^{(a)}(x^1,x^2)=_aA_{\mathrm{cl}}^i(x^1,x^2),a=1,2.$$ (3.65) These two bosonic massless modes consist of two left-moving modes and two right-moving modes. On the other hand, we have seen already that there are two left-moving massless Nambu-Goldstone fermions and one right-moving massless Nambu-Goldstone fermion. These two left-moving Nambu-Goldstone bosons and fermions form two doublets of the $`(1,0)`$ supersymmetry algebra. The right-moving modes are asymmetric in bosons and fermions: two Nambu-Goldstone bosons and a single Nambu-Goldstone fermion. These three states are all singlets of the $`(1,0)`$ supersymmetry algebra in accordance with our analysis in sect.2.3. Therefore we obtained a chiral structure of Nambu-Goldstone fermions on the junction background configuration. ### 3.3 Non-normalizability of the Nambu-Goldstone fermions We would like to argue that our observation is a generic feature of the Nambu-Goldstone fermions on the domain wall junction in a flat space in the bulk: Nambu-Goldstone fermions are not localized at the junction and hence are not normalizable, if they are associated with the supersymmetry breaking due to the coexistence of nonparallel domain walls. The following observation is behind this assertion. A single domain wall breaks only a half of supercharges. Nonparallel wall also breaks half of supercharges, some of which may be linear combinations of the supercharges already broken by the first wall. If the junction configuration is a $`1/4`$ BPS state, linearly independent ones among these two sets of broken supercharges of nonparallel walls become $`\frac{3}{4}`$ of the original supercharges. To see in more detail, let us first note that the junction configuration reduces asymptotically to a wall if one goes along the wall, say the wall $`1`$. On this first wall, a half of the original supersymmetry ($`Q^{(1)},\mathrm{},Q^{(N)}`$) is broken. Denoting the number of original supercharges to be $`N`$, we call these broken supercharges as $`Q^{(1)},\mathrm{},Q^{(N/2)}`$. Consequently we have Nambu-Goldstone fermions localized around the core of the wall and is constant along the wall. In the junction configuration, we have other walls which are not parallel to the first wall. Asymptotically far away along one of such walls, say wall $`2`$, another half of the supersymmetry $`Q^{(1)},\mathrm{},Q^{(N/2)}`$ is broken. If the junction is a $`1/4`$ BPS state, a half of these, say $`Q^{(1)},\mathrm{},Q^{(N/4)}`$, is a linear combination of $`Q^{(1)},\mathrm{},Q^{(N/2)}`$ broken already on the wall $`1`$. The other half, $`Q^{(\frac{N}{4}+1)},\mathrm{},Q^{(\frac{N}{2})}`$ are unbroken on the wall $`1`$. Altogether a quarter of the original supercharges remain unbroken. Consequently the Nambu-Goldstone fermions corresponding to $`Q^{(1)},\mathrm{},Q^{(N/4)}`$ have a wave function which extends to infinity and approaches a constant profile along both the walls $`1`$ and $`2`$. Those modes corresponding to $`Q^{(\frac{N}{4}+1)},\mathrm{},Q^{(\frac{N}{2})}`$ have support only along the wall $`2`$, and those corresponding to the linear combinations of $`Q^{(1)},\mathrm{},Q^{(N/2)}`$ orthogonal to $`Q^{(1)},\mathrm{},Q^{(N/4)}`$ have support only along the wall $`1`$. Thus we find that any linear combinations of the Nambu-Goldstone fermions have to be infinitely extended along at least one of the walls which form the junction configuration. Therefore the Nambu-Goldstone fermions associated with the coexistence of nonparallel domain walls are not localized at the junction and are not normalizable. In our exact solution, domain wall junction configuration reduces asymptotically to the wall $`1`$ at $`x^2\mathrm{}`$ with fixed $`x^1`$. On the wall, only two supercharges in Eq.(3.54) are broken $$Q_\mathrm{I}=\frac{1}{\sqrt{2}}(e^{i\pi /4}Q_2+e^{i\pi /4}\overline{Q}_{\dot{2}}),Q_{\mathrm{II}}=\frac{1}{\sqrt{2}}(e^{i\pi /4}Q_1+e^{i\pi /4}\overline{Q}_{\dot{1}}),$$ (3.68) and there are two corresponding Nambu-Goldstone fermions which become domain wall zero modes asymptotically $`\psi _0^{(\mathrm{I})i}(x^1,x^2)`$ $`=`$ $`\left(\begin{array}{c}4_zA_{\mathrm{cl}}^i(x^1,x^2)e^{i\pi /4}\\ 0\end{array}\right)\left(\begin{array}{c}2_1A_{\mathrm{cl}}^{i\mathrm{wall}}(x^1)e^{i\pi /4}\\ 0\end{array}\right),`$ (3.73) $`\psi _0^{(\mathrm{II})i}(x^1,x^2)`$ $`=`$ $`\left(\begin{array}{c}0\\ 2_1A_{\mathrm{cl}}^i(x^1,x^2)e^{i\pi /4}\end{array}\right)\left(\begin{array}{c}0\\ 2_1A_{\mathrm{cl}}^{i\mathrm{wall}}(x^1)e^{i\pi /4}\end{array}\right).`$ (3.78) These wave functions are localized on the core of the wall 1 in the $`x^1`$ direction and are constant along the wall. Along the other walls we find two broken supercharges one of which is identical to one of the broken supercharges, $`Q_\mathrm{I}`$. The other broken supercharge is $`Q_{\mathrm{II}}^{}`$ on the wall 2 and $`Q_{\mathrm{II}}^{\prime \prime }`$ on the wall $`3`$. There are only two independent supercharges among $`Q_{\mathrm{II}}`$, $`Q_{\mathrm{II}}^{}`$, and $`Q_{\mathrm{II}}^{\prime \prime }`$. Together with $`Q_\mathrm{I}`$ we obtain three independent broken supercharges. We can construct a linear combination of the Nambu-Goldstone fermions to have no support along one out of the three walls. However, any linear combination has nonvanishing wave function which becomes fermion zero mode on at least one of the wall asymptotically. Therefore the associated Nambu-Goldstone fermions have support which is infinitely extended at least along two of the walls. If a single wall is present, we can explicitly construct a plane wave solution propagating along the wall, which may be called a spin wave and is among massive modes on the wall background. Even if there are several walls forming a junction configuration, we can consider excitation modes which reduce to the spin wave modes along each wall. They should be a massive mode on the domain wall junction background. The Nambu-Goldstone mode on the domain wall junction is the zero wave number limit of such a spin wave mode. This physical consideration suggests that the massless Nambu-Goldstone fermion is precisely the vanishing wave number (along the wall) limit of the massive spin wave mode. Let us note that our argument does not apply to models with the bulk cosmological constant. In such models, massless graviton is localized on the background of intersection of walls . In that case, massless mode is a distinct mode different from the massless limit of the massive continuum, although the massless mode is buried at the tip of the continuum of massive modes. The normalizability of the massless graviton is guaranteed by the Anti de Sitter geometry away from the junction or intersection including the direction along the wall. ## 4 Boundary conditions and central charges For a $`1/4`$ BPS state, there are two sets of BPS equations, Eqs.(2.13)–(2.15) and (2.16)–(2.17), corresponding to the two kinds of BPS domain wall junctions. In this section we make explicit the relation between the boundary conditions and the choice of these BPS equations. BPS domain wall junction is formed when nonparallel BPS walls meet at a junction. In regions far away from the junction, the configuration approaches to isolated walls asymptotically. BPS domain wall is a $`1/2`$ BPS state and conserves two supercharges. These two supercharges are given, from Eqs.(2.11) and (2.12), in terms of central charges $`Z_1`$ and $`Z_2`$ for the wall. Let us take a general Wess-Zumino model in Eq.(2.3) and examine if a domain wall junction can be formed where $`N`$ different vacua appear in asymptotic regions. These $`N`$ vacua correspond to $`N`$ points in the complex plane of superpotential $`𝒲`$. The field configuration of the junction at infinity is mapped to a straight line connecting these $`N`$ vertices , . In order to have a balance of force, this polygon has to be convex . We set the origin of the $`𝒲`$ space at an arbitrary point inside this BPS polygon and denote the value of the superpotential at the $`I`$-th vacuum as $`𝒲_I`$, for $`I=1,\mathrm{},N`$, as illustrated in Fig. 3(b). Let us take the origin in $`x^1,x^2`$ space as the junction point of these BPS walls. If we denote $`\theta _{IJ}`$ the angle of the half wall separating two vacua, $`I`$ and $`J`$, as illustrated in Fig. 3(a), the central charges $`Z_1`$ and $`Z_2`$ of this wall are given by Eq.(2.4) as $`\stackrel{}{Z}_{IJ}(Z_1,Z_2)_{IJ}=2[𝒲_J^{}𝒲_I^{}]\stackrel{}{\omega }_{IJ}(\text{Area}),`$ (4.79) $`\stackrel{}{\omega }_{IJ}(\mathrm{cos}(\theta _{IJ}+\pi /2),\mathrm{sin}(\theta _{IJ}+\pi /2)).`$ (4.80) Thus two supercharges conserved at this wall are $$Q_1+e^{i(\alpha _{IJ}\theta _{IJ})}\overline{Q}_{\dot{1}},Q_2+e^{i(\alpha _{IJ}+\theta _{IJ})}\overline{Q}_{\dot{2}},$$ (4.81) where $`\alpha _{IJ}=\mathrm{arg}(𝒲_J𝒲_I)`$. BPS domain wall junction is a $`1/4`$ BPS state and conserves only one supercharge. Let us consider the case of $`H=H_{\mathrm{II}}`$ where a linear combination of $`Q_2`$ and $`\overline{Q}_{\dot{2}}`$ is conserved as shown in Eq.(2.12). This must be the common conserved supercharge for all the walls $$\mathrm{}=Q_2+e^{i(\alpha _{IJ}+\theta _{IJ})}\overline{Q}_{\dot{2}}=Q_2+e^{i(\alpha _{JK}+\theta _{JK})}\overline{Q}_{\dot{2}}=\mathrm{}.$$ (4.82) Then the relative angle of the two neighboring walls must be equal to the difference of two phases of the differences $`\mathrm{\Delta }𝒲`$ of the superpotentials for the two walls $$\mathrm{},\theta _{JK}\theta _{IJ}=\alpha _{JK}\alpha _{IJ},\mathrm{}.$$ (4.83) Moreover the field configuration at infinity should move counterclockwise in $`𝒲`$ space, as we go around the origin counterclockwise in $`x^1,x^2`$ space. Similarly, a linear combination of $`Q_1`$ and $`\overline{Q}_{\dot{1}}`$ is the common conserved supercharge in the case of $`H=H_\mathrm{I}`$. We obtain in this case $$\mathrm{},\theta _{JK}\theta _{IJ}=(\alpha _{JK}\alpha _{IJ}),\mathrm{}$$ (4.84) and that the field configuration at infinity should move clockwise in $`𝒲`$ space, as we go around the origin counterclockwise in $`x^1,x^2`$ space. Therefore we find that the BPS equations (2.16)–(2.17) for the case $`H=H_{\mathrm{II}}`$ should be used if the phase of the superpotential $`𝒲`$ increases as we go around the origin counterclockwise in $`x^1,x^2`$ space. If the phase of the superpotential $`𝒲`$ decreases as we go around the origin counterclockwise in $`x^1,x^2`$ space, the other BPS equations (2.13)–(2.15) for $`H=H_\mathrm{I}`$ should be used. Next we discuss the sign of the contribution of the central charge $`Y_3`$ to the mass of the junction configuration. We can use the Stokes theorem to obtain an expression for the central charge $`Y_3`$ as a contour integral $$Y_3=𝑑x^3id^2x\left[_1\left(K_i_2A^i\right)_2\left(K_i_1A^i\right)\right]=𝑑x^3iK_i𝑑A^i,$$ (4.86) where $`K_iK/A^i`$ is a derivative of the Kähler potential $`K`$. This contour integral in the field space should be done as a map from a counterclockwise contour in the infinity of $`z=x^1+ix^2`$ plane. Only complex fields can contribute to $`Y_3`$. Let us assume for simplicity that there is only one field which can contribute to $`Y_3`$ as in our exact solution. Eq.(4.86) shows that the central charge $`Y_3`$ becomes negative (positive), if the asymptotic counterclockwise contour in $`x^1,x^2`$ is mapped into a counterclockwise (clockwise) contour in field space. On the other hand, the sign of the contribution of the central charge $`Y_3`$ to the mass of the junction configurtion is determined by the formula $`H=H_{\mathrm{II}}=|iZ_1Z_2|+Y_3`$, or $`H=H_\mathrm{I}=|iZ_1Z_2|Y_3`$. The choice of these mass formulas are in turn deterined by the map of the asymptotic counterclockwise contour in $`x^1,x^2`$ space to a counterclockwise or clockwise contour in the superpotential space $`𝒲`$. Combining these two observations, we conclude that the contribution of the central charge $`Y_3`$ to the mass of the junction configuration is negative if the sign of rotations is the same in field space $`A^i`$ and in superpotential space $`𝒲`$, and positive if the sign of rotations is opposite. The field configuration moves counterclockwise in field space in our exact solution in (2.24) and then the central charge is negative in this solution. Since the exact solution satisfies the BPS equation for the case $`H=H_{\mathrm{II}}`$, the central charge contributes to the mass of the junction configuration negatively. Therefore we should not consider the central charge $`Y_3`$ alone as the physical mass of the junction at the center. In the junction configuration, the junction at the center cannot be separated from the walls. We also can find a solution for the other case of $`H=H_\mathrm{I}`$ in our model. The solution is just a configuration obtained by a reflection $`x^1x^1`$. Then the central charge is positive, but the contribution to the mass $`H=H_\mathrm{I}`$ becomes again negative. In either solution, the rotation in field $`T`$ space has the same sign as the rotation in superpotential $`𝒲`$ space. Therefore central charge $`Y_3`$ contributes negatively to the mass of the junction, irrespective of the choice of $`H=H_\mathrm{I}`$ or $`H=H_{\mathrm{II}}`$. More recently this feature of negative contribution of $`Y_3`$ to the junction mass is studied from a different viewpoint and it is argued that this feature is valid in most situations except possibly in contrived models . These models, if they exist, should correspond to the case of opposite sign of rotations in $`𝒲`$ space and field space. ## 5 Energy density and central charges ### 5.1 Charge densities Our exact solution is useful to examine how the topological charges $`Z_k`$, $`Y_k`$ and energy of the domain wall junction are distributed in $`x^1,x^2`$ space. We shall study their densities and integrated quantities in finite regions in this section. #### 5.1.1 $`Y`$ charge density The $`Y_3`$ charge density $`𝒴_3=iϵ^{3nm}_n\left(T^{}_mT\right)`$ is given in our exact solution by $`𝒴_3`$ $`=`$ $`24\mathrm{\Lambda }^4{\displaystyle \frac{e^{\sqrt{3}\mathrm{\Lambda }x^2}}{\left[e^{\sqrt{3}\mathrm{\Lambda }x^2}+2\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)\right]^3}}`$ (5.87) $`=`$ $`24\mathrm{\Lambda }^4{\displaystyle \frac{1}{\left[e^{\frac{2\mathrm{\Lambda }r}{\sqrt{3}}\mathrm{sin}\theta }+e^{\frac{2\mathrm{\Lambda }r}{\sqrt{3}}\mathrm{sin}\left(\theta +\frac{2\pi }{3}\right)}+e^{\frac{2\mathrm{\Lambda }r}{\sqrt{3}}\mathrm{sin}\left(\theta \frac{2\pi }{3}\right)}\right]^3}}`$ where the cylindrical coordinates $`r`$ and $`\theta `$ is used to make $`Z_3`$ symmetry explicit. A bird’s eye view of the $`𝒴_3`$ is given in Fig. 4. Here and the following, we shall take the unit of $`\mathrm{\Lambda }1`$ in drawing figures. The density is localized near the origin and the $`Z_3`$ symmetry is manifest. #### 5.1.2 $`Z`$ charge density We obtain the superpotential as the function of $`x^1`$ and $`x^2`$, by inserting the solution (2.24) $`\mathrm{Re}𝒲^{}`$ $`=`$ $`8\mathrm{\Lambda }^3{\displaystyle \frac{\left(2+3e^{\sqrt{3}\mathrm{\Lambda }x^2}\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)+\mathrm{cosh}\left(2\mathrm{\Lambda }x^1\right)\right)\mathrm{sinh}\left(\mathrm{\Lambda }x^1\right)}{\left[e^{\sqrt{3}\mathrm{\Lambda }x^2}+2\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)\right]^3}}`$ $`\mathrm{Im}𝒲^{}`$ $`=`$ $`2\sqrt{3}\mathrm{\Lambda }^3{\displaystyle \frac{e^{\sqrt{3}\mathrm{\Lambda }x^2}\left[2+e^{2\sqrt{3}\mathrm{\Lambda }x^2}+6e^{\sqrt{3}\mathrm{\Lambda }x^2}\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)\right]}{\left[e^{\sqrt{3}\mathrm{\Lambda }x^2}+2\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)\right]^3}}.`$ (5.88) The $`Z`$ charge densities are given by $`𝒵_k=2_k𝒲^{},(k=1,2)`$ and are found to be $`\mathrm{Re}𝒵_1`$ $`=`$ $`48\mathrm{\Lambda }^4{\displaystyle \frac{2+e^{2\sqrt{3}\mathrm{\Lambda }x^2}\mathrm{cosh}\left(2\mathrm{\Lambda }x^1\right)+3e^{\sqrt{3}\mathrm{\Lambda }x^2}\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)}{\left[e^{\sqrt{3}\mathrm{\Lambda }x^2}+2\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)\right]^4}}`$ $`\mathrm{Im}𝒵_1`$ $`=`$ $`\mathrm{Re}𝒵_2=48\sqrt{3}\mathrm{\Lambda }^4{\displaystyle \frac{e^{\sqrt{3}\mathrm{\Lambda }x^2}\mathrm{sinh}\left(\mathrm{\Lambda }x^1\right)\left(1+2e^{\sqrt{3}\mathrm{\Lambda }x^2}\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)\right)}{\left[e^{\sqrt{3}\mathrm{\Lambda }x^2}+2\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)\right]^4}}`$ $`\mathrm{Im}𝒵_2`$ $`=`$ $`48\mathrm{\Lambda }^4{\displaystyle \frac{e^{\sqrt{3}\mathrm{\Lambda }x^2}\left[\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)+e^{\sqrt{3}\mathrm{\Lambda }x^2}\left(2+3\mathrm{cosh}\left(2\mathrm{\Lambda }x^1\right)\right)\right]}{\left[e^{\sqrt{3}\mathrm{\Lambda }x^2}+2\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)\right]^4}}.`$ (5.89) We can define the effective value of the $`Z`$ charge which contributes to the energy of the junction as $`Z_{\mathrm{eff}}=\mathrm{Re}Z_1+\mathrm{Im}Z_2`$. Corresponding effective charge density is given by $$𝒵_{\mathrm{eff}}=96\mathrm{\Lambda }^4\frac{1+2e^{\sqrt{3}\mathrm{\Lambda }x^2}\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)+e^{2\sqrt{3}\mathrm{\Lambda }x^2}\left(1+2\mathrm{cosh}\left(2\mathrm{\Lambda }x^1\right)\right)}{\left[e^{\sqrt{3}\mathrm{\Lambda }x^2}+2\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)\right]^4}.$$ (5.90) Let us note that the effective charge density is $`Z_3`$ symmetric, whereas individual charges $`𝒵_1,𝒵_2`$ are not. #### 5.1.3 Energy density Adding $`𝒵_{\mathrm{eff}}`$ and $`𝒴_3`$ together, the energy density of the junction is obtained, $$=24\mathrm{\Lambda }^4\frac{4+6e^{\sqrt{3}\mathrm{\Lambda }x^2}\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)+e^{2\sqrt{3}\mathrm{\Lambda }x^2}\left(3+8\mathrm{cosh}\left(2\mathrm{\Lambda }x^1\right)\right)}{\left[e^{\sqrt{3}\mathrm{\Lambda }x^2}+2\mathrm{cosh}\left(\mathrm{\Lambda }x^1\right)\right]^4}$$ (5.91) A bird’s eye view of $``$ is shown in Fig. 5. The energy density is $`Z_3`$ symmetric as expected. A cross section of the densities, $`,𝒵_{\mathrm{eff}}`$ and $`𝒴_3`$ along one of the walls (e.g. negative $`x^2`$ direction) is shown in Fig. 6. The $`Z_{\mathrm{eff}}`$ charge contributes to the energy positively while $`Y_3`$ does negatively. Since the decrease of $`𝒵_{\mathrm{eff}}`$ is faster than the increase of $`𝒴_3`$, a small dip is found around $`R\mathrm{\Lambda }1`$. Far from the origin there is practically no difference between $`𝒵_{\mathrm{eff}}`$ and $``$ because $`𝒴_3`$ is localized near the origin. ### 5.2 Charge densities integrated over a region of finite radius In this subsection we shall evaluate the central charge densities integrated over a triangular or circular region depicted in Fig. 7. #### 5.2.1 $`Y_3`$ charge Integrating $`𝒴_3`$ over a triangle whose inscribed circle has a radius $`R`$ as shown in Fig. 7, we obtain for large $`R`$ ($`R\mathrm{\Lambda }1`$) $$Y_3^{\mathrm{triangle}}(R)=2\sqrt{3}\mathrm{\Lambda }^2L_3\left[1\frac{3\pi }{4}e^{\sqrt{3}R\mathrm{\Lambda }}+𝒪\left(e^{2\sqrt{3}R\mathrm{\Lambda }}\right)\right],$$ (5.92) where $`L_3`$ denotes the length along the $`x^3`$ direction. The leading term agrees with our previous evaluation by the step function approximation and the subleading terms vanish exponentially as $`R\mathrm{}`$. On the other hand, for small $`R`$ ($`R\mathrm{\Lambda }1`$), we obtain the $`Y_3`$ charge $$Y_3^{\mathrm{triangle}}(R)=\frac{8}{\sqrt{3}}\mathrm{\Lambda }^2L_3\left[R^2\mathrm{\Lambda }^2R^4\mathrm{\Lambda }^4\frac{4\sqrt{3}}{45}R^5\mathrm{\Lambda }^5+𝒪\left(R^6\mathrm{\Lambda }^6\right)\right].$$ (5.93) Notice that the leading term comes from the density at the origin multiplied by the area of the circle. Although there are no terms of the first or third degree in $`R`$, there is a fifth degree term. We can also integrate the $`𝒴_3`$ over a circle of radius $`R`$. In this case it is more convenient to use cylindrical coordinates ($`r,\theta `$, and $`x^3`$), and the following surface integral formula obtained from the Stokes theorem, $`Y_3^{\mathrm{circle}}(R)={\displaystyle _{L_3/2}^{L_3/2}}𝑑x^3{\displaystyle _0^{2\pi }}𝑑\theta \left[iT^{{}_{}{}^{}}{\displaystyle \frac{}{\theta }}T^{^{}}\right]`$ $`=`$ $`{\displaystyle \frac{8R\mathrm{\Lambda }^3L_3}{\sqrt{3}}}{\displaystyle _0^{2\pi }}𝑑\theta {\displaystyle \frac{\mathrm{sin}\left(\theta \frac{2\pi }{3}\right)e^{2R\mathrm{\Lambda }\mathrm{cos}\theta }+\mathrm{sin}\theta e^{2R\mathrm{\Lambda }\mathrm{cos}\left(\theta +\frac{2\pi }{3}\right)}+\mathrm{sin}\left(\theta +\frac{2\pi }{3}\right)e^{2R\mathrm{\Lambda }\mathrm{cos}\left(\theta \frac{2\pi }{3}\right)}}{\left[e^{\frac{2R\mathrm{\Lambda }}{\sqrt{3}}\mathrm{sin}\left(\theta +\frac{2\pi }{3}\right)}+e^{\frac{2R\mathrm{\Lambda }}{\sqrt{3}}\mathrm{sin}\left(\theta \frac{2\pi }{3}\right)}+e^{\frac{2R\mathrm{\Lambda }}{\sqrt{3}}\mathrm{sin}\theta }\right]^3}}`$ where the $`Z_3`$ symmetry is manifest. Expanding the integrand for small $`R(R\mathrm{\Lambda }1)`$, we obtain $$Y_3^{\mathrm{circle}}(R)=\frac{8\pi }{9}\mathrm{\Lambda }^2L_3\left[R^2\mathrm{\Lambda }^2\frac{1}{2}R^4\mathrm{\Lambda }^4+𝒪(R^6\mathrm{\Lambda }^6)\right].$$ (5.95) The leading term is again the density at the origin multiplied by the area of the circle. In contrast to the triangle case, there is no term of odd degree in $`R`$. For large $`R(R\mathrm{\Lambda }1)`$, we have to perform numerical integration to evaluate the $`Y_3^{\mathrm{circle}}(R)`$. We compare the $`Y_3(R)`$ evaluated for triangle, inscribed and circumscribed circle in Fig. 8. In the limit of $`R\mathrm{}`$, $`Y_3(R)`$ for all the regions converge to $`2\sqrt{3}\mathrm{\Lambda }^2L_3`$ as expected. #### 5.2.2 $`Z`$ charges Since the $`Z_1`$ charge is given by a total derivative in $`x^1`$, we can rewrite the $`Z_1`$ charge as $$Z_1(R)=2𝑑x^2𝑑x^1_1𝒲^{}(A^{})=2_{x^2}^{x^{2+}}𝑑x^2\left[𝒲^{}(x^{1+}\left(x^2\right),x^2)𝒲^{}(x^1\left(x^2\right),x^2)\right],$$ (5.96) where $`x^{i\pm }(i=1,2)`$ denote the upper and lower bound of the domain of integration. Since Eq.(5.88) shows that $`\mathrm{Im}𝒲^{}(x^1,x^2)`$ is even and $`\mathrm{Re}𝒲^{}(x^1,x^2)`$ is odd in $`x^1`$, we obtain $`\mathrm{Im}Z_1(R)=0`$ and $`\mathrm{Re}Z_2(R)=0`$ for an integration region symmetric in $`x^1`$ which we shall use. Let us note that $`\mathrm{Re}𝒵_1`$ ($`\mathrm{Im}𝒵_2`$) is negative (positive) definite. Firstly we choose as a domain of integration the triangle region whose inscribed circle has a radius $`R`$. For large $`R(R\mathrm{\Lambda }1)`$, we obtain $`\mathrm{Re}Z_1^{\mathrm{triangle}}(R)`$ $`=`$ $`12\mathrm{\Lambda }^2L_3\left[R\mathrm{\Lambda }+{\displaystyle \frac{\sqrt{3}\pi }{4}}e^{\sqrt{3}R\mathrm{\Lambda }}+𝒪\left(e^{2\sqrt{3}R\mathrm{\Lambda }}\right)\right]`$ $`\mathrm{Im}Z_2^{\mathrm{triangle}}(R)`$ $`=`$ $`12\mathrm{\Lambda }^2L_3\left[R\mathrm{\Lambda }+{\displaystyle \frac{\sqrt{3}}{3}}\left({\displaystyle \frac{\pi }{8}}{\displaystyle \frac{4}{3}}\right)e^{\sqrt{3}R\mathrm{\Lambda }}+𝒪\left(e^{2\sqrt{3}R\mathrm{\Lambda }}\right)\right].`$ (5.97) The leading linear term represents the contribution of charge density per unit length of the wall. It is interesting to observe that there are no constant terms. The exponentially suppressed terms represent the way the domain wall junction configuration converges to isolated walls as $`R\mathrm{}`$. The effective value of the $`Z`$ charge becomes $`Z_{\mathrm{eff}}^{\mathrm{triangle}}(R)`$ $`=`$ $`\mathrm{Re}Z_1^{\mathrm{triangle}}(R)+\mathrm{Im}Z_2^{\mathrm{triangle}}(R)`$ (5.98) $`=`$ $`24\mathrm{\Lambda }^3L_3R+\sqrt{3}\left({\displaystyle \frac{7\pi }{2}}{\displaystyle \frac{16}{3}}\right)\mathrm{\Lambda }^2L_3e^{\sqrt{3}R\mathrm{\Lambda }}+𝒪\left(e^{2\sqrt{3}R\mathrm{\Lambda }}\right).`$ For small $`R(R\mathrm{\Lambda }1)`$, $`Z`$ charges become $`\mathrm{Re}Z_1^{\mathrm{triangle}}(R)`$ $`=`$ $`{\displaystyle \frac{32}{\sqrt{3}}}\mathrm{\Lambda }^2L_3\left[R^2\mathrm{\Lambda }^2+{\displaystyle \frac{1}{2}}R^4\mathrm{\Lambda }^4{\displaystyle \frac{2\sqrt{3}}{9}}R^5\mathrm{\Lambda }^5+𝒪\left(R^6\mathrm{\Lambda }^6\right)\right]`$ $`\mathrm{Im}Z_2^{\mathrm{triangle}}(R)`$ $`=`$ $`{\displaystyle \frac{32}{\sqrt{3}}}\mathrm{\Lambda }^2L_3\left[R^2\mathrm{\Lambda }^2{\displaystyle \frac{1}{2}}R^4\mathrm{\Lambda }^4+{\displaystyle \frac{2\sqrt{3}}{9}}R^5\mathrm{\Lambda }^5+𝒪\left(R^6\mathrm{\Lambda }^6\right)\right].`$ (5.99) Notice that the leading term represents the density at the origin multiplied by the area of the triangle, and that there is no term of the first or third degree in $`R`$. The effective value of $`Z`$ charge is $$Z_{\mathrm{eff}}^{\mathrm{triangle}}(R)=\frac{64}{\sqrt{3}}R^2\mathrm{\Lambda }^4L_3+𝒪\left(R^6\mathrm{\Lambda }^6\right).$$ (5.100) We can also choose a circle of radius $`R`$ as a domain of integration. For small $`R(R\mathrm{\Lambda }1)`$ we obtain $`\mathrm{Re}Z_1^{\mathrm{circle}}(R)`$ $`=`$ $`{\displaystyle \frac{32\pi }{9}}\mathrm{\Lambda }^2L_3\left[R^2\mathrm{\Lambda }^2{\displaystyle \frac{1}{4}}R^4\mathrm{\Lambda }^4+𝒪\left(R^6\mathrm{\Lambda }^6\right)\right]`$ $`\mathrm{Im}Z_2^{\mathrm{circle}}(R)`$ $`=`$ $`{\displaystyle \frac{32\pi }{9}}\mathrm{\Lambda }^2L_3\left[R^2\mathrm{\Lambda }^2{\displaystyle \frac{1}{4}}R^4\mathrm{\Lambda }^4+𝒪\left(R^6\mathrm{\Lambda }^6\right)\right].`$ (5.101) The leading term is again given by the densities at the origin multiplied by the area of the circle. The effective value of the $`Z`$ charge is $$Z_{\mathrm{eff}}^{\mathrm{circle}}(R)=\frac{64\pi }{9}\mathrm{\Lambda }^2L_3\left[R^2\mathrm{\Lambda }^2\frac{1}{4}R^4\mathrm{\Lambda }^4+𝒪\left(R^6\mathrm{\Lambda }^6\right)\right].$$ (5.102) A numerical evaluation is needed for large $`R`$. We compare the effective $`Z`$ value evaluated for triangle, inscribed circle and circumscribed circle in Fig. 9. As $`R\mathrm{}`$, the asymptotic slope of $`Z_{\mathrm{eff}}^{\mathrm{ins}.\mathrm{circle}}(R)`$ for inscribed circle converges to the same value as that for the triangle. In the case of the circumscribed circle, the $`Z_{\mathrm{eff}}`$ becomes twice as large as those of the other cases for large $`R`$, since the total length of the walls is twice as long as those of the other cases. #### 5.2.3 Energy of the junction Since our exact solution satisfies the BPS equation corresponding to $`H=H_{\mathrm{II}}`$, the energy of the junction is obtained by adding $`Y_3`$ and $`Z_{\mathrm{eff}}`$ together $`H=Z_{\mathrm{eff}}+\mathrm{Y}_3=\mathrm{Re}Z_1+\mathrm{Im}Z_2+\mathrm{Y}_3`$. Firstly we choose the triangle region whose inscribed circle has radius $`R`$. For large $`R(R\mathrm{\Lambda }1)`$, the energy is $$H_{\mathrm{triangle}}(R)=24\mathrm{\Lambda }^3L_3R2\sqrt{3}\mathrm{\Lambda }^2L_3+\sqrt{3}\left(5\pi \frac{16}{3}\right)\mathrm{\Lambda }^2L_3e^{\sqrt{3}R\mathrm{\Lambda }}+𝒪\left(e^{2\sqrt{3}R\mathrm{\Lambda }}\right).$$ (5.103) The first linear term can be regarded as the contribution from the walls and the second constant term can be regarded as the contribution from the junction at the center. For small $`R(R\mathrm{\Lambda }1)`$, the energy is $$H_{\mathrm{triangle}}(R)=\frac{56}{\sqrt{3}}\mathrm{\Lambda }^2L_3\left[R^2\mathrm{\Lambda }^2+\frac{1}{7}R^4\mathrm{\Lambda }^4+\frac{4\sqrt{3}}{315}R^5\mathrm{\Lambda }^5+𝒪\left(R^6\mathrm{\Lambda }^6\right)\right].$$ (5.104) In the case of the circle of radius $`R`$, the energy is given for small $`R(R\mathrm{\Lambda }1)`$ as $$H_{\mathrm{circle}}(R)=\frac{56\pi }{9}\mathrm{\Lambda }^2L_3\left[R^2\mathrm{\Lambda }^2\frac{3}{14}R^4\mathrm{\Lambda }^4+𝒪\left(R^6\mathrm{\Lambda }^6\right)\right].$$ (5.105) The energy of the triangle region is compared to those of inscribed and circumscribed circles in Fig. 10. For large $`R(R\mathrm{\Lambda }1)`$, the energy $`H`$ reduces to $`Z_{\mathrm{eff}}`$. Finally we plot the $`\theta `$–dependence of the energy and charges that are obtained by integrating the densities from $`r=0`$ to $`r=R`$ with $`\theta `$ fixed (see Fig. 11). In each figure the energy $`H`$ (solid line) is the sum of the $`Z_{\mathrm{eff}}`$ (short dashed line) and $`Y_3`$ (long dashed line) and $`Z_3`$ symmetry is manifest in their $`\theta `$–dependence. In Fig. 11(a), all the quantities are almost uniform in $`\theta `$ near the junction at the center, reflecting the fact that the junction is a string–like object and symmetric around the $`x^3`$ axis. As we move away from the origin, main contribution comes from the direction of the walls (in our case $`\pi /2,\pi /6,`$ and $`5\pi /6`$) (see Fig. 11(b) and (c)). As $`R`$ grows, $`Y_3`$ disappears and the energy $`H`$ approaches $`Z_{\mathrm{eff}}`$ (see Fig. 11(d)). ### Acknowledgments One of the authors (H.O.) gratefully acknowledges support from the Iwanami Fujukai Foundation. This work is supported in part by Grant-in-Aid for Scientific Research from the Japan Ministry of Education, Science and Culture for the Priority Area 291 and 707. ## A Fermionic contributions to central charges We shall derive the central charges including fermionic contributions in the case of a general Wess-Zumino model with an arbitrary superpotential $`𝒲`$. For simplicity, Kähler metric is assumed to be minimal $`K_{ij^{}}=\delta _{ij^{}}`$. $``$ $`=`$ $`_\mu A^j^\mu A^j+F^jF^j+{\displaystyle \frac{i}{2}}_\mu \overline{\psi }^j\overline{\sigma }^\mu \psi ^j{\displaystyle \frac{i}{2}}\overline{\psi }^j\overline{\sigma }^\mu _\mu \psi ^j`$ (A.1) $`+`$ $`F^j{\displaystyle \frac{𝒲}{A^j}}{\displaystyle \frac{1}{2}}\psi ^i\psi ^j{\displaystyle \frac{𝒲}{A^iA^j}}+F^j{\displaystyle \frac{𝒲^{}}{A^j}}{\displaystyle \frac{1}{2}}\overline{\psi }^i\overline{\psi }^j{\displaystyle \frac{𝒲^{}}{A^iA^j}}.`$ We have added a surface term to Eq.(2.3) to make the variational principle meaningful. This is the starting Lagrangian to derive central charges and we will not neglect any total divergences from now on. The canonical supercurrent is found to be $`J_\alpha ^\mu `$ $`=`$ $`\sqrt{2}\left(\sigma ^\nu \overline{\sigma }^\mu \psi ^i\right)_\alpha _\nu A^i+i\sqrt{2}\left(\sigma ^\mu \overline{\psi }^i\right)_\alpha F^i`$ (A.2) $`=`$ $`\sqrt{2}\left(\sigma ^\nu \overline{\sigma }^\mu \psi ^i\right)_\alpha _\nu A^ii\sqrt{2}\left(\sigma ^\mu \overline{\psi }^i\right)_\alpha {\displaystyle \frac{𝒲^{}}{A^i}}`$ $`\overline{J}^{\mu \dot{\alpha }}`$ $`=`$ $`\sqrt{2}\left(\overline{\sigma }^\nu \sigma ^\mu \overline{\psi }^i\right)^{\dot{\alpha }}_\nu A^i+i\sqrt{2}\left(\overline{\sigma }^\mu \psi ^i\right)^{\dot{\alpha }}F^i`$ (A.3) $`=`$ $`\sqrt{2}\left(\overline{\sigma }^\nu \sigma ^\mu \overline{\psi }^i\right)^{\dot{\alpha }}_\nu A^ii\sqrt{2}\left(\overline{\sigma }^\mu \psi ^i\right)^{\dot{\alpha }}{\displaystyle \frac{𝒲}{A^i}}`$ The canonical energy momentum tensor is given by $`T^{\mu \nu }=^\mu A^j^\nu A^j+^\mu A^j^\nu A^j+{\displaystyle \frac{i}{2}}\overline{\psi }^j\overline{\sigma }^\mu ^\nu \psi ^j+{\displaystyle \frac{i}{2}}\psi ^j\sigma ^\mu ^\nu \overline{\psi }^j`$ (A.4) $`+g^{\mu \nu }\left[_\lambda A^j^\lambda A^j\left|{\displaystyle \frac{𝒲}{A^j}}\right|^2+{\displaystyle \frac{i}{2}}_\lambda \overline{\psi }^j\overline{\sigma }^\lambda \psi ^j{\displaystyle \frac{i}{2}}\overline{\psi }^j\overline{\sigma }^\lambda _\lambda \psi ^j{\displaystyle \frac{1}{2}}\psi ^i\psi ^j{\displaystyle \frac{𝒲}{A^iA^j}}{\displaystyle \frac{1}{2}}\overline{\psi }^i\overline{\psi }^j{\displaystyle \frac{𝒲^{}}{A^iA^j}}\right].`$ Canonical quantization gives (anti-) commutation relations $$[A^i(x),_0A^j(y)]_{x^0=y^0}=i\delta ^3(xy)\delta ^{ij},\{\psi _\alpha ^i(x),\overline{\psi }_{\dot{\beta }}^j(y)\}_{x^0=y^0}=\delta ^3(xy)\delta ^{ij}\sigma _{\alpha \dot{\beta }}^0.$$ (A.5) The anticommutator between supercharges of the same chirality gives the supersymmetry algebra (2.1) with the central charge $`Z_k`$ in Eq.(2.4) $$Z_k=2d^3x_k𝒲^{}(A^{})$$ (A.6) which turns out to have only bosonic contributions. The anticommutator between supercharges of the opposite chirality gives the supersymmetry algebra (2.2) with the central charge $`Y_k`$ $$Y_k=iϵ^{knm}d^3x_n\left(A^j_mA^j\frac{1}{2}\overline{\psi }^j\overline{\sigma }_m\psi ^j\right),ϵ^{123}=1,$$ (A.7) which has both bosonic and fermionic contributions. ## B Gamma matrices and fermion mode equations In order to separate variables $`(x^1,x^2)`$ and $`(x^0,x^3)`$ for spinors, it is most convenient to use a gamma matrix representation where the direct product structure of $`2\times 2`$ matrices in $`(x^1,x^2)`$ and $`(x^0,x^3)`$ becomes manifest. One such representation is $$\gamma ^0=\rho ^1,\gamma ^3=i\rho ^2,\gamma ^1=i\sigma ^1\rho ^3,\gamma ^2=i\sigma ^2\rho ^3,$$ (B.1) where $`\sigma ^a`$ are Pauli matrices acting on $`2\times 2`$ matrices and $`\rho ^a`$ acting on indices of blocks of these $`2\times 2`$ matrices. The four component spinor can be decomposed into a pair of two component spinors $`\xi `$ and $`\chi `$ in $`0+1`$ dimensions $$\psi =\left[\begin{array}{c}\xi \\ i\chi \end{array}\right],$$ (B.2) The $`B`$ matrix for $`1+3`$ dimensions can be defined as a product of $`B`$ matrices $`B^{(1)}`$ for $`1+1`$ dimensions and $`B^{(2)}`$ for $`0+2`$ dimensions $$B=B^{(1)}B^{(2)},B^{(1)}=\rho ^3,B^{(2)}=i\sigma ^1.$$ (B.3) The Majorana condition for the $`1+3`$ dimensional spinor and the pseudo-Majorana condition for the $`0+2`$ dimensional spinor are given by $$\psi =B\psi ^{},\xi =B^{(2)}\xi ^{},\chi =B^{(2)}\chi ^{},$$ (B.4) which implies $`\xi _1=i\xi _2^{}`$ for components of the two component spinor $`\xi =(\xi _1,\xi _2)^T`$, and similarly for $`\chi `$. The Dirac equation for four component fermions in the general Wess-Zumino model reads $$i\gamma ^\mu _\mu \psi ^i\left(\frac{^2𝒲}{A^iA^j}\frac{1+i\gamma ^5}{2}+\frac{^2𝒲^{}}{A^iA^j}\frac{1i\gamma ^5}{2}\right)\psi ^j=0.$$ (B.5) The mode equation in $`x^1,x^2`$ space is defined in terms of two component spinors $`\xi _n^j`$ and $`\chi _n^j`$ $`\left[\delta ^{ij}\left(\sigma ^1_1+\sigma ^2_2\right){\displaystyle \frac{^2𝒲}{A^iA^j}}{\displaystyle \frac{1\sigma ^3}{2}}{\displaystyle \frac{^2𝒲^{}}{A^iA^j}}{\displaystyle \frac{1+\sigma ^3}{2}}\right]\xi _n^j`$ $`=`$ $`m_n^{(1)}\chi _n^i,`$ $`\left[\delta ^{ij}\left(\sigma ^1_1+\sigma ^2_2\right){\displaystyle \frac{^2𝒲}{A^iA^j}}{\displaystyle \frac{1+\sigma ^3}{2}}{\displaystyle \frac{^2𝒲^{}}{A^iA^j}}{\displaystyle \frac{1\sigma ^3}{2}}\right]\chi _n^j`$ $`=`$ $`m_n^{(2)}\xi _n^i.`$ (B.6) The two component spinors $`\xi _n^i,\chi _n^i`$ satisfy the pseudo-Majorana condition (B.4). Then we find that mass eigenvalues are real $$m_n^{(1)}=\left(m_n^{(1)}\right)^{},m_n^{(2)}=\left(m_n^{(2)}\right)^{}.$$ (B.7) We can make a separation of variables for the Dirac equation (B.5) by means of real fermionic fields $`c_n`$ and $`b_n`$ $$\psi ^i=\underset{n}{}\left[\begin{array}{c}c_n(x^0,x^3)\xi _n(x^1,x^2)\\ ib_n(x^0,x^3)\chi _n(x^1,x^2)\end{array}\right].$$ (B.8) Using these mode functions we find that the fermion fields $`(c_n,ib_n)^T`$ satisfy the Dirac equation in $`1+1`$ dimensions in Eq.(3.46). This representation can be related to the usual Weyl representation in ref. by the following unitary matrix $`U`$ $$U=\frac{1\sigma ^3}{2}\rho ^3+\frac{1+\sigma ^3}{2}\rho ^2,\gamma _{\mathrm{Weyl}}^\mu =U^{}\gamma ^\mu U,B_{\mathrm{Weyl}}=U^{}BU^{}=\sigma ^2\rho ^2.$$ (B.9) The two component spinor in the Weyl representation is related to the components of the four component spinor (B.2) in the representation (B.1) in this appendix as $$\left[\begin{array}{c}\psi _\alpha \\ \overline{\psi }^{\dot{\alpha }}\end{array}\right]_{\mathrm{Weyl}}=\left[\begin{array}{c}\chi _1\\ \xi _2\\ i\xi _1\\ i\chi _2\end{array}\right].$$ (B.10) Thus we obtain the mode equation (3.41)– (3.44) in the Weyl representation.
warning/0004/hep-ph0004070.html
ar5iv
text
# 1 Introduction ## 1 Introduction Non-topological solitons, in particular Q–balls, are extended objects with finite mass and spatial extension, and arise in scalar field theories when there is an exact continious symmetry and some kind of attractive interaction, as already classified by Coleman <sup>?</sup>. As first pointed out by Kusenko <sup>?</sup>, Q-balls naturally exist in supersymmetric theories thanks to global baryon (B) and lepton (L) number symmetries. Besides, theories with an extended scalar sector can support non–baryonic Q–balls <sup>?</sup>. Q–balls have found applications in modelling several physical processes ranging from leptoquarks to dark matter <sup>?,?</sup>. It is well known that the MSSM has unremovable physical phases which can be identified with the phases of $`\mu `$ parameter and $`A`$ terms <sup>?</sup>. According to the vacuum stability arguments these phases relax to CP conserving points <sup>?</sup>. However, the same arguments are not sufficient to relax the phases in the minimal extensions <sup>?</sup>. Hence, it is plausable to take these phases finite and look for their effects in low energy processes. As summarized above, the MSSM predicts the existence of both non–topological solitons and finite CP violation. However, so far the investigations on the Q–ball formation in the MSSM have not dealt with the effects of the CP violation. In this short note we investigate the effects of explicit CP violation in the MSSM on the Q–ball formation. In the next section we discuss this issue in detail using the MSSM scalar potential. We particularly analyze the effects of the phases in the trilinear couplings and the $`\mu `$ parameter. In the last section we summarize the main findings. ## 2 MSSM with explicit CP violation and Q-balls The MSSM scalar sector contains two Higgs doublets (with opposite hypercharge) and scalar partners of quarks and leptons. The supersymmetry and gauge symmetry are broken by the soft supersymmetry breaking terms which introduce a number of mass parameters to the potential. Both the Higgsino Dirac mass term $`\mu `$ and the trilinear couplings in soft terms are complex, and they lead to CP violation beyond the CKM matrix already present in the SM. Denoting the neutral components of the Higgs doublets as $`\varphi _1`$ and $`\varphi _2`$, the MSSM scalar potential for one generation of sfermions reads as: $`V_{MSSM}`$ $`=`$ $`[(h_uA_u\varphi _2h_u\mu ^{}\varphi _1^{})\stackrel{~}{u}_L\stackrel{~}{u}_R^{}(h_dA_d\varphi _1h_d\mu ^{}\varphi _2^{})\stackrel{~}{d}_L\stackrel{~}{d}_R^{}`$ (1) $`(h_eA_e\varphi _1h_e\mu ^{}\varphi _2^{})\stackrel{~}{e}_L\stackrel{~}{e}_R^{}+h.c.]`$ $`+m_{\stackrel{~}{Q}}^2|\stackrel{~}{Q}|^2+m_{\stackrel{~}{u}}^2|\stackrel{~}{u}_R|^2+m_{\stackrel{~}{d}_R}^2|\stackrel{~}{d}_R|^2+m_{\stackrel{~}{L}}^2|\stackrel{~}{L}|^2+m_{\stackrel{~}{e}_R}^2|\stackrel{~}{e}_R|^2`$ $`+m_1^2|\varphi _1|^2+m_2^2|\varphi _2|^2+|\mu |^2|\varphi _2|^2+|\mu |^2|\varphi _1|^2+h_e^2|\stackrel{~}{L}^2||\stackrel{~}{e}_R|^2`$ $`+h_d^2|\stackrel{~}{Q}^2||\stackrel{~}{d}_R|^2+h_e^2|\stackrel{~}{L}^2||\stackrel{~}{e}_R|^2+h_u^2|\stackrel{~}{Q}^2||\stackrel{~}{u}_R|^2+h_e^2|\varphi _1|^2|\stackrel{~}{e}_R|^2`$ $`+h_d^2|\varphi _1|^2|\stackrel{~}{d}_R|^2+h_e^2|\varphi _1\stackrel{~}{e}_L|^2+h_u^2|\varphi _2\stackrel{~}{u}_L|^2+h_d^2|\varphi _1\stackrel{~}{d}_L|^2`$ plus the $`D`$ term contributions which will not be shown explicitely. This scalar potential has two global symmetries: $`U(1)_B`$ and $`U(1)_L`$ corresponding to baryon number and lepton number symmetries, respectively. These are the exact symmetries of the theory and their breaking (spontaenous and otherwise) lead to B– and L– violating processes. As was emphasized in Refs. 1 and 2 it is mainly the cubic couplings that generate B–ball or L–ball type solitonic solutions. It is appearent that slepton doublet $`\stackrel{~}{L}`$ and right–handed slepton $`\stackrel{~}{e}_R`$ contribute to L–balls whereas one needs the squark doublet $`\stackrel{~}{Q}`$ and right–handed squarks $`\stackrel{~}{u}_R`$ and $`\stackrel{~}{d}_R`$ to form B–balls. In both cases Higgs fields are necessary. In this form the scalar potential involves several scalar fields, and a true analysis of the Q–ball formation requires a minimization of the multi–field quantity <sup>?</sup> $`m_{eff}^2(\varphi _1,\mathrm{},\stackrel{~}{e}_R)2V_{MSSM}/{\displaystyle \underset{Ball}{}}charge\times |field|^2`$ (2) that guarantees the stability of the Q–ball against decaying into its constitutents. In this formula $`charge`$ and $`field`$ denote the baryon number (lepton number) and squark fields (slepton fields). Instead of dealing with coupled equations of motion for fields contributing to a particular Q–ball, practically one can describe the nature of the Q–ball by using a single scalar degree of freedom <sup>?</sup>. This approximation is quite accurate especially for $`D`$– and $`F`$–flat potentials as the degrees of freedom orthogonal to flat directions will be much more massive <sup>?,?</sup>. For this purpose it is convenient to introduce a scalar field $`\phi `$ representing the Q-ball matter, and decompose the component fields in terms of $`\phi `$ using the dimensionless parameters $`\xi _i`$, with $`\xi <1`$ and $`_i\xi _i^2=1`$, as follows: $`\stackrel{~}{e}_{L,R}=\xi _{e_{L,R}}\phi ,\stackrel{~}{u}_{L,R}=\xi _{u_{L,R}}\phi ,\stackrel{~}{d}_{L,R}=\xi _{d_{L,R}}\phi ,\varphi _{1,2}=\xi _{1,2}\phi .`$ (3) Using this decomposition, the scalar potential $`V_{MSSM}`$ takes the form $`V_\phi =M_\phi ^2|\phi |^2+M_cRe\phi |\phi |^2M_sIm\phi |\phi |^2+\lambda |\phi |^4`$ (4) for both $`L`$– and $`B`$–balls. Here $`M_\phi ^2`$ is the linear combination of the scalar quadratic mass parameters in $`V_{MSSM}`$. The quartic coupling $`\lambda `$ is a linear combination of Yukawa couplings $`h_{u,d,e}`$ and gauge couplings $`g_{3,2,1}`$ following, respectively, from the $`F`$–term and $`D`$–term contributions. As the general analyses of Q–ball formation Refs. 1, 2, 3 show explicitly, the crucial parameters in Eq. (4) are the trilinear mass parameters $`M_c`$ and $`M_s`$. For $`L`$–balls one has $`M_c`$ $`=`$ $`2\left[\mathrm{cos}(\varphi _{A_e})\stackrel{~}{A}_e+\mathrm{cos}(\varphi _\mu )\stackrel{~}{\mu }\right]`$ (5) $`M_s`$ $`=`$ $`2\left[\mathrm{sin}(\varphi _{A_e})\stackrel{~}{A}_e+\mathrm{sin}(\varphi _\mu )\stackrel{~}{\mu }\right]`$ (6) where $`\stackrel{~}{A_e}`$ $`=`$ $`h_e|A_e|\xi _1\xi _{e_L}\xi _{e_R}`$ $`\stackrel{~}{\mu }`$ $`=`$ $`h_e|\mu |\xi _2\xi _{e_L}\xi _{e_R}`$ (7) where $`\varphi _{A_e}`$ and $`\varphi _\mu `$ are the phases of $`A_e`$ and $`\mu `$ parameter, respectively. For $`B`$–balls, however, $`M_c`$ and $`M_s`$ have the following expressions $`M_c`$ $`=`$ $`2\left[\mathrm{cos}(\varphi _{A_u})\stackrel{~}{A_u}\mathrm{cos}(\varphi _{A_d})\stackrel{~}{A_d}+\mathrm{cos}(\varphi _\mu )\stackrel{~}{\mu }\right]`$ (8) $`M_s`$ $`=`$ $`2\left[\mathrm{sin}(\varphi _{A_u})\stackrel{~}{A_u}\mathrm{sin}(\varphi _{A_d})\stackrel{~}{A_d}+\mathrm{sin}(\varphi _\mu )\stackrel{~}{\mu }\right]`$ (9) where $`\stackrel{~}{A_u}`$ $`=`$ $`h_u|A_u|\xi _2\xi _{u_L}\xi _{u_R}`$ $`\stackrel{~}{A_d}`$ $`=`$ $`h_d|A_d|\xi _1\xi _{d_L}\xi _{d_R}`$ $`\stackrel{~}{\mu }`$ $`=`$ $`|\mu |(h_d\xi _2\xi _{d_L}\xi _{d_R}h_u\xi _1\xi _{u_L}\xi _{u_R}).`$ (10) The form of the scalar potential Eq. (4) is such that $`U(1)_B`$ or $`U(1)_L`$ symmetries are not manifest at all. In particular, $`Im\phi `$ and $`Re\phi `$ refer to Higgs fields $`\varphi _{1,2}`$ which do not contribute to the charge of the Q–ball. For instance, in thin–wall approximation and in the notation of Ref. 2, the total charge and the effective mass of the $`B`$–ball are respectively given by $`B`$ $`=`$ $`2a\omega |\phi |^2V`$ $`m_{eff}^2`$ $`=`$ $`{\displaystyle \frac{1}{a}}\left[M_\phi ^2+M_cRe\phi M_sIm\phi +\lambda \{(Re\phi )^2+(Im\phi )^2\}\right]`$ $`\mathrm{where}\mathrm{a}`$ $`=`$ $`{\displaystyle \frac{1}{3}}\left[\xi _{u_L}^2+\xi _{u_R}^2+\xi _{d_L}^2+\xi _{d_R}^2\right]`$ (11) Thus, the total charge vanishes if squarks are absent; that is, the Higgs fields do not play a role in charge accumulation. However, there is no stable Q–ball if the Higgs fields are absent, as can be seen from vanishing of the trilinear couplings $`M_c`$ and $`M_s`$. In this sense, $`Re`$ and $`Im`$ parts of $`\phi `$ in Eq. (4) refer to the time–independent phase of $`\phi `$ generated by the non–trivial phases in $`\mu `$ and $`A_{u,d,e}`$ parameters. As is seen from Eq. (4), the main effect of complex $`\mu `$ and $`A_{u,d,e}`$ is to introduce $`M_s0`$ which is proportional to $`Im\phi `$. There are three distinct limits in which the Q–ball matter gains different CP characteristics depending on the values of $`M_c`$ and $`M_s`$. So far discussions of Q–balls have been based on purely real $`\mu `$ and $`A_{u,d,e}`$ in which case $`M_s0`$. Then only $`Re\phi `$ has a trilinear coupling and stable Q–matter is thus composed of $`Re\phi `$ which is purely CP even. In the opposite limit, that is, for purely imaginary $`\mu `$ and $`A_{u,d,e}`$ one has $`M_c0`$. Then only $`Im\phi `$ has a trilinear coupling, and thus, the resulting Q–ball is composed of $`Im\phi `$, and Q–matter is purely CP odd. In the general case, where $`\mu `$ and $`A_{u,d,e}`$ are complex parameters with nonvanishing real and imaginary parts, both $`Im\phi `$ and $`Re\phi `$ contribute to the Q–matter. Namely, $`m_{eff}^2`$ is minimized for $`Re\phi =M_c/2\lambda `$ and $`Im\phi =M_s/2\lambda `$, so that mass per unit charge for $`B`$–ball reads as $`m_{eff}^2(Bball)={\displaystyle \frac{1}{a}}\left[M_\phi ^2{\displaystyle \frac{1}{4\lambda }}(M_c^2+M_s^2)\right].`$ (12) According to Coleman’s theorem <sup>?</sup>, if $`0<m_{eff}^2<M_\phi ^2/a`$ then the resulting B-ball is stable. One also notices that $`m_{eff}^2`$ depends explicitly on the relative phase between any pair of $`\stackrel{~}{A}_d,\stackrel{~}{A}_u`$ and $`\stackrel{~}{\mu }`$, using Eq.(8): $`m_{eff}^2(Bball)={\displaystyle \frac{1}{a}}[M_\phi ^2{\displaystyle \frac{1}{\lambda }}(\stackrel{~}{A_u}^2+\stackrel{~}{A_d}^2+\stackrel{~}{\mu }^22\stackrel{~}{A_u}\stackrel{~}{A_d}\mathrm{cos}(\varphi _{A_u}\varphi _{A_d})`$ $`+2\stackrel{~}{A_u}\stackrel{~}{\mu }\mathrm{cos}(\varphi _{A_u}\varphi _\mu )2\stackrel{~}{A_d}\stackrel{~}{\mu }\mathrm{cos}(\varphi _{A_d}\varphi _\mu ))]`$ (13) Depending on the values of these CP phases $`m_{eff}^2`$ takes on a range of values. Therefore, mass per unit charge in the $`B`$–ball varies with the CP violating phases. To illustrate this, one can consider the simple case of $`\stackrel{~}{u}_L\stackrel{~}{u}_R`$ $`B`$–ball, that is, $`A_d=0`$. In this case $`m_{eff}^2`$ varies from its minimum value ($`\varphi _{A_u}\varphi _\mu =0`$) $`\left[m_{eff}^2(Bball)\right]_{min}={\displaystyle \frac{1}{a}}\left[M_\phi ^2{\displaystyle \frac{1}{\lambda }}\left(\stackrel{~}{A_u}+\stackrel{~}{\mu }\right)^2\right]`$ (14) to the maximal value ($`\varphi _{A_u}\varphi _\mu =\pi `$) $`\left[m_{eff}^2(Bball)\right]_{max}={\displaystyle \frac{1}{a}}\left[M_\phi ^2{\displaystyle \frac{1}{\lambda }}\left(\stackrel{~}{A_u}\stackrel{~}{\mu }\right)^2\right]`$ (15) with the mean ($`\varphi _{A_u}\varphi _\mu =\pi /2`$) $`\left[m_{eff}^2(Bball)\right]_{mean}={\displaystyle \frac{1}{a}}\left[M_\phi ^2{\displaystyle \frac{1}{\lambda }}\left(\stackrel{~}{A_u}^2+\stackrel{~}{\mu }^2\right)\right]`$ (16) taking $`\stackrel{~}{\mu }`$ positive. One notices that none of the conditions above implies a specific value for the phases $`\varphi _{A_u}`$ and $`\varphi _\mu `$. Indeed the mass parameters $`M_c`$ and $`M_s`$ take on the following values $`M_c=2\left(\stackrel{~}{A_u}+\stackrel{~}{\mu }\right)\mathrm{cos}\varphi _{A_u},M_s=2\left(\stackrel{~}{A_u}+\stackrel{~}{\mu }\right)\mathrm{sin}\varphi _{A_u}`$ $`M_c=2\left(\stackrel{~}{A_u}\stackrel{~}{\mu }\right)\mathrm{cos}\varphi _{A_u},M_s=2\left(\stackrel{~}{A_u}\stackrel{~}{\mu }\right)\mathrm{sin}\varphi _{A_u}`$ (17) $`M_c=2\left(\stackrel{~}{A_u}\mathrm{cos}\varphi _{A_u}+\stackrel{~}{\mu }\mathrm{sin}\varphi _{A_u}\right),M_s=2\left(\stackrel{~}{A_u}\mathrm{sin}\varphi _{A_u}\stackrel{~}{\mu }\mathrm{cos}\varphi _{A_u}\right)`$ for $`\varphi _{A_u}\varphi _\mu =0`$, $`\pi `$ and $`\pi /2`$, respectively. Hence, despite the phase independence of mass per unit charge (14)–(16), the value of the condensate, determined by $`M_c`$ and $`M_s`$, is an explicit function of the CP phases. This follows from the fact that the Q–matter $`\phi `$ depends on the individual phases additively whereas $`m_{eff}^2`$ depends only on the relative phases. This particular pattern of phase structure can be important for Q–ball formation. In the purely CP–conserving limit, one would have $`m_{eff}^2=\left[m_{eff}^2(Bball)\right]_{min}`$ together with $`M_c=2\left(\stackrel{~}{A_u}+\stackrel{~}{\mu }\right)`$ and $`M_s=0`$, correspondig to the first line of (2) with $`\varphi _{A_u}=0`$. However, as the Eqs. (14)–(16) and (2) suggest clearly the nonvanishing CP phases offer more alternatives. To see the implications of such a phase dependence, one can consider the special case of $`\stackrel{~}{A_u}=\stackrel{~}{\mu }`$ and $`\stackrel{~}{\mu }^2/\lambda =M_\phi ^2/4`$. Then, for $`\varphi _{A_u}\varphi _\mu =0`$, one has $`m_{eff}^2=0`$ with nonvanishing $`M_c`$ and $`M_s`$. Therefore, for this parameter set one has a massless Q–ball, or equivalently, a Q–matter distribution over entire space. In this case, at least in the thin–wall approximation, there is no macroscopic structure with finite size. On the other hand, for $`\varphi _{A_u}\varphi _\mu =\pi `$, $`m_{eff}^2=M_\phi ^2/a`$, and this corresponds to the critical value of $`m_{eff}^2`$ below which there would be a stable Q–ball solution. A careful look at $`M_c`$ and $`M_s`$ shows that they vanish identically and Q–ball formation without the trilinear couplings is already out of question. Therefore this possibility leaves no room for Q–ball formation. Finally, for $`\varphi _{A_u}\varphi _\mu =\pi /2`$, however, one obtaines $`m_{eff}^2=M_\phi ^2/2a`$ leaving both $`M_c`$ and $`M_s`$ nonvanishing. This is a regular Q–ball solution, and it corresponds to maximal CP–violation, for instance, in the Higgs sector <sup>?</sup>. Indeed, $`M_c`$ and $`M_s`$ can vary with $`\varphi _{A_u}`$ further and this does not affect the Q–ball structure obtained for $`\left[m_{eff}^2(Bball)\right]_{mean}`$. Therefore, depending on the amount of CP violation, there may be regions of the parameter space that support a stable Q–ball solution though the strictly CP–conserving MSSM does not. Another implication of these phases would be on the scattering of fermions from the Q–ball. Indeed, a typical cross section has the form $`\sigma 4\pi R^2`$ where $`R^21/m_{eff}^2`$. Therefore, the supersymmetric CP phases influence the formation as well as interactions of the Q–balls with surrounding plasma. In fact, this expectation is confirmed by the recent analysis of the scattering of the dark matter particles from the nucleons <sup>?</sup>. So far we have discussed Q–ball formation without imposing any $`D`$– and/or $`F`$–flatness. However, the scalar potential of the low–energy supersymmetric theories has many flat directions. Such flat potentials are phenomenologically relevant as it is possible to produce large enough Q–balls that can resist the evaporation on time scales of the order of the age of the universe <sup>?,?</sup>. As has been listed in Ref. , there are slepton as well as squark flat directions with corresponding Q–balls. In fact, radiatively corrected flat directions in the MSSM induce a potential of the form $`V_\phi =M_\phi ^2|\phi |^2+\lambda ^2{\displaystyle \frac{|\phi |^{2(d1)}}{M_{Pl}^{2(d3)}}}+(\lambda A{\displaystyle \frac{\phi ^d}{M_{Pl}^{(d3)}}}+h.c.)`$ (18) where $`d`$ is the mass dimension of the nonrenormalizable operator in the potential, and $`A`$ is a typical trilinear coupling in the soft supersymmetry breaking part. For the purpose of this work the essential piece in this formula is $`A`$ dependent part. As discussed above, if $`A`$–terms are complex parameters, in general, the $`Re\phi `$ as well as $`Im\phi `$ can develop nonvanishing vacuum expectation values leading to CP violating Q–matter. ## 3 Discussions In this work we have discussed Q-ball formation in the MSSM with explicit CP violation. It is seen that the complex $`\mu `$ parameter and $`A`$–terms can affect the Q-ball formation process. In particular, it is shown that the scalar vacuum expectation value in the Q–ball depends additively on the soft CP phases whereas the mass per unit charge of the Q–ball depends only on the relative phases. There are regions of the parameter space where finite CP phases induce a stable Q–ball where the strictly CP–conserving does not. Once the Q–ball is formed, as long as the Q–matter is CP violating one, one expects that the scatterings of the light fermions from the Q–ball as well as its decay to fermions (neutralino LSP, for example) can affect their phenomenology. Particularly interesting is the Q–matter contributing to dark matter where the detection rates will change with the CP violating phases considerably <sup>?</sup>. It is worthy of reemhasizing that: 1. The vacuum expectation values of the fields, that is, the Q–ball matter $`\phi `$ depends on the individual relative CP phases additively; however, 2. The mass per unit charge in the Q-ball depends only on the relative phases. Hence, there are regions of the parameter space where the Q–matter is nonvanishing; however, the Q–ball solutions are not stable. Moreover, there are regions of the parameter space where finite CP phases suport a stable Q–ball whereas the CP conserving MSSM does not. These discussions equally apply to $`L`$–balls as well, and one may obtain new kinds of Q–ball solutions by including appropriate soft terms such as $`R`$–parity violating ones. Acknowledgments D. A. D. thanks Alexander Kusenko for useful conversations.
warning/0004/gr-qc0004002.html
ar5iv
text
# Gravitational Waves from the Dynamical Bar Instability in a Rapidly Rotating Star ## I Introduction A self–gravitating, axisymmetric fluid body with a sufficiently high rotation rate can be dynamically unstable to non-axisymmetric perturbations. Typically, the fastest growing unstable mode is the $`m=2`$ “bar” mode which acts to transform the body from a disk–like shape to an elongated bar that tumbles end–over–end. This instability has been described analytically for the case of uniform density bodies, and has been the subject of numerous numerical studies<sup>*</sup><sup>*</sup>*For a brief review, see Ref. . The numerical results show that bar formation is accompanied by the ejection of mass and angular momentum, and that the ejected matter forms long spiral arms in the equatorial plane. The subsequent evolution is less certain. Some simulations indicate that the bar shape is short lived, with the star returning to a predominantly disk–like shape after a few bar–rotation periods. Other simulations predict that the bar persists for many bar–rotation periods. In recent work, New, Centrella, and Tohline address this issue with a series of simulations using two different codes at various resolutions, and conclude that the bar shape is persistent. In their highest resolution run the bar decayed after roughly $`6`$ or $`7`$ bar–rotation periods. This was believed to be caused by numerical errors that induced an unphysical center of mass motion. In a lower resolution run a symmetry condition was imposed that prevented any center of mass motion. In that case the star maintained its bar shape throughout the simulation. The purpose of the present work is to simulate the long–time evolution of a rapidly rotating, self–gravitating star using Newtonian hydrodynamics and gravity. As discussed in Sec. III, the numerical code is substantially different from the codes that have been used previously to address this problem. The initial data for this study consists of a $`\gamma =5/3`$ polytrope with stability parameter $`\beta =0.30`$. The stability parameter is defined by $`\beta =T/|W|`$, where $`T`$ is rotational kinetic energy and $`W`$ is gravitational potential energy. The results here suggest that the bar shape is indeed long–lived—the star displays a prominent bar shape at the end of the simulation, which includes more than $`10`$ bar rotation periods. Numerical studies of fluids with various equations of state and initial rotation profiles have shown that the dynamical bar instability appears when the stability parameter $`\beta `$ exceeds a certain critical value, typically close to $`0.27`$. For the secular instability, which arises through dissipative mechanisms, the critical value of $`\beta `$ is near $`0.14`$. A neutron star might reach the critical value of $`\beta `$ for dynamical or secular instability by accreting matter and angular momentum from a binary companion. A stellar core that has exhausted its nuclear fuel might reach a critical rotation rate as it collapses. A star or stellar core that develops a rotating bar–like configuration will generate large amounts of gravitational radiation. Depending on the distance of the source, this radiation might be strong enough to be detected by the world–wide network of gravitational–wave detectors currently under construction. Here, the question of persistence of the bar shape becomes very important. The detectability of a source depends on its characteristic amplitude $`h_ch\sqrt{n}`$, where $`h`$ is the amplitude of the waves with frequency $`f`$ and $`n`$ is the number of wave cycles in a bandwidth near $`f`$. Thus, long–duration signals with large $`n`$ can be more easily detected than short–duration signals. Should we expect the bar shape to persist or decay? One reason why the bar might decay is the loss of mass and angular momentum from the ends of the bar. The accompanying drop in rotational kinetic energy could reduce the stability parameter and allow the star to return to axisymmetry. Loss of rotational kinetic energy through shock heating might also occur. For the simulation presented here $`\beta `$ has an initial value of $`0.30`$, large enough to dynamically trigger the growth of the bar mode. During the initial period of bar formation, mass and angular momentum are shed from the ends of the bar and long, spiral arms are formed. The stability parameter rapidly drops below $`0.27`$ and eventually settles to a value of about $`0.24`$. However, these losses of mass and angular momentum, with the accompanying drop in $`\beta `$, are insufficient to completely rob the star of its bar shape. The results here are consistent with the conjecture that the star will retain its bar shape indefinitely on a dynamical time scale, as long as $`\beta `$ stays above the critical value for secular instability. If this is correct, the bar should still decay due to secular processes. Imamura, Durisen, and Pickett have argued that the bar continues to shed small amounts of angular momentum to the spiral arms, and these loses cause the bar to decay. Other secular mechanisms, such as energy and angular momentum loss through gravitational radiation, could play a role as well. It is possible that, for the equation of state and angular velocity profile used here, the star is able to retain its bar shape simply because the critical value of $`\beta `$ for dynamical instability is $`0.24`$ or less. This possibility was tested by conducting a second simulation using the same equation of state and scaling the angular velocity by a factor of $`3/4`$. The resulting model has an initial value of $`\beta `$ equal to $`0.25`$. The bar mode showed no signs of growth in this simulation. Thus, for the models tested here, the bar mode does not spontaneously grow unless $`\beta `$ exceeds a critical value greater than $`0.25`$. However, once the bar shape is established, it can persist for many bar–rotation periods with $`\beta `$ equal to $`0.24`$ or less. In the present numerical calculation there are two sources of error in angular momentum. First, some of the angular momentum that is lost off the edge of the computational grid might, ideally, fall back onto the star. Second, numerical errors at each time step introduce a purely artificial loss of angular momentum. Both of these errors act to decrease the angular momentum of the star. One expects such losses to cause the bar shape to deteriorate. What is seen, both in the gravitational wave analysis and in the Fourier analysis of the density, is a relatively persistent signal with a modest decline in amplitude over the duration of the simulation. The physical model and initial data that form the basis of the simulation are described in Sec. II. Section III contains a discussion of the numerical code. The results of the simulation are described qualitatively and quantitatively in Sec. IV. Conclusions and results are summarized in Sec. V. Details of the initial data code are given in the Appendix. ## II Physical model The initial data for the simulation consists of a stationary, axisymmetric Newtonian fluid star with polytropic equation of state, $`P=K\rho ^\gamma `$. The equation of state parameters are chosen to coincide roughly with those of a low density neutron star with soft equation of state, $`\gamma =5/3`$ and $`K=5.38\times 10^9\mathrm{cgs}`$. The bar instability in stellar cores with stiff equations of state has been investigated by Houser and Centrella. The equations of hydrostatic equilibrium are solved using an algorithm described in the Appendix, in which the freely specifiable data are the central density $`\rho _\mathrm{c}`$ and the angular velocity distribution $`\omega (r)`$. Here, $`r`$ is the distance from the rotation axis. With the choices $`\rho _\mathrm{c}`$ $`=`$ $`2.00\times 10^{14}\mathrm{g}/\mathrm{cm}^3,`$ (2) $`\omega (r)`$ $`=`$ $`(4000/\mathrm{s})e^{(r/4.80\times 10^6\mathrm{cm})^2},`$ (3) the equilibrium configuration has mass, equatorial radius, polar radius, angular momentum, and stability parameter given by $`M`$ $`=`$ $`2.37M_{},`$ (5) $`R_{\mathrm{eq}}`$ $`=`$ $`4.91\times 10^6\mathrm{cm},`$ (6) $`R_\mathrm{p}`$ $`=`$ $`1.11\times 10^6\mathrm{cm},`$ (7) $`J`$ $`=`$ $`6.98\times 10^{49}\mathrm{g}\mathrm{cm}^2/\mathrm{s},`$ (8) $`\beta `$ $`=`$ $`0.300,`$ (9) respectively. Equations (1b) and (2b) show that the angular velocity has values $$\omega (0)=4000/\mathrm{s},\omega (R_{\mathrm{eq}})=1400/\mathrm{s},$$ (10) on the rotation axis and at the equator. The azimuthal velocity at the equator is $`6.89\times 10^9\mathrm{cm}/\mathrm{s}`$, below the Kepler velocity of $`8.90\times 10^9\mathrm{cm}/\mathrm{s}`$. Note that the mass (2a) is considerably larger than the masses of observed neutron stars, and exceeds the limit for nonrotating neutron stars for most equations of state. However, as discussed by Baumgarte, Shapiro, and Shibata, differentially rotating “hypermassive” neutron stars with $`M>2M_{}`$ might appear as the immediate products of core–collapse supernovae or binary neutron star mergers. The rotation law for a uniform density Maclaurin spheroid is $$j_{\mathrm{sp}}(m_{\mathrm{en}})=(5J/2M)[1(1m_{\mathrm{en}}/M)^{2/3}],$$ (11) where $`j_{\mathrm{sp}}=r^2\omega `$ is the specific angular momentum and $`m_{\mathrm{en}}`$ is the mass enclosed in a cylindrical radius. Many previous works on the dynamical bar instability have employed this same rotation law (see, for example, Refs. ). For the initial data code used here the angular velocity is specified as a function of radius $`r`$, so the rotation law (4) cannot be used directly. However, the rotation law (1b) was chosen for its similarity to the Maclaurin law, as Fig. 1 shows. In Fig. 1, the numerical data for $`j_{\mathrm{sp}}`$ and $`m_{\mathrm{en}}`$ are plotted for the rotation law (1b), along with the Maclaurin rotation law (4). For clarity, only one in five numerical data points is shown. The data and the rotation law (4) agree on the value of $`j_{\mathrm{sp}}`$ to within $`2\%`$ throughout most of the star, out to about $`m_{\mathrm{en}}/M=0.8`$. At the surface of the star, the rotation law (1b) differs from the Maclaurin rotation law by about $`8\%`$. The dimensionful model described here can be rescaled by introducing parameters $`\lambda `$, $`\mu `$, and $`\tau `$ that scale the lengths, masses, and times. To be precise, a quantity $`Q`$ with dimensions $`[Q]=L^AM^BT^C`$, where $`L`$ is length, $`M`$ is mass, and $`T`$ is time, is rescaled according to $`Q_{\mathrm{new}}=Q_{\mathrm{old}}/(\lambda ^A\mu ^B\tau ^C)`$. A physical, dimensionful model is retained if the parameters are dimensionless and Newton’s constant $`G`$ is unchanged by the rescaling. This requires $`\lambda `$, $`\mu `$, and $`\tau `$ to satisfy $`\lambda ^3=\mu \tau ^2`$. Alternatively, the model can be converted to polytropic units in which $`G=K=M=1`$ by setting $`\lambda =4.81\times 10^5\mathrm{cm}`$, $`\mu =4.71\times 10^{33}\mathrm{g}`$, and $`\tau =1.88\times 10^5\mathrm{s}`$. This leads to $`\rho _\mathrm{c}`$ $`=`$ $`4.73\times 10^3,`$ (13) $`R_{\mathrm{eq}}`$ $`=`$ $`10.2,`$ (14) $`\omega (0)`$ $`=`$ $`7.53\times 10^2,`$ (15) for the central density, equatorial radius, and central angular velocity. For the model with the scaling displayed in Eqs. (1) and (2), the maximum flow velocity is $`8.2\times 10^9\mathrm{cm}/\mathrm{s}`$. This occurs at a distance of $`3.4\times 10^6\mathrm{cm}`$ from the rotation axis. At this peak velocity, the special relativistic gamma factor $`(1v^2/c^2)^{1/2}`$, where $`c`$ is the speed of light, differs from unity by about 4%. General relativistic effects should be greatest near the surface of the star at the poles, where the gravitational potential is approximately $`GM/R_\mathrm{p}=2.8\times 10^{20}\mathrm{cm}^2/\mathrm{s}^2`$. Comparing this result with $`c^2`$, we find that general relativistic effects represent a correction of roughly 30% at the poles. Thus, with the scaling used in Eqs. (1) and (2), the errors that we introduce by ignoring special and general relativity are substantial. The errors are reduced if the model is rescaled appropriately. For example, with $`\lambda =1/2`$, $`\mu =2`$, and $`\tau =1/4`$, the mass of the star is cut in half, its linear size is increased by a factor of $`2`$, and its angular velocity is decreased by a factor of $`4`$. In this case the model might represent a stellar core that has partially collapsed, but is prevented from contracting to nuclear density by centrifugal forces. (Note that with this scaling Newton’s constant is unchanged. Also note that the speed of light should not be scaled, since it does not appear in the Newtonian hydrodynamical equations or the Poisson equation.) The scaling reduces the maximum speed by a factor of $`2`$ and reduces the gravitational potential by a factor of $`4`$. For this model, special and general relativistic effects represent corrections of about 2% and 8%, respectively. ## III Numerical code The axisymmetric initial data was calculated on a grid in the $`r`$$`z`$ plane consisting of $`512\times \mathrm{\hspace{0.17em}511}`$ zones. Details and tests of the initial data code are contained in the Appendix. In preparation for the evolution of this data, the grid was chosen to cover a physical domain of $`1.77\times 10^7\mathrm{cm}`$ in radius and $`2.50\times 10^7\mathrm{cm}`$ in the $`z`$–direction. Thus, the star was contained in a small region of the grid, roughly the inner $`142\times \mathrm{\hspace{0.17em}45}`$ zones. The quality of the data can be checked with the diagnostic expression $`𝒱=(2T+W+3P𝑑V)/W`$, where $`T`$ is kinetic energy, $`W`$ is gravitational potential energy, and $`P𝑑V`$ is the volume integral of pressure. According to the virial equations, this quantity should vanish. For the numerically generated model, this diagnostic has a value of $`𝒱=3.2\times 10^5`$. At the beginning of the evolution, the density and velocity are interpolated onto a three–dimensional Cartesian grid. The grid contains $`128^3`$ zones and covers a cubic domain with sides of length $`2.50\times 10^7\mathrm{cm}`$. The equatorial radius of the star spans $`25`$ zones, while the polar radius spans $`6`$ zones. The density in each zone is modified with a random perturbation ranging from $`10\%`$ to $`+10\%`$ of the unperturbed value. Initially, the specific internal energy $`e`$ is obtained from the density by the relation $`e=K\rho ^{\gamma 1}/(\gamma 1)`$. Thereafter, the star is evolved with the equation of state $$P=(\gamma 1)\rho e$$ (16) with $`\gamma =5/3`$. The evolution code includes the nonrelativistic hydrodynamics code VH-1, written by Blondin, Hawley, Lindahl and Lufkin. VH-1 is based on the piecewise–parabolic method (PPM) as described by Colella and Woodward. The PPM scheme is a higher–order extension of Godunov’s method. It uses parabolas as interpolation functions within each zone, and characteristics to determine the domains of dependence for zone interfaces. The average values of density, pressure and velocity within these domains are used as inputs to the Riemann problem between adjacent zones. The solution of the Riemann problem determines time averaged values of pressure and velocity, which in turn are used to compute hydrodynamical fluxes. In VH-1, each time step is reduced to three one–dimensional evolutionary “sweeps” via operator splitting. For each one–dimensional sweep, the fluid variables are evolved in Lagrangian coordinates and then remapped onto the original Eulerian grid. The order of the sweeps is cycled through all possible permutations. For the simulation presented here, I used a “Courant number” of $`0.3`$. That is, the timestep was set to $`0.3`$ times the maximum allowed by the CFL condition. This is a relatively small timestep for a PPM code. It was chosen to help minimize the artificial loss of angular momentum, discussed below. The Poisson equation for the gravitational potential is solved using multigrid methods. The finest grid has size $`129^3`$, with grid points that lie at the corners of the $`128^3`$ hydrodynamical zones. The boundary conditions are computed from a multipole expansion of the potential that includes monopole, dipole, quadrupole, and octupole terms. The PPM method is designed to evolve the fluid mass density $`\rho `$, linear momentum density, and total energy density, and in the process to conserve the mass, linear momentum, and total energy of the system. Unfortunately, for such a rapidly rotating star, the total energy density $`E`$ is dominated by kinetic energy density $`K`$. This is a problem because the internal energy density $`\rho e=EK`$ is needed for the calculation of pressure from the equation of state (6). Since $`\rho e`$ is a difference of large numbers, it is subject to large numerical errors. As long as the fluid flow is smooth, one can solve this problem by modifying the code so that internal energy, rather than total energy, is evolved. However, this modification of the PPM algorithm leads to the wrong jump conditions at shock fronts. Thus, for the present simulation, a compromise was struck: in regions of smooth flow the code evolves internal energy, and in the neighborhood of shocks the code evolves total energy. Similar schemes have been described in Refs. . VH-1 uses the shock flattening algorithm described by Colella and Woodward as a dissipation mechanism. The algorithm determines a “flattening coefficient” $`f`$ for each computational zone, which ranges from $`0`$ for smooth flow to a maximum of $`0.5`$ in the presence of a strong shock. The flattening coefficient is also used to determine whether internal energy or total energy is used to update the fluid variables in each zone. Thus, internal energy is used when $`f`$ is less than some threshold value, $`f_t`$, while total energy is used for $`f>f_t`$. In practice, the performance of the code was found to be insensitive to the value of the threshold $`f_t`$, for values ranging from $`0.1`$ to $`0.5`$. The ability of this “hybrid” code to switch from internal energy to total energy at shock fronts was demonstrated on a variety of one–dimensional test problems. These include the standard Sod shock tube problem with density and pressure ratios of up to $`1.0\times 10^6`$, and strong standing shocks. Figures 2 and 3 show the results of the most extreme test, a standing shock wave with an upstream Mach number of $`1.0\times 10^8`$. Results are shown for the hybrid code, a total energy code (which evolves total energy exclusively) and an internal energy code (which evolves internal energy exclusively). The simulations use 100 zones to cover the computational domain, $`0x1`$, and a Courant number of $`0.6`$. The upstram fluid, in the region $`0.5<x1`$, has density $`\rho =1.0`$, pressure $`P=1.0`$, and velocity $`u=1.3\times 10^8`$. The downstream fluid, in the region $`0x<0.5`$, has density $`\rho =4`$, pressure $`P=1.25\times 10^{16}`$, and velocity $`u=3.2\times 10^7`$. Ideally the discontinuities should be maintained at $`x=0.5`$. Figures 2 and 3 show the pressure and density after $`250`$ timesteps. The CFL condition for this simulation is determined by the upstream velocity, so the upstream fluid travels a distance of $`0.006`$ in each timestep. Figure 2 shows that the total energy code produces extremely large errors in the upstream pressure. As expected, Figs. 2 and 3 show that the results from the internal energy code are quite poor. Altogether the hybrid code is the most successful at maintaining the proper pressure and density profiles. For this extreme test, the largest errors from the hybrid code appear in the downstream fluid where the density and pressure differ from their ideal values by at most $`11\%`$ and $`1.6\%`$, respectively. The hybrid code also lost about $`2.7\%`$ of the total energy during the $`250`$–timestep simulation. Because the present code conserves linear momentum, there is no erroneous center of mass motion due to numerical errors. On the other hand, the code does not conserve angular momentum. This is a fairly serious problem because the angular motion of the star is the central effect we wish to study. The amount of angular momentum that is lost, or gained, through numerical error is found by computing the total angular momentum on the grid, $`J_{\mathrm{grid}}`$, and correcting for the amount of angular momentum that is lost (or gained) through the grid boundaries, $`J_{\mathrm{lost}}`$. The angular momentum passing through the grid boundaries is obtained by computing the angular momentum that moves off the Eulerian grid during each Lagrangian time step. Ideally, the sum of the angular momentum on the grid and the angular momentum lost through the boundaries, $`J_{\mathrm{grid}}+J_{\mathrm{lost}}`$, should be constant. As seen in the next section, over the course of the simulation ($`8000`$ time steps) about 8% of the star’s angular momentum is lost through the boundaries. During this same time, the total angular momentum drops by nearly 25%. Thus, numerical errors account for a drop of about 17% in the angular momentum of the star. In the Newtonian/quadrupole approximation, the gravitational wave signal $`h_{ij}`$ is formed from linear combinations of components of the second time derivative of the reduced quadrupole moment, $`\ddot{I\text{-}}_{ij}`$. The relationship is $`h_{ij}=(2G/Rc^4)P_{ij}^k\mathrm{}\ddot{I\text{-}}_k\mathrm{}`$, where $`P_{ij}^k\mathrm{}`$ is the projection operator for transverse directions and $`R`$ is the distance from the source to the observation point. Using the hydrodynamical equations, the time derivatives that appear in $`\ddot{I\text{-}}_{ij}`$ can be eliminated so that, to within boundary terms, we have $$\ddot{I\text{-}}_{ij}=\mathrm{STF}d^3x\mathrm{\hspace{0.17em}2}\rho \left[v_iv_jx_i_j\mathrm{\Phi }\right].$$ (17) Here, “STF” stands for the symmetric, trace free part of the expression that follows, $`v_i`$ is the fluid velocity, $`_j=/x_j`$ is a spatial derivative, and $`\mathrm{\Phi }`$ is the gravitational potential. The gravitational–wave signal is computed numerically by using expression (7) for $`\ddot{I\text{-}}_{ij}`$. This calculation is subject to errors due to the finite size of the computational grid. Recall that during the course of the simulation, matter is expelled from the star. Some of this matter, a total of $`0.048M_{}`$, passes through the grid boundaries. As a consequence, the relationship $`h_{ij}=(2G/Rc^4)P_{ij}^k\mathrm{}\ddot{I\text{-}}_k\mathrm{}`$ is not correct, even in the Newtonian/quadrupole approximation, because the lost matter is not included in the calculation of $`\ddot{I\text{-}}_{ij}`$. A second source of error stems from the fact that boundary terms were discarded in the derivation of Eq. (7). These boundary terms would vanish if the density were always zero on the grid boundary. Fortunately, the matter that reaches the boundary has relatively low density and the errors that arise from the missing boundary terms in Eq. (7) are small. This has been verified by comparing the results from Eq. (7) with the results obtained by computing numerically the second time derivative of the reduced quadrupole moment. The numerical derivatives were obtained by performing a least–squares fit to a quadratic using $`5`$ or $`7`$ consecutive data points. The second derivative at the central time is found from the curvature of the quadratic. This calculation is subject to a certain amount of numerical noise, but otherwise the results agree quite closely with those obtained from Eq. (7). Since the matter that passes through the grid boundaries does not greatly effect the calculation of $`\ddot{I\text{-}}_{ij}`$ with Eq. (7), it is plausible to assume that the errors obtained by equating $`h_{ij}`$ and $`(2G/Rc^4)P_{ij}^k\mathrm{}\ddot{I\text{-}}_k\mathrm{}`$ are also small. The frequency spectrum for the gravitational–wave signal is computed with the Lomb normalized periodogram. This method was chosen for its ability to handle the unevenly sampled data. In the Newtonian/quadrupole approximation, the gravitational–wave luminosity is given by $`L=G\ddot{I\text{-}}\text{.}{}_{ij}{}^{}\ddot{I\text{-}}\text{.}{}_{}{}^{ij}/(5c^5)`$, and the radiated energy $`\mathrm{\Delta }E`$ is the time integral of $`L`$. The calculation of the time derivative $`\ddot{I\text{-}}\text{.}{}_{ij}{}^{}=d\ddot{I\text{-}}_{ij}/dt`$ must be handled with care since, otherwise, even small errors in $`L`$ will accumulate in the integration over time. For this calculation, I use a Savitzky–Golay approach. Specifically, the time derivative of the function $`\ddot{I\text{-}}_{ij}`$ in each time interval is computed from the slope of a quadratic polynomial that is obtained from a least–squares fit to several nearest neighbor data points. The number of points used is typically between $`10`$ and $`20`$. The evolution code was run on a Cray T90 vector computer, while the initial data, gravitational wave spectrum, and gravitational wave luminosity were computed on local workstations. The evolution required $`8000`$ timesteps and about $`45`$ CPU hours. ## IV Description of the evolution The star retains its predominantly circular, disk–like shape for about $`10\mathrm{ms}`$, which amounts to approximately $`6`$ rotation periods for the fluid near the center of the star. Over the next few milliseconds the bar mode grows rapidly, so that by $`13\mathrm{ms}`$ the star is highly elongated as shown in Fig 4. After a few more milliseconds the central regions of the star recircularize, and by $`16\mathrm{ms}`$ the inner density contours have nearly lost their bar shape as seen in Fig. 5. The cycle of bar formation and recircularization repeats; by $`19\mathrm{ms}`$ the star again shows a strong bar–like shape, as shown in Fig. 6. The episodes of bar formation and recircularization continue with diminishing amplitude and a period of about $`6.5\mathrm{ms}`$. During each episode, the bar undergoes a little more than $`1\frac{1}{2}`$ revolutions. Near the end of the simulation, at a time of $`50\mathrm{ms}`$, the oscillations become quite weak and the star settles into a bar–shaped configuration of modest strength. See Fig. 7. During the initial episode of rapid bar formation, matter is thrown outward from the ends of the bar. The ejected matter forms long spiral arms, clearly visible in Figs. 4 and 5. Some of this matter is lost off the edge of the computational grid—between $`14`$ and $`20\mathrm{ms}`$ the mass loss totals $`0.040M_{}`$. The subsequent episodes of bar formation are less violent, with only small amounts of ejected mass reaching the grid boundaries. The mass loss between $`20\mathrm{ms}`$ and the end of the simulation amounts to $`0.008M_{}`$. When the star assumes its bar shape, the individual particles in the inner regions of the star circulate along trajectories that roughly coincide with the constant density contours. For example, at the time shown in Fig. 4, $`13.58\mathrm{ms}`$, the bar is rotating about the center of mass with a period of approximately $`4\mathrm{ms}`$. Individual particles flow along the $`10^{14}\mathrm{g}/\mathrm{cm}^3`$ density contour with fairly uniform speeds ranging from about $`6.5\times 10^9\mathrm{cm}/\mathrm{s}`$ to about $`7.5\times 10^9\mathrm{cm}/\mathrm{s}`$. The density contour at $`10^{14}\mathrm{g}/\mathrm{cm}^3`$ has a perimeter length of approximately $`1.7\times 10^7\mathrm{cm}`$. Thus, the particles orbit along this contour with period $`2.5\mathrm{ms}`$ as the contour precesses about the star’s center with period $`4\mathrm{ms}`$. The shape of the star can be quantified by Fourier analyzing the density in a circle in the equatorial plane. The Fourier coefficients are defined by $$A_m+iB_m=\frac{1}{2\pi }_0^{2\pi }𝑑\phi \rho (\phi )e^{im\phi },$$ (18) where $`\rho (\phi )`$ is the density in the equatorial plane at an arbitrarily chosen distance of $`2.0\times 10^6\mathrm{cm}`$ from the center of the star. Here, $`\phi `$ is the azimuthal angle. The amplitude of the $`m^{\mathrm{th}}`$ Fourier mode is defined by $`C_m=\sqrt{A_m^2+B_m^2}`$, and the phase angle is $`\varphi _m=\mathrm{arctan}(B_m/A_m)`$. Figure 8 shows the natural logarithm of the ratio of the amplitude $`C_m`$ and the average density $`C_0`$ for the $`m=2`$ bar mode and the $`m=4`$ mode. The $`m=3`$ mode is small and is not displayed in Fig. 8. The coefficient $`C_3/C_0`$ remains less than $`0.01`$ through most of the evolution, reaching peak values of $`0.02`$ in the late stages. The periodic growth and decay of the bar shape is clearly seen in the peaks and valleys of the graph in Fig. 8. The amplitudes grow exponentially for several milliseconds during the initial episode of bar formation. For the $`m=2`$ bar mode, the growth rate near $`10\mathrm{ms}`$ is $`d\mathrm{ln}C_2/dt820/\mathrm{s}`$. For the $`m=4`$ mode, the growth rate is $`d\mathrm{ln}C_2/dt1400/\mathrm{s}`$. In polytropic units ($`\tau =1.88\times 10^5\mathrm{s}`$), the growth rates for the $`m=2`$ and $`m=4`$ modes are $`0.015`$ and $`0.026`$, respectively. For $`m=2`$, this value differs by a few percent from the result $`0.0145`$ reported in Ref.. In units of the “dynamical time” $`t_D=[R_{\mathrm{eq}}^3/(GM)]^{1/2}`$, where $`R_{\mathrm{eq}}`$ is the initial equatorial radius, the growth rates are $`0.50/t_D`$ and $`0.86/t_D`$ for the $`m=2`$ and $`m=4`$ modes. Comparing these results to the values $`0.55/t_D`$ and $`1.1/t_D`$ obtained from the highest resolution run in Ref. , we find a fairly large discrepancy in the $`m=4`$ case. This difference might be caused by numerical errors in the present code associated with the cartesian grid. Near the end of the simulation, the oscillations in the Fourier amplitudes have subsided and the star settles into a bar shape with strength $`C_2/C_00.18`$. The downward drift in $`C_2/C_0`$, visible in Fig. 8, is possibly due to numerical error as discussed below. The phase angles for the $`m=2`$ and $`m=4`$ modes are shown in Fig. 9. The eigenfrequencies are $`d\varphi _2/dt=3.24/\mathrm{ms}`$ and $`d\varphi _4/dt=6.52/\mathrm{ms}`$. The pattern speeds $`m^1d\varphi _m/dt`$ are $`1.62/\mathrm{ms}`$ for the $`m=2`$ mode and $`1.63/\mathrm{ms}`$ for the $`m=4`$ mode. These speeds, which equal $`0.99/t_D`$ and $`1.0/t_D`$ in dynamical time units, agree with the results obtained in Refs. to within $`1\%`$. The approximate equality of the pattern speeds shows that the $`m=4`$ mode is a harmonic of the $`m=2`$ bar mode. The stability parameter $`\beta `$ undergoes fluctuations, dropping to a local minimum when the bar amplitude $`C_2/C_0`$ is a local maximum, and rising to a local maximum when the bar amplitude is a local minimum. This behavior is shown in Fig. 10. As the fluctuations diminish, $`\beta `$ settles to a value of $`0.24`$ with a slight downward drift. Note that this value of $`\beta `$ is well below the critical value $`0.27`$ for the growth of the bar mode in a constant density star. The fact that the bar shape persists with $`\beta `$ below $`0.27`$ might indicate that the threshold for bar formation in a $`\gamma =5/3`$ star with rotation law given by Eq. (1b) is actually $`0.24`$ or less. This possibility was tested by evolving a second model star, obtained by modifying the freely specifiable data in Eq. (1) so that the central value of angular velocity is $`3000/\mathrm{s}`$ rather than $`4000/\mathrm{s}`$. The parameters describing this model are $`M`$ $`=`$ $`1.34M_{},`$ (20) $`R_{\mathrm{eq}}`$ $`=`$ $`5.88\times 10^6\mathrm{cm},`$ (21) $`R_\mathrm{p}`$ $`=`$ $`1.40\times 10^6\mathrm{cm},`$ (22) $`J`$ $`=`$ $`2.90\times 10^{49}\mathrm{g}\mathrm{cm}^2/\mathrm{s},`$ (23) $`\beta `$ $`=`$ $`0.253,`$ (24) so, in particular, the stability parameter is between $`0.24`$ and $`0.27`$. This model was evolved for approximately $`23\mathrm{ms}`$, which equals about $`11`$ rotation periods for the fluid near the center of the star. During this time the $`m=2`$ bar mode showed no signs of growth. We are thus led to conclude that for a $`\gamma =5/3`$ fluid star with rotation law of the type considered here, the dynamical $`m=2`$ bar mode will not spontaneously grow unless $`\beta `$ exceeds a critical value near $`0.27`$; however, once a bar shape is established, it can persist for many rotation periods with $`\beta =0.24`$ or less. As discussed in the previous section, the numerical code does not conserve angular momentum. In Fig. 11, the total angular momentum on the grid, $`J_{\mathrm{grid}}`$, is plotted as a function of time along with the difference between the initial angular momentum $`J_0=6.98\times 10^{49}\mathrm{g}\mathrm{cm}^2/\mathrm{s}`$ and the angular momentum $`J_{\mathrm{lost}}`$ that flows off the edge of the numerical grid. Ideally, these two curves should coincide; the fact that $`J_{\mathrm{grid}}`$ falls below $`J_0J_{\mathrm{lost}}`$ indicates that numerical errors artificially remove angular momentum from the system. The results for $`J_{\mathrm{lost}}`$ were checked by computing the angular momentum flux from each of the saved data sets, which were generated every 50 timesteps. The numerically computed time derivative of $`J_{\mathrm{lost}}`$ is in very close agreement with the angular momentum flux. I have not yet determined the source of the numerical errors in angular momentum. Note, however, that the errors are not significant until the star begins to develop a noticeable bar shape, at about $`10\mathrm{ms}`$. This suggests that the code has difficulty tracking the leading and trailing edges of the bar–shaped star, with their sharp density gradients, as the bar rotates in the $`x`$$`y`$ plane. Since numerical errors cause angular momentum to be lost, it is plausible to expect that rotational kinetic energy is artificially lost as well. This might account for the downward drift in the stability parameter at late times, as seen in Fig. 10. In turn, one might expect a drop in the stability parameter to cause the star’s bar shape to decay. Figure 8 shows that the Fourier amplitude $`C_2`$ is fairly robust at late times with a slight downward trend on average. If angular momentum were not artificially lost, the simulation might show that after the fluctuations cease the bar shape remains unchanged on a dynamical time scale. ## V Gravitational waves In the Newtonian/quadrupole approximation, the gravitational wave amplitude for the plus polarization state, $`h_+`$, as measured at a distance $`R`$ in the equatorial plane of the source, is $`{\displaystyle \frac{c^4R}{G}}h_+(\mathrm{eq})`$ $`=`$ $`(\mathrm{cos}^2\phi 2)\ddot{I\text{-}}_{xx}+(\mathrm{sin}^2\phi 2)\ddot{I\text{-}}_{yy}`$ (26) $`+2\mathrm{cos}\phi \mathrm{sin}\phi \ddot{I\text{-}}_{xy}.`$ Here, $`\ddot{I\text{-}}_{ij}`$ is the second time derivative of the reduced quadrupole moment and $`\phi `$ is the azimuthal angle of the observation point relative to the (arbitrary) $`x`$–axis of the source. Figure 12 shows the results for $`h_+`$ with $`R=20\mathrm{Mpc}`$ and $`\phi =\pi /2`$. (Recent estimates indicate that the distance to the Virgo cluster is $`20.9\mathrm{Mpc}`$.) In the equatorial plane, the cross polarization amplitude, $`h_\times (\mathrm{eq})`$, is a linear combination of the components $`\ddot{I\text{-}}_{xz}`$ and $`\ddot{I\text{-}}_{yz}`$. Since the star is approximately symmetric under reflections in the equatorial plane, $`h_\times (\mathrm{eq})`$ as measured in the equatorial plane is nearly zero. For an observation point above the north pole of the star, the plus polarization amplitude is $`{\displaystyle \frac{c^4R}{G}}h_+(\mathrm{p})`$ $`=`$ $`(2\mathrm{cos}^2\phi 1)\ddot{I\text{-}}_{xx}+(2\mathrm{sin}^2\phi 1)\ddot{I\text{-}}_{yy}`$ (28) $`+4\mathrm{cos}\phi \mathrm{sin}\phi \ddot{I\text{-}}_{xy}.`$ Since the star is highly flattened and has little motion in the $`z`$–direction, the $`zz`$ component of the second time derivative of the unreduced quadrupole moment, $`\ddot{I}_{zz}`$, is very small. Furthermore, because the shape of the bar does not evolve rapidly compared to its rotation rate, the $`xx`$ and $`yy`$ components satisfy $`\ddot{I}_{xx}+\ddot{I}_{yy}0`$. These observations imply $`\ddot{I\text{-}}_{xx}\ddot{I\text{-}}_{yy}`$. As a consequence, we see from Eqs. (10) and (11) that $`h_+(\mathrm{p})2h_+(\mathrm{eq})`$. This is confirmed by the results of the simulation. As expected, the amplitude $`h_\times (\mathrm{p})`$ is almost identical to $`h_+(\mathrm{p})`$, just phase shifted by $`45^{}`$. Although the reliability of the results for the gravitational–wave signal is rather poor, due both to the crudeness of the physical model and to numerical errors, it is still of some interest to consider the detectability of these waves. Thus, consider the characteristic amplitude of the source, approximated by $$h_c\left(\frac{3G\mathrm{\Delta }E}{2\pi ^2c^3f_cR^2}\right)^{1/2}.$$ (29) Here, $`\mathrm{\Delta }E`$ is the total energy radiated in gravitational waves and $`f_c`$ is the characteristic frequency. The gravitational–wave frequency is sharply peaked about $`490\mathrm{Hz}`$, as shown in Fig. 13. The total energy radiated in gravitational waves, as computed in the Newtonian/quadrupole approximation up to the time the simulation was halted, is $`\mathrm{\Delta }E5\times 10^4\mathrm{M}_{}\mathrm{c}^2`$. From these data we find the characteristic amplitude of the source to be $`h_c4\times 10^{22}`$ at a distance of $`20\mathrm{Mpc}`$. This value for $`h_c`$ is a lower bound, since the radiated energy $`\mathrm{\Delta }E`$ will continue to increase as long as the bar persists. Also note that the result $`h_ch\sqrt{n}4\times 10^{22}`$ coincides roughly with $`n25`$ wave cycles with amplitude $`h0.08\times 10^{21}`$. This is consistent with the signal displayed in Fig. 12. The signal $`h_c4\times 10^{22}`$ is probably too weak to be seen by the first generation of interferometric gravitational–wave detectors, which should have a sensitivity of about $`10^{20}`$ at $`500\mathrm{Hz}`$. The signal might be detectable by the planned advanced detectors, which are expected to have a sensitivity below $`10^{21}`$ at $`500\mathrm{Hz}`$. For a source at a distance of $`25\mathrm{kpc}`$, comparable to the diameter of the Milky Way galaxy, the characteristic amplitude is $`h_c3\times 10^{19}`$. This is a relatively large signal that should be detected easily by the first generation of interferometers. If the simulation is scaled as described in section II, the amplitude as measured at a fixed (unscaled) distance $`R`$ changes by a factor of $`\tau ^2/(\mu \lambda ^2)`$ and the frequency changes by a factor of $`\tau `$. For example, with $`\lambda =1/2`$, $`\mu =2`$, and $`\tau =1/4`$, the amplitude $`h_+(\mathrm{eq})`$ at $`R=20\mathrm{Mpc}`$ is reduced from that shown in Fig. 12 by a factor of $`8`$. The characteristic amplitude $`h_c`$ is also reduced by a factor of $`8`$, to about $`5\times 10^{23}`$ at $`20\mathrm{Mpc}`$, and the peak frequency occurs at $`120\mathrm{Hz}`$. This signal is somewhat below the expected sensitivity of the advanced gravitational–wave detectors, which is approximately $`10^{22}`$ at $`120\mathrm{Hz}`$. At a distance of $`25\mathrm{kpc}`$, the characteristic amplitude is almost $`4\times 10^{20}`$, well above the expected sensitivity of $`4\times 10^{21}`$ for the first generation detectors. ## VI Conclusions The simulation presented here shows that a rapidly rotating $`\gamma =5/3`$ fluid star with rotation law Eq. (1b) is dynamically unstable to bar formation, and that the bar shape, once established, is relatively long lived. Over the final $`10`$ or $`20\mathrm{ms}`$ of the simulation, the star maintains a modest but fairly persistent bar shape with stability parameter $`\beta `$ equal to about $`0.24`$. It is possible that, for $`\gamma =5/3`$ stars with certain initial angular velocity profiles, the threshold for bar formation is actually $`0.24`$ or less. However, further numerical study showed that an axisymmetric star with $`\beta =0.25`$ initially is stable against growth of the bar mode. Thus, for the models considered here, the dynamical bar mode requires a value of $`\beta `$ near $`0.27`$ to grow but the bar shape can persist for many bar–rotation periods with $`\beta =0.24`$ or less. Imamura, Durisen, and Pickett suggest that as long as $`\beta `$ is greater than the critical value for growth of the secular instability, about $`0.14`$, a bar–shaped star is dynamically stable against forming an axisymmetric disk. The results here are consistent with this conjecture. Recent numerical studies by New, Centrella, and Tohline also indicate that the bar shape is long lived. The strength of the gravitational–wave signal emitted by an initially axisymmetric, dynamically unstable star depends on the rotating star’s length and time scales. With the numerical model scaled to represent a low–density neutron star with soft equation of state, the characteristic amplitude is $`h_c4\times 10^{22}`$ at a distance of $`20\mathrm{Mpc}`$. With the model scaled to represent a stellar core that has partially collapsed and is prevented from contracting to nuclear densities by centrifugal forces, the characteristic amplitude is $`h_c5\times 10^{23}`$ at $`20\mathrm{Mpc}`$. These signals are probably too weak to be detected by the first generation of interferometers, although the neutron star signal might be detectable by the planned advanced interferometers. At a distance of $`25\mathrm{kpc}`$, both signals are quite large and should be detected easily by the first generation interferometers. ## Acknowledgments I would like to thank John Blondin for helpful discussions. Numerical calculations were carried out on the Cray T90 at the North Carolina Supercomputing Center. ## Appendix The technique used in this work to solve the equations of hydrostatic equilibrium is similar to the self–consistent field method, first developed by Ostriker and Mark and later refined by Hachisu. The self–consistent field method is based on the integral form of the Euler equation which reads, for the equation of state $`P=K\rho ^\gamma `$, $$\frac{\gamma K}{\gamma 1}\rho ^{\gamma 1}+\mathrm{\Psi }+\mathrm{\Phi }=C.$$ (30) In this equation, $`\mathrm{\Psi }=𝑑rr\omega ^2`$ is the “rotational potential” derived from the angular velocity $`\omega (r)`$, $`\mathrm{\Phi }`$ is the gravitational potential, and $`C`$ is an integration constant. In the self–consistent field method, one begins with an initial guess for the density $`\rho `$, solves the Poisson equation for $`\mathrm{\Phi }`$, then uses Eq. (13) to obtain a corrected density distribution. The process is then iterated. Typically, the Poisson equation is solved by expanding the potential $`\mathrm{\Phi }`$ in terms of Legendre polynomials, and the constant $`C`$ appearing in Eq. (13) is determined by specifying certain properties at the star’s surface. For example, Hachisu fixes the “axis ratio” $`R_\mathrm{p}/R_{\mathrm{eq}}`$. For the method used in this paper, the constant $`C`$ is determined by properties at the center of the star. The rotational potential is chosen to vanish at the star’s center, $`\mathrm{\Psi }(0)=0`$, and the constant $`C`$ is written as $$C=\gamma K\rho _\mathrm{c}^{\gamma 1}+\mathrm{\Phi }_\mathrm{c}.$$ (31) Here, the subscript c denotes the center of the star. The central density $`\rho _\mathrm{c}`$ is specified as input data, along with the rotation law $`\omega (r)`$. The initial guess for $`\rho `$ is chosen to be a static, spherically symmetric polytrope with central density $`\rho _\mathrm{c}`$, obtained by solving numerically the Lane–Emden equation. The potential $`\mathrm{\Phi }`$, and its central value, are obtained from the Poisson equation using a multigrid algorithm, as discussed below. Note that with this scheme the center of the star must have nonzero density. This precludes the possibility of generating stars with toroidal surfaces. Equation (13) can be written as $`F=0`$ where the function $`F`$ is defined by $$F\frac{\gamma K}{\gamma 1}\rho ^{\gamma 1}+\mathrm{\Psi }+\mathrm{\Phi }C$$ (32) and the constant $`C`$ is given by Eq. (14). The corrections to the density are obtained from a Newton–Raphson algorithm applied to $`F=0`$. Note that $`F(x)`$ is a nonlocal function of $`\rho (x)`$, where $`x`$ labels points in space. The nonlocal dependence of $`F(x)`$ on $`\rho (x)`$ enters through the gravitational potential $`\mathrm{\Phi }(x)`$. One might expect that the value of $`\mathrm{\Phi }`$ at the point $`x`$ is insensitive to the density $`\rho (x)`$ at the same point $`x`$, since the potential $`\mathrm{\Phi }`$ is determined primarily by the global properties of the star, such as its mass. With this observation in mind, we compute the variation of $`F`$ at point $`x`$ with respect to $`\rho `$ at point $`x`$ by dropping the nonlocal terms. This leads to the approximate result $`\delta F\gamma K\rho ^{\gamma 2}\delta \rho `$. Then setting $`F+\delta F=0`$, we find the density correction $$\delta \rho =\frac{1}{\gamma K}\rho ^{2\gamma }(C\mathrm{\Phi }\mathrm{\Psi })\frac{1}{\gamma 1}\rho ,$$ (33) to be used for each Newton–Raphson iteration. I have not constructed a mathematical argument to justify the above assumption, namely, that the nonlocal terms in $`F`$ can be ignored in computing $`\delta F`$. However, this same assumption is implicit in the usual self–consistent field method. As described in Refs. , for the self–consistent field method the density is updated by solving Eq. (13) for $`\rho `$, ignoring the $`\rho `$–dependence in $`\mathrm{\Phi }`$. That is, the change in $`\rho `$ is $$\delta \rho =\left[\frac{\gamma 1}{\gamma K}(C\mathrm{\Phi }\mathrm{\Psi })\right]^{\frac{1}{\gamma 1}}\rho ,$$ (34) where $`\mathrm{\Phi }`$ is the solution of the Poisson equation. Formula (17) can be justified by the same reasoning that led to Eq. (16), but with the following difference: enthalpy $`H\gamma K\rho ^{\gamma 1}/(\gamma 1)`$, rather than density $`\rho `$, is treated as the independent variable. The variation of $`F`$ with respect to $`H`$, ignoring the $`\mathrm{\Phi }`$ dependence on $`H`$, is $`\delta F\delta H`$. Setting $`F+\delta F=0`$ leads to a new enthalpy distribution, $`H_{\mathrm{new}}=HF`$. Solving this equation for the new density $`\rho _{\mathrm{new}}`$, we find the density correction $`\delta \rho =\rho _{\mathrm{new}}\rho `$ of Eq. (17). The initial data code solves the equations of hydrostatic equilibrium at the zone centers of a $`513\times 513`$ grid in the $`r`$$`z`$ plane of a cylindrical coordinate system. The symmetry axis concides with the edge of the first column of zones. The density $`\rho `$ is computed on the inner $`512\times 511`$ zones. One layer of zones at the top, bottom, and side opposite the symmetry axis are reserved for setting boundary conditions for the gravitational potential $`\mathrm{\Phi }`$. The boundary values are obtained by expanding the potential in a multipole series that includes monopole and quadrupole terms. (The dipole and octupole moments vanish due to rotational symmetry and reflection symmery about the equatorial plane.) The Poisson equation is solved with a multigrid algorithm, adapted from the code described in Ref. to handle the boundary conditions and the cylindrical geometry. The initial data described in Sec. II was generated by a code that relies on the Newton–Raphson iteration scheme (16) and the multigrid Poisson solver described above. The results of three tests of this code (JDB) are shown in Table I. For comparison, the results obtained by Hachisu are also listed. The values reported in the table are scaled as in Hachisu’s self–consistent field method, with Newton’s constant $`G`$, the equatorial radius $`R_{\mathrm{eq}}`$, and the maximum density $`\rho _{\mathrm{max}}`$ equal to unity. Hats denote these scaled quantities. Also, $`\mathrm{\Pi }`$ denotes the volume integral of pressure, $`P𝑑V`$. | | test | $`\widehat{R}_\mathrm{p}`$ | $`\widehat{M}`$ | $`\widehat{J}`$ | $`\widehat{T}`$ | $`\widehat{W}`$ | $`3\widehat{\mathrm{\Pi }}`$ | $`\widehat{P}_{\mathrm{max}}`$ | $`𝒱`$ | | --- | --- | --- | --- | --- | --- | --- | --- | --- | --- | | JDB | $`r`$ | $`0.6667`$ | $`0.3288`$ | $`0.02575`$ | $`0.006641`$ | $`0.1164`$ | $`0.1031`$ | $`0.2044`$ | $`1.8\times 10^4`$ | | Hachisu | $`r`$ | $`0.667`$ | $`0.328`$ | $`0.0257`$ | $`0.00663`$ | $`0.116`$ | $`0.103`$ | $`0.204`$ | | | JDB | $`v`$ | $`0.3332`$ | $`0.6413`$ | $`0.1378`$ | $`0.06392`$ | $`0.3733`$ | $`0.2454`$ | $`0.2020`$ | $`4.7\times 10^5`$ | | Hachisu | $`v`$ | $`0.333`$ | $`0.639`$ | $`0.137`$ | $`0.0638`$ | $`0.372`$ | $`0.244`$ | $`0.202`$ | | | JDB | $`j`$ | $`0.1662`$ | $`0.8419`$ | $`0.1036`$ | $`0.04559`$ | $`0.5982`$ | $`0.5070`$ | $`0.3272`$ | $`1.6\times 10^5`$ | | Hachisu | $`j`$ | $`0.167`$ | $`0.843`$ | $`0.104`$ | $`0.0456`$ | $`0.599`$ | $`0.508`$ | $`0.327`$ | | The “$`r`$” test assumes a rigid rotation law, $`\omega (r)=\omega _0`$, with the scaled angular velocity given by $`\widehat{\omega }_0^2=0.266`$. The “$`v`$” test assumes the “$`v`$–constant” rotation law, $`\omega (r)=v_0/\sqrt{d^2+r^2}`$, with $`\widehat{v}_0^2=0.215`$ and $`\widehat{d}=0.100`$. The “$`j`$” test assumes the “$`j`$–constant” rotation law, $`\omega (r)=j_0/(d^2+r^2)`$, with $`\widehat{j}_0^2=0.0176`$ and $`\widehat{d}=0.100`$. For each test, $`\gamma =5/3`$. For Hachisu’s data the virial theorem diagnostic $`𝒱`$ is typically less than a few times $`10^4`$. The test results show that the present code and Hachisu’s code differ by less than $`1\%`$ in every case. Subsequent testing has shown that the density update in Eq. (16) is no better, in terms of accuracy and convergence rate, than the usual self–consistent field method update (17). After $`30`$ Newton–Raphson iterations, the results obtained from the two update formulas agree to a few parts in $`10^9`$, at worst. For both update formulas, the errors (as compared to the $`30`$–iteration results) were at most a few tenths of a percent after $`10`$ iterations.
warning/0004/hep-th0004003.html
ar5iv
text
# References 1. Introduction The observed smallness of the cosmological constant requires a remarkable cancellation of all vacuum energy contributions . Supersymmetry seems at present the only known symmetry that could naturally explain this cancellation, but thus far no mechanism for supersymmetry breaking is known that would not destroy this property. Nonetheless, in searching for a possible resolution of the cosmological constant problem, it seems natural to include supersymmetry as a central ingredient. What would then be needed, however, is a mechanism – some UV/IR correspondence – by which the short distance cancellations of supersymmetry can somehow be translated into a long distance stability of the cosmological evolution equations. In this paper we would like to propose a possible candidate for such a mechanism. The basic idea is as follows. Consider a string realization of the Randall-Sundrum compactification scenario . In such a compactification, space-time consists of a slice of $`AdS_5`$ (times some small compact 5-manifold) bounded by two brane-like structures, which we will call the Planck and matter brane, respectively. Both branes must be thought of as effective descriptions of more elaborate structures, defined by the full high energy string theory and the low energy quantum field theory , respectively. Now, via the holographic UV/IR duality of the AdS/CFT correspondence , we may identify the 5-d physics at different radial locations in the bulk region with 4-d physics at corresponding intermediate energy scales. The fact that the matter degrees of freedom are localized on a matter brane separate from the Planck brane, is therefore just a reflection of the fact that they spend most of their time at much lower energy scales than the Planck scale. This 4-d matter still feels ordinary long distance gravity via its interaction with the 5-d bulk . The two ingredients that we would like to add to this scenario are the following: * Since all observable (non-supersymmetric) matter is localized at the matter brane, it seems an allowed assumption that the physics on and near the Planck brane is supersymmetric, at least to a very good approximation. In particular, the classical action that describes the effective dynamics of the Planck brane could reasonably be taken to be of a particular supersymmetric form. Via the AdS/CFT dictionary, this assumption can be thought of as the holographic image of requiring 4-d high energy supersymmetry. * Since almost all of the 5-d volume is contained in the Planck region, the large distance structure of the effective 4-d geometry will be determined by the shape of the Planck brane. Locality of the 5-d supergravity in turn implies that this shape is determined by local 5-d equations of motion, satisfied by the 5-d supergravity fields in the direct neighborhood of the Planck brane. Both these assumptions, if indeed realizable, are most likely valid only within a certain level of approximation. It is clear however that, if both are truly valid, combined they imply that the equations of motion that determine the large scale 4-d geometry can indeed be protected by supersymmetry. This possible reappearance of supersymmetry at large distance scales can be seen as related to the fact that – unlike in the conventional non-compact set-up – the UV/IR mapping of the AdS/CFT correspondence now acts on one single space-time that combines both the 4-d boundary field theory and the 5-d bulk gravity. Via this UV/IR duality, the infra-red bulk region of the AdS-space near the Planck brane becomes the natural home base for both the shortest and longest distance physics. In the following, we will try to test the consistency of these two assumptions. To this end, we will address the following obvious and most serious counter-argument. Intuitively, one would expect that the low energy matter sector on the matter brane will produce some quite arbitrary effective tension, that (without some unnatural or non-local fine-tuning) is expected to induce a non-zero cosmological constant for the total 4-d effective field theory. If indeed present, its backreaction would curve the Planck brane and consequently break its supersymmetry. A different version of the same objection is that the AdS/CFT dictionary tells us that the normal variations of the local supergravity fields near the Planck brane in fact know about low energy quantities of the dual field theory, such as vacuum expectation values, etc. In particular, the normal variation of the bulk metric (or more precisely, the extrinsic curvature at the Planck brane ) knows about the full vacuum energy produced by the low energy field theory. It would seem quite unnatural to expect that the Planck brane dynamics could be chosen such that, without any pre-knowledge of the IR dynamics, it exactly cancels this matter contribution to the vacuum energy. In the following sections we will describe a mechanism that will neutralize this counter-argument. In the final section we address some other aspects of our proposal, and discuss its relation with other recently proposed scenarios . 2. Set-up We first briefly state our assumptions about the physics inside the three different regions – the bulk region, the matter brane, and the Planck brane. The bulk supergravity The 5-dimensional bulk region is a negatively curved space, with a varying 5-d cosmological constant $`V(\varphi )`$, typically of planckian magnitude and dependent on various bulk matter fields $`\varphi `$. The metric thus takes a warped form, such that the ratio between the warp factor on the Planck and matter brane matches the hierarchy between the Planck and typical matter scale . The classical equations of motion for the bulk region are prescribed by a 5-d gauge supergravity action, of the schematic form (omitting fermionic fields) $$S_{\mathrm{bulk}}=d^5x\sqrt{G}(V(\varphi )+\kappa R\frac{1}{2}(\varphi )^2).$$ (1) The matter brane The matter brane hosts all visible matter. It typically represents a strongly curved or even singular region of the 5-d geometry. Because the supergravity approximation breaks down inside this region, we will introduce an artificial line of separation between the matter brane and the bulk, located at some arbitrary scale (see fig 1). On this line, we consider the values of the 5-d supergravity fields $`(g,\varphi )`$. Here $`g`$ is short-hand for the 4-d metric $`g_{\mu \nu }`$ and $`\varphi `$ determine the couplings and expectation values of the matter theory. The matter brane effective action $$\mathrm{\Gamma }_{\mathrm{matter}}(g,\varphi )$$ (2) is obtained by integrating out all degrees of freedom to the left of the dividing line, with fixed boundary values for $`g`$ and $`\varphi `$. Since supersymmetry is broken in this sector, we will not require any special symmetries of this effective action. Moreover, we will in principle allow the matter brane to have a finite temperature and/or matter density, which may vary with time and position; $`\mathrm{\Gamma }`$ will then contain specific terms whose variation with respect to $`g_{\mu \nu }`$ will equal the corresponding stress-energy contributions. The effective action $`\mathrm{\Gamma }_{\mathrm{matter}}`$ will not be completely arbitrary, however: consistency of the geometric set-up requires that the total partition function of the bulk and matter region does not depend on the location of the artificial dividing line between the two regions. In AdS/CFT dual terms, this means that we imagine that around the energy scale corresponding to this line, the matter theory has become equivalent to the exact holographic dual of the 5-d bulk supergravity. Shifting the position of the line then corresponds to changing some arbitrarily chosen RG scale . We will formulate this invariance requirement in more detail in section 5. The Planck brane The region near the Planck brane represents the internal compactification geometry of the full string theory outside the AdS-like region . In accordance with the holographic interpretation of the 5-d geometry, its internal structure is directly probed only by Planck scale 4-d physics. In our scenario, this physics is assumed to be supersymmetric. To make this symmetry manifest, we will now postulate, as an e͡ffective description of the Planck brane brane dynamics, the following special boundary action $$S_{\mathrm{planck}}(g,\varphi )=d^4x\sqrt{g}W(\varphi ).$$ (3) Here $`W(\varphi )`$ is the superpotential of the 5-d supergravity, related to $`V(\varphi )`$ via $$V(\varphi )=\frac{1}{2}(_\varphi W(\varphi ))^2\frac{1}{3}W(\varphi )^2.$$ (4) In particular, we will be assuming that the Planck brane will not carry any independent matter density on its own world volume.<sup>1</sup><sup>1</sup>1This is a simplification, that in effect amounts to a restriction on the 4-d physics, namely that it is always sub-Planckian. This in particular means that we will be ignoring the possible stress-energy contributions of gravitationally collapsed matter inside 4-d black holes. As we will see shortly, the choice of action (3) implies that the Planck brane geometry satisfies an effective 4-d Einstein equation with vanishing cosmological constant. Ultimately, it will therefore be important to determine how well protected (3)-(4) are against quantum corrections induced via the presence of the non-supersymmetric matter brane. For now, however, we will simply define the classical action of our model via (3)-(4). 3. Equation of Motion I The boundary conditions that follow from (3) are $$\theta _{\mu \nu }\theta g_{\mu \nu }=W(\varphi )g_{\mu \nu }_n\varphi =_\varphi W(\varphi ),$$ (5) where $`g_{\mu \nu }`$ denotes the induced metric and $`\theta _{\mu \nu }`$ the extrinsic curvature of the Planck brane; $`_n`$ the derivative normal to its world-volume. Now the normal component of the 5-d bulk Einstein equation, when written in terms of 4-d geometric quantities, reads $$\frac{1}{4}(\theta ^2(\theta _{\mu \nu })^2)\frac{1}{2}(_n\varphi )^2+V(\varphi )+\kappa R\frac{1}{2}(\varphi )^2=0.$$ (6) Here $``$ and $`R`$ are the 4-d gradient and curvature scalar. When combined with (5) and (4), we deduce from (6) that on the Planck brane world volume $$\kappa R\frac{1}{2}(\varphi )^2=0.$$ (7) The classical supersymmetric Planck brane geometry thus solves the trace of the 4-d Einstein equations, with possibly non-zero matter density, but always with zero cosmological constant. This result holds independently of what happens inside the bulk: possible corrections to (7) can only arise from local physics that happens at or near the location of the Planck brane. In the following sections we will present a rederivation of (7) within the quantum context. 4. Partition Function Consider the total quantum mechanical partition function of our system. We can first divide it into three parts, corresponding to the three sub-regions. The bulk supergravity partition function can best be thought of as an evolution operator $`\widehat{𝒰}`$ that describes the propagation through the bulk from the matter to the Planck brane. Its matrix element between eigen states $`|g,\varphi `$ of the boundary fields has the formal path integral expression $$g_1,\varphi _1|\widehat{𝒰}|g_2,\varphi _2=\underset{(g_1,\varphi _1)}{\overset{(g_2,\varphi _2)}{}}[dG][d\varphi ]e^{{\scriptscriptstyle \frac{i}{\mathrm{}}}S_{\mathrm{bulk}}(G,\varphi )}.$$ (8) Since in our set-up the matter brane is defined as the region left to some quite arbitrary line of separation with the bulk (see fig 1), its partition function is indeed most appropriately thought of as a wave-function $$\mathrm{\Psi }_{\mathrm{matter}}|g_1,\varphi _1=e^{{\scriptscriptstyle \frac{i}{\mathrm{}}}\mathrm{\Gamma }_{\mathrm{matter}}(g_1,\varphi _1)}.$$ (9) Similarly we can define $$g_2,\varphi _2|\mathrm{\Psi }_{\mathrm{planck}}=e^{{\scriptscriptstyle \frac{i}{\mathrm{}}}S_{\mathrm{planck}}(g_2,\varphi _2)}.$$ (10) The total partition function is obtained by taking the inner-product of these two wave-functions, with the bulk evolution operator inserted in between $$Z=\mathrm{\Psi }_{\mathrm{matter}}|\widehat{𝒰}|\mathrm{\Psi }_{\mathrm{planck}}.$$ (11) Inserting the above definitions, this matrix element represents the complete functional integral over all matter and gravity fields. We will imagine that it can be given a well-defined definition via the underlying fundamental string theory. 5. Holographic RG Now let us consider the evolution of the matter state under local variations of the position of the dashed line in fig 1. Thinking about the radial direction as a Euclidean time directon, we may identify the corresponding Hamilton operator $`\widehat{}`$ of the bulk supergravity as the generator of these variations. However, as in any gravity theory, physical wave functions should be invariant under local variations of the particular time-slicing. We may write this condition as $$\widehat{}\widehat{𝒰}|\mathrm{\Psi }_{\mathrm{matter}}=\mathrm{\hspace{0.17em}0}.$$ (12) The Hamilton operator associated with the bulk action (1) reads $$\widehat{}=\frac{\mathrm{}^2}{\sqrt{g}}(\frac{1}{3}\left(\frac{\delta }{\delta g_\lambda ^\lambda }\right)^2\left(\frac{\delta }{\delta g^{\mu \nu }}\right)^2\frac{1}{2}\left(\frac{\delta }{\delta \varphi }\right)^2)+\sqrt{g}(V(\varphi )+\kappa R\frac{1}{2}(\varphi )^2).$$ (13) Classically, the condition $`\widehat{}=0`$ is simply the 5-d bulk Einstein equation (6). Indeed, we emphasize that eqn (12) is not some special symmetry requirement on the matter sector, but a property of any state that has traveled through the bulk supergravity region. After applying the AdS/CFT dictionary, our geometric set-up is equivalent to assuming that the complete matter theory above some energy scale – corresponding roughly to the line of separation between the matter brane and the bulk – becomes equivalent to the exact holographic dual of the bulk supergravity. From this 4-d perspective, the relation (12) acquires a new meaning as the evolution of the matter partition function under the holographic RG flow . Note however, that for finite rank $`N`$ and gauge coupling, the bulk theory is a complete string theory with finite string length and string coupling. The constraint (12) for finite $`\mathrm{}`$ includes $`1/N`$ effects, but the Hamiltonian (13) will receive various string corrections. We expect however that these will not significantly alter the basic form of the RG equation . 6. Equation of Motion II In our set-up, the boundary state at the Planck brane has been chosen to be invariant under 4-d global supersymmetry transformations, $`Q_\alpha |\mathrm{\Psi }_{\mathrm{planck}}=0`$. Using the explicit form (10) of $`|\mathrm{\Psi }_{\mathrm{planck}}`$ and the classical relation (4), we find that $$\widehat{}|\mathrm{\Psi }_{\mathrm{planck}}=\sqrt{g}(\kappa R\frac{1}{2}(\varphi )^2)|\mathrm{\Psi }_{\mathrm{planck}}.$$ (14) Here we used that the one-loop term, proportional to $`\mathrm{}(\frac{4}{3}W\frac{1}{2}_\varphi ^2W)`$, is cancelled by the corresponding fermionic contribution, which is true provided the fermionic part of the wavefunction is chosen such that $`|\mathrm{\Psi }_{\mathrm{planck}}`$ indeed represents a supersymmetric ground state. Combined with the $`\widehat{}=0`$ condition (12), eqn (14) implies $$\kappa R\frac{1}{2}(\varphi )^2=\frac{1}{Z}\mathrm{\Psi }_{\mathrm{matter}}|\widehat{𝒰}\widehat{}|\mathrm{\Psi }_{\mathrm{planck}}=\mathrm{\hspace{0.17em}0}.$$ (15) This equation generalizes the classical result of section 3. As emphasized before, possible corrections to this identity can only arise from matter and/or supersymmetry breaking terms at the direct location of the Planck brane. 7. Supersymmetry breaking Suppose we insist on maintaining the Planck brane as supersymmetric as needed to make the cosmological constant as small as observed today. The question then arises as to whether the evolution into the 5-d bulk, that describes the holographic RG flow towards the IR, can still lead to a realistic non-supersymmetric low energy field theory on the matter brane. A possible scenario by which this may happen is when the RG flow at some scale enters a strongly coupled region (coupling $`g`$ of order 1 or larger) from both the 5-d and 4-d point of view. The dynamics in this region could then be sufficiently non-trivial to induce dynamical breaking of 4-d supersymmetry. To gain some insight into how this may happen, suppose we define the matter state by means of some non-supersymmetric low energy matter sector, so that $$Q_\alpha |\mathrm{\Psi }_{\mathrm{matter}}\mathrm{\hspace{0.17em}0}.$$ (16) The vacuum amplitude is then still supersymmetric, since the Planck state via (11) projects out the supersymmetric part of the matter wave-function. The violation of supersymmetry becomes visible, however, as soon as we consider non-trivial expectation values. Generalizing our prescription of section 4, we can formally represent these via $$\mathrm{\Psi }_{\mathrm{matter}}|𝒪_{phys}\widehat{𝒰}|\mathrm{\Psi }_{\mathrm{planck}}$$ (17) where $`𝒪_{phys}`$ represents some physical observable inside the matter brane region. Suppose we now try to derive a supersymmetry Ward identity. Because of (16), it will be violated $$\mathrm{\Psi }_{\mathrm{matter}}|[Q_\alpha ,𝒪_{phys}]\widehat{𝒰}|\mathrm{\Psi }_{\mathrm{planck}}\mathrm{\hspace{0.17em}0}.$$ (18) The 4-d low energy observer would thus conclude that supersymmetry is broken. On the other hand, since physical operators $`𝒪_{phys}`$ must commute with the generator $`\widehat{}`$, the derivation of eqn (15) as given in the previous section still goes through (see also Appendix A). 8. 4-d Effective Field Theory The reasoning in the above sections is rather formal. To make things somewhat more concrete, let us from now on assume that the 4-d high energy gauge theory is at large $`N`$ and strong ’t Hooft coupling, so that the 5-d supergravity is in its classical regime. In terms of pure 4-d language the situation is then summarized as follows. Consider the complete matter effective action, computed in some fixed 4-d background $`(g,\varphi )`$, obtained by integrating out both the low energy energy matter, as well as the high energy matter dual to the 5-d bulk supergravity. We will still call this effective action $`\mathrm{\Gamma }_{\mathrm{matter}}`$, since indeed it is just the same action (2) evolved via the RG flow (12) to a larger scale factor for $`g_{\mu \mu }`$. Note that $`\mathrm{\Gamma }_{\mathrm{matter}}`$ now contains the full Einstein term of the 4-d gravity. The total 4-d action is obtained by just adding the Planck brane action (3) $$\mathrm{\Gamma }_{\mathrm{total}}(g,\varphi )=\mathrm{\Gamma }_{\mathrm{matter}}(g,\varphi )+d^4x\sqrt{g}W(\varphi ).$$ (19) The claim is that, regardless of the details of the low energy matter theory, the classical equations of motion of this action $`{\displaystyle \frac{1}{\sqrt{g}}}{\displaystyle \frac{\delta \mathrm{\Gamma }_{\mathrm{matter}}}{\delta g^{\mu \nu }}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}W(\varphi )g_{\mu \nu }`$ $`{\displaystyle \frac{1}{\sqrt{g}}}{\displaystyle \frac{\delta \mathrm{\Gamma }_{\mathrm{matter}}}{\delta \varphi }}`$ $`=`$ $`_\varphi W(\varphi )`$ (20) are such that the trace of the Einstein equation always takes the form (7). This is possible since $`\mathrm{\Gamma }_{\mathrm{matter}}`$ is not completely arbitrary, but due to the high energy equivalence with the 5-d supergravity satisfies (12)-(13). In the classical approximation, this identity reduces to the Hamilton-Jacobi relation, which looks identical to eqn. (6) with the replacement $`\theta _{\mu \nu }\theta g_{\mu \nu }`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{g}}}{\displaystyle \frac{\delta \mathrm{\Gamma }_{\mathrm{matter}}}{\delta g^{\mu \nu }}}`$ $`_n\varphi `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{g}}}{\displaystyle \frac{\delta \mathrm{\Gamma }_{\mathrm{matter}}}{\delta \varphi }}.`$ (21) Via this same replacement, the classical equations of motion (20) just amount to the boundary conditions (5). Combining these two results gives (7), just as before. 9. Stabilizing the Matter Brane Finally, let us address how to stabilize the matter brane. This is an important issue, since its location relative to the Planck brane represents an invariant physical scale . As expected, some subtleties will arise here. A first subtlety is that the 5-d physics in the matter brane region is likely to be singular. Thus far, however, we did not need to worry about this issue, because we chose to describe the matter brane region by means of the dual low energy field theory. No special assumptions about the properties of the singularity were needed, except that at some distance away from it we can use the geometric supergravity language to derive the constraint (12). It is not important for our argument, for example, whether there are discrete or continuously many solutions to this constraint. Regardless of this issue, it is clear that the matter brane effective action should be allowed to be as arbitrary as possible. Nonetheless one would hope to naturally stabilize its location by means of its interaction with the Planck brane. We will first describe two possible mechanisms, which however both will fall short in that they either remain unstable or need unnatural fine-tuning. We will then describe a third scenario that will resolve these problems. Attempt 1: Goldberger-Wise mechanism Suppose that, in spite of the fact that it describes a singular space-time region, we would assume that the matter brane dynamics is well approximated by that of some classical brane with some arbitrary tension $`\mathrm{\Lambda }(\varphi )`$. We can then look for a classically stable location for matter brane as follows . A static bulk solution, when matched onto the supersymmetric boundary conditions set by the Planck brane, is described by scalar fields $`\varphi (r)`$ and a metric $$ds^2=a^2(r)\eta _{\mu \nu }dx^\mu dx^\nu +dr^2,$$ (22) satisfying the supersymmetric flow equations $$\frac{a^{}}{a}=\frac{1}{6}W(\varphi )\varphi ^{}=_\varphi W.$$ (23) From this one finds that the matching relations at the matter brane, that are required for there to be a flat solution, are that (cf. eqn (20)) $$_\varphi \mathrm{\Lambda }(\varphi _c)=_\varphi W(\varphi _c)\mathrm{\Lambda }(\varphi _c)=W(\varphi _c)$$ (24) for some critical value of $`\varphi _c`$ for the scalar fields. The first equation generically gives at most a discrete set of solutions for $`\varphi _c`$. The second relation, however, is then valid only if the value of $`\mathrm{\Lambda }`$ at such a critical point $`\varphi _c`$ is exactly equal to that of the superpotential $`W`$. This amounts to an unnatural fine-tuning of the matter brane action. However, even if we would choose $`\mathrm{\Lambda }(\varphi )`$ such that this condition is satisfied, the equations of motion (24) still fail to stabilize the relative location of the Planck and matter brane, as they do not pick out one particular preferred relative ratio for the scale factors $`a`$ at the two branes. Because of both these problems, we will instead choose a different route. Attempt 2: Hamilton-Jacobi action In accordance with our definition of the matter brane as the region behind the dashed line in fig 1, we will now reinstate the condition that its effective action satisfies (12). Let us further assume that for slowly varying fields it takes the general form $$\mathrm{\Gamma }_{\mathrm{matter}}=\sqrt{g}(\mathrm{\Lambda }(\varphi )+\mathrm{\Phi }(\varphi )R+\mathrm{})$$ (25) Using the classical supergravity approximation, (12) then reduces to the Hamilton-Jacobi relation $$\frac{1}{2}(_\varphi \mathrm{\Lambda })^2\frac{1}{3}\mathrm{\Lambda }^2+(_\varphi \mathrm{\Lambda }_\varphi \mathrm{\Phi }\frac{1}{3}\mathrm{\Lambda }\mathrm{\Phi })R+\mathrm{}=V+\kappa R+\mathrm{}$$ (26) with $`V`$ and $`\kappa `$ the 5-d potential and Newton constant. Apart from this constraint, the functions $`\mathrm{\Lambda }(\varphi )`$ and $`\mathrm{\Phi }(\varphi )`$ will be allowed to be arbitrary. In principle, one could read (26) as a condition on classical field configurations $`(g_c,\varphi _c)`$, that determines the matter brane curvature $`R`$ for given tension $`\mathrm{\Lambda }(\varphi _c)`$, etc. In our set-up, however, (25) defines a true Hamilton-Jacobi action, for which (26) in fact amounts to a functional identity valid for all field configurations $`(g,\varphi )`$. So in particular the relation $$\frac{1}{2}(_\varphi \mathrm{\Lambda })^2\frac{1}{3}\mathrm{\Lambda }^2=V$$ (27) represents a relation between $`\mathrm{\Lambda }(\varphi )`$ and $`V(\varphi )`$ that holds for all values of $`\varphi `$. Next let us again look for a consistent flat matter brane solution in the presence of the Planck brane. Following the same derivation as before, we again arrive at the condition (24) for $`\mathrm{\Lambda }(\varphi )`$. Now the situation is somewhat different: if we can find a solution to the first relation in (24) then via (27) we automatically satisfy the second relation. At first this seems like good news, but in fact the situation is now worse than before: for generic $`V(\varphi )`$ the only solution to the H-J relation (27) that will ever be tangent to $`W(\varphi )`$ is $`W(\varphi )`$ itself. Hence if we require that (24) is satisfied, we must have that $`\mathrm{\Lambda }(\varphi )=W(\varphi )`$. But this is a hyper fine-tuned situation, since it says that the matter brane is just as supersymmetric as the Planck brane. This is not what we want. Attempt 3: Holographic effective action The problem we just encountered is in essence Weinberg’s counter-argument against the use of various adjustment mechanisms for cancelling the cosmological constant . We now propose a mechanism that will evade this counter-argument. See also . There are many indications that the matter brane action $`\mathrm{\Gamma }_{\mathrm{matter}}(\varphi ,g)`$, at finite values for the scale factor of $`g_{\mu \nu }`$, in fact corresponds to a Wilsonian effective action of the dual 4-d field theory, defined with a finite UV cut-off $`ϵ`$. Therefore, a more accurate formula for the matter effective action is $$\mathrm{\Gamma }_{\mathrm{matter}}=\sqrt{g}(\mathrm{\Lambda }(\varphi ,ϵ)+\mathrm{\Phi }(\varphi ,ϵ)R+\mathrm{}).$$ (28) This cut-off dependence of $`\mathrm{\Gamma }_{\mathrm{matter}}`$ will turn out to be crucial. The holographic dictionary furthermore relates variations in $`ϵ`$ to variations in the scale factor of $`g_{\mu \nu }`$ via $$ϵ\frac{}{ϵ}=\mathrm{\hspace{0.17em}2}g^{\mu \nu }\frac{\delta }{\delta g^{\mu \nu }}=a\frac{}{a}$$ (29) with $`a`$ the overall warp factor in (22). We can use this relation to replace the short-distance cut-off $`ϵ`$ in (28) by $`a`$. The effective action (28) thus acquires a new and unexpected dependence on $`g_{\mu \nu }`$. The holographic relation (29) between RG transformations and constant Weyl rescalings, and the resulting extra dependence of $`\mathrm{\Gamma }_{\mathrm{matter}}`$ on $`g_{\mu \nu }`$, has quite non-trivial implications. In particular, it implies that the Hamilton-Jacobi relation (27) and matching relations (24) now take the more general form: $$\frac{1}{2}(_\varphi \mathrm{\Lambda })^2\frac{1}{48}\left(4\mathrm{\Lambda }a\frac{\mathrm{\Lambda }}{a}\right)^2=V,$$ (30) $$_\varphi \mathrm{\Lambda }(\varphi _c,a)=_\varphi W(\varphi _c),$$ (31) $$\left(4a\frac{}{a}\right)\mathrm{\Lambda }(\varphi _c,a)=\mathrm{\hspace{0.17em}4}W(\varphi _c).$$ (32) In fact, now we are in business: it is easy to show that these equations in general do have solutions. The reason is that, via the non-trivial dependence on $`a`$, we now have an extra parameter that we can adjust. Things are however still less trivial than one might expect: the above equations have an RG invariance and generically allow for a whole critical line of solutions $`\varphi _c(a)`$, generated by the holographic RG-flow (see eqn (23)): $$a\frac{}{a}\varphi _c(a)=\beta _\varphi (\varphi _c(a))$$ (33) $$\beta _\varphi =\frac{6_\varphi W}{W}.$$ (34) Eqn (31) does therefore not pick out one particular preferred value for the scale $`a`$ at which the matter brane is localized: $`a`$ is still left free. This is in accordance with the definition of the matter brane as the region behind a relatively arbitrary line of separation with the bulk region. Nonetheless, we have achieved a stabilization of the matter brane location. For a given matter brane effective action (28), eqn (31) selects out a particular RG trajectory. This trajectory will generically become unstable around a particular scale $`a_{crit}`$, where one or more of the scalar fields starts to run off a steep slope and induce a large 5-d curvature. This critical scale $`a_{crit}`$ is the natural scale of the matter brane. and also the scale where non-supersymmetric physics can start to take place. 10. An Explicit Example To illustrate the above stabilizing mechanism, we now describe a specific example. Consider an effective potential of the matter brane of the form $$\mathrm{\Lambda }(\varphi ,a)=W(\varphi )+\omega (\varphi ,a)$$ (35) where $`\omega (\varphi ,a)`$ can be treated as a small perturbation. Eqns (30)-(32) then reduce to $$\left(a\frac{}{a}\beta _\varphi _\varphi 4\right)\omega (\varphi ,a)=\mathrm{\hspace{0.17em}0}$$ (36) $$_\varphi \omega (\varphi _c,a)=\mathrm{\hspace{0.17em}0}$$ (37) $$\left(a\frac{}{a}4\right)\omega (\varphi _c,a)=0,$$ (38) with $`\beta _\varphi `$ as defined in (34). The first equation states that $`\omega `$ is invariant under the holographic RG flow generated by $`W`$. The last two equations further show that $`\omega `$ is the potential that needs to be minimized to obtain the classical vacuum values of the fields. We can thus identify $`\omega (\varphi ,a)`$ with the total effective potential. Now as a simple example, let us consider the case with just one scalar field $`\varphi `$ – the dilaton – and assume that $$W(\varphi )=e^{{\scriptscriptstyle \frac{1}{\sqrt{3}}}\varphi }.$$ (39) is $$\beta _\varphi =2\sqrt{3}.$$ (40) Now suppose that $`\omega (\varphi ,a)`$ takes the form<sup>2</sup><sup>2</sup>2 This potential has been deliberately chosen to take the typical form of a one-loop correction, obtained by integrating out a 4-d matter field, in this particular case with a mass proportional to $`e^{\frac{1}{2\sqrt{3}}\varphi }`$. The extra $`a`$-dependence of the potential can be derived for example by computing the one-loop determinant via dimensional regularization, in the 4-d induced background metric $`g_{\mu \nu }=a^2\eta _{\mu \nu }`$ on the matter brane. $$a^4\omega (\varphi ,a)=\frac{1}{4}\lambda a^4e^{{\scriptscriptstyle \frac{2}{\sqrt{3}}}\varphi }\left(\mathrm{log}\left(\mu a^2e^{{\scriptscriptstyle \frac{1}{\sqrt{3}}}\varphi }\right)\frac{1}{2}\right)$$ (41) with $`\lambda `$ and $`\mu `$ some free parameters. It is easily verified that this expression, for arbitrary $`\lambda `$ and $`\mu `$, solves the linearized RG flow equation (36). The minimum of the effective potential lies at $$\mu a^2e^{{\scriptscriptstyle \frac{1}{\sqrt{3}}}\varphi _c}=\mathrm{\hspace{0.17em}1}$$ (42) This minimum describes an RG trajectory $$\varphi _c(a)=\sqrt{3}\mathrm{log}(\mu a^2).$$ (43) Via (23) this corresponds to a classical bulk geometry (22), with warp factor $$a(r)=\sqrt{1\frac{r}{3\mu }}0r3\mu .$$ (44) Here we normalized $`a`$ such that the Planck brane is located at $`a=1`$. We see that the $`r`$-direction describes a compact interval: the geometry closes off by means of a naked singularity, located at the critical distance $`r_{crit}=3\mu `$ from the Planck brane. The matter brane is the strongly curved neighborhood of the naked singularity. Its location is therefore indeed dynamically stabilized by the presence of the Planck brane. Furthermore, by virtue of our presumption that the Planck brane remains supersymmetric, both branes are flat. This result holds for arbitrary values of the free parameters $`\mu `$ and $`\lambda `$. Note that the effective 4-d cosmological constant of the matter brane is indeed not equal to the minimum value of the effective potential $$a^4\omega (\varphi _c,a)=\frac{\lambda }{8\mu ^2}.$$ (45) Rather, it is equal to its variation with respect to the overal scale $`a`$ $$\mathrm{\Lambda }_{\mathrm{eff}}=a\frac{d}{da}(a^4\omega (\varphi _c,a))=\mathrm{\hspace{0.17em}0}.$$ (46) This last equality of course follows by inspection from (45), but more fundamentally, arises as a direct consequence of the (linearized) RG constraint (36). The RG trajectory (43) and geometry (44) describe a supersymmetric bulk solution. This is the classical version of the result that in the vacuum amplitude (11), the Planck state always projects out the supersymmetric component of the matter state. We expect, however, that for general non-supersymmetric observables, the non-supersymmetric form (35)-(41) of the matter brane effective action will become manifest. Supersymmetry remains intact only for the BPS-type limit $`\lambda 0`$ and $`\mu 0`$ (in some fixed ratio). 11. Discussion We have explored some physical consequences of a Randall-Sundrum scenario , in which the Planck brane region is assumed to be supersymmetric. Via the UV/IR mapping of the AdS/CFT duality, this condition appears equivalent to that of 4-d high energy supersymmetry. As we have shown, however, the same supersymmetry also stabilizes the large scale 4-d geometry. It appears therefore that our set-up implies a stronger restriction on the 4-d effective field theory than ordinary high energy supersymmetry. From the higher dimensional viewpoint this stronger restriction seems natural, however, as it pertains to just one single region of the 5-d geometry.<sup>3</sup><sup>3</sup>3In , a somewhat similar suggestion was made for a scenario in which our 4-d world might be stabilized via the embedding into an unobservable 5-d supersymmetric world. Our set-up differs in a number of aspects from other recent proposals . In these papers, the effective 4-d flatness appears as a consequence of a self-tuning mechanism of a single 4-d brane-world, embedded inside a fine-tuned 5-d bulk. This mechanism however relies on rather strong assumptions about the physics of the IR singularity of the 5-d geometry. In our case, on the other hand, the 4-d stability is the consequence of a fine-tuned Planck brane action. This fine-tuning is justified by our assumption of supersymmetry, and furthermore local, both from the 5-d and the 4-d RG perspective. Given our observation that the 4-d flatness is proportional to the amount of supersymmetry breaking at the Planck brane, it will clearly be important to look for possible mechanisms that may prevent or suppress the mediation of supersymmetry breaking via the bulk. It seems clear that holography (both 5-d and 4-d) will be an important ingredient in this discussion. First, it will be important to know how best to describe the local degrees of freedom on the Planck brane. Via the string theory description , much is known in principle about its internal structure, but its effective dynamics still seems hard to extract. Let us nonetheless imagine computing some supersymmetry violating one loop correction to the Plank brane physics, induced by a bulk graviton traveling back and forth to the matter brane region. A priori, one would not expect any large suppression from this propagator, and it would thus appear that supersymmetry will indeed be violated on the Planck brane at roughly the same length scale as that set by the matter brane location. This conclusion, however, disregards the implications of 5-d holography, which suggest that the local quantum fluctuations at the Planck brane – in as far present – all have roughly Planckian frequencies and wave-lengths. If this is the case, the 5-d graviton modes emitted by these fluctuations must also have Planckian wavelengths along the brane direction. Such modes, however, decay exponentially along the $`r`$-direction with a Planckian length. Hence, if this physical picture is correct, this would indeed give a mechanism by which the bulk mediation of supersymmetry breaking gets exponentially suppressed via the separation between the two branes. Additional restrictions on the use of naive effective field theory may come from 4-d holography. It is clear, for example, that a direct AdS/CFT interpretation of the 5-d bulk region, as dual to a 4-d quantum field theory with Planckian cut-off, still introduces far too many degrees of freedom. It is tempting to speculate that supersymmetry near the Planck region could be instrumental in hiding this excess in degrees of freedom, thus protecting the 4-d holographic bound. Finally: the UV/IR connection explored here, just like all other ones pointed out earlier in , seems directly connected with the fundamental equivalence between open string loops and closed string propagators. Roughly, our assumption here has been that long closed string propagators and short open string loops are protected by supersymmetry, even though short closed string propagators and long open string loops are not. It is reassuring that these two types of diagrams represent clearly separated regions in the parameter space of open string Feynman diagrams. Acknowledgements This work is supported by NSF-grant 98-02484. I would like to thank Micha Berkooz, Lisa Randall, Raman Sundrum and Erik Verlinde for helpful discussions. Appendix: A Small Matter Perturbation Here we consider the effect of a small matter perturbation on the geometry of the Planck brane. The perturbation induces a small extra contribution $`\delta \mathrm{\Gamma }(g,\varphi )`$ to the matter+bulk effective action. The variation of this extra contribution with respect to the metric and field $`\varphi `$ represent the expectation values of the corresponding dual operators $$\frac{1}{\sqrt{g}}\frac{\delta (\delta \mathrm{\Gamma })}{\delta g^{\mu \nu }}=T_{\mu \nu }\frac{1}{\sqrt{g}}\frac{\delta (\delta \mathrm{\Gamma })}{\delta \varphi }=𝒪_\varphi $$ (A.1) which represent the physical effect of the matter perturbation. The RG flow relation (12) must still hold for the total effective action, including the extra term. (This is true since we assume that all matter is located on the matter brane.) This RG condition implies that the above two expectation values cannot be independent. Working to linearized order in $`T_{\mu \nu }`$, and within the classical supergravity approximation, one finds that $$T=\beta _\varphi 𝒪_\varphi +\frac{6}{W}𝒪_\varphi 𝒪_\varphi $$ (A.2) Here $`T`$ is the trace of the stress energy tensor and $`\beta _\varphi `$ the holographic beta-function defined in (34). Now if we consider the linearized equations of motion of the small metric variation $`h_{\mu \nu }=\delta g_{\mu \nu }`$ induced by the matter perturbation, then because of the relation (A.2) it is possible to combine the trace of the $`h`$ equation with the $`\varphi `$ equation such that the source term $`T`$ cancels. The resulting equation takes the form $$\kappa \mathrm{}h+\frac{1}{2}(\varphi )^2=0.$$ (A.3) The trace of the stress energy tensor of matter away from the Planck brane thus always gets represented by means of variations in the $`\varphi `$ field.
warning/0004/gr-qc0004081.html
ar5iv
text
# Einstein–Proca model: spherically symmetric solutions ## I Introduction A massive vector meson (spin 1 particle with a non-trivial mass) is described by a one-form field which obeys the Proca wave equation . Early development of the Proca theory was concerned with the classical and quantum electrodynamics of a massive photon. However, strong experimental limits on photon’s mass (see, e.g., ) in combination with theoretical arguments based on the idea of gauge invariance (which ultimately led to the standard model of electroweak interactions) have closed the electrodynamical chapter in the history of this theory. A further discussion of the differences between the Proca and electromagnetic fields can be found in . At present, interest in the Proca field is twofold. Firstly, the Proca model presents a convenient theoretical “laboratory” for the study of Lagrangian and Hamiltonian theories with second class constraints . Secondly, although it is irrelevant for the electrodynamics, a massive vector meson often appears in the spectra of many non-trivial field theoretical models, including some classes of generalized theories of gravity. In connection with this, it is interesting to investigate the specific physical effects arising in such models due to the interaction of Proca particles with electromagnetic, gravitational and other physical and geometrical fields. The interaction of spin 1 field with electromagnetic field is known to be free of algebraic inconsistencies as well as of acausal wave propagation ($`v>c`$) when the coupling is minimal (or modified by the addition of an anomalous magnetic dipole moment). However, acausal propagation anomalies arise for more general interaction Lagrangians . Similarly, acausal propagation takes place (along with algebraic inconsistencies) for a Proca field coupled minimally to external torsion field . Different aspects of the interaction of classical and quantum vector field with torsion have been analyzed recently in . Early studies of the gravitational interaction of the Proca field were centered around the black hole issue. Qualitative analysis of the self-consistent Einstein-Proca system revealed the absence of black hole type solutions (possessing a regular horizon) with an external vector meson “hair” . At the same time several exact spherically symmetric solutions of the Proca wave equation on the classical Schwarzschild background spacetime were obtained . Assuming that a massive vector field source was located on a thin spherical shell outside the Schwarzschild horizon, it was demonstrated in that the meson field may change the structure of the spacetime near the central singularity. For point vector field sources located at the origin , or at a finite distance from the origin , it was shown that the range of the meson field is reduced by the metric gravitational field. The energy-momentum invariant was found to be divergent on the Schwarzschild horizon. However, it should be noted that, contrary to the Abelian Proca case, the non-Abelian massive vector field (with mass of a Yang-Mills field coming from a spontaneous symmetry breaking mechanism) may form a black hole type configuration . The results of numerical analysis of the spherically symmetric gravitationally interacting complex spin 1 field have been reported recently in . In this case the Einstein-Proca system admits everywhere regular “boson star”-type solutions (cf. with massive scalar boson stars ). A direct motivation for our current study comes from the metric-affine theory of gravity (MAG). In Einstein’s general relativity the spacetime geometry is described by the curvature 2-form $`R^{\alpha \beta }`$. In MAG two post-Riemannian structures are introduced: the 1-form of nonmetricity $`Q_{\alpha \beta }`$ and the torsion 2-form $`T^\alpha `$. For a comprehensive review of this theory see . Already the early investigations of the models with the simplest possible MAG Lagrangians, which include only a linear Hilbert term, quadratic segmental curvature invariant, and a single trace torsion or Weyl nonmetricity square term, have shown that an effective Einstein–Proca theory arises naturally from the vacuum MAG field equations (cf. also ). This result was subsequently extended to a very general family of MAG Lagrangians . In all these models the effective Proca field describes the triplet of post-Riemannian one-forms which are proportional to each other: the Weyl covector $`Q:=g^{\alpha \beta }Q_{\alpha \beta }/4`$, the torsion trace $`T:=e_\alpha T^\alpha `$, and the nonmetricity one-form $`\mathrm{\Lambda }:=\vartheta ^\alpha e^\beta Q_{\alpha \beta }Q`$. The mass of the effective vector particle is constructed from the coupling constants of the MAG Lagrangian. For a complete review of the known exact solutions of MAG see . In this paper we study the spherically symmetric static solutions of the coupled Einstein-Proca system of field equations. A preliminary analysis of the limiting cases of this problem shows a possibility of solutions which combine the exponential “Yukawa” type behavior of the Proca potential at the origin with the asymptotically Schwarzschild solution far away from the source. We will present the corresponding solutions which have been obtained by the application of numerical integration techniques. Our main conventions and notation are taken from . In particular, the $`\eta `$-basis of the exterior algebra is constructed from a coframe one-form $`\vartheta ^\alpha `$ with the help of the Hodge duality operator: $`\eta ^\alpha ={}_{}{}^{}\vartheta _{}^{\alpha },\eta ^{\alpha \beta }={}_{}{}^{}(\vartheta ^\alpha \vartheta ^\beta ),\eta ^{\alpha \beta \gamma }={}_{}{}^{}(\vartheta ^\alpha \vartheta ^\beta \vartheta ^\gamma )`$. The dual frame is denoted as $`e_\alpha `$. The Greek indices $`\alpha ,\beta ,\mathrm{}=0,\mathrm{},3`$ label anholonomic components, and the metric signature is $`(,+,+,+)`$. ## II Einstein-Proca theory The Lagrangian four-form of the Einstein-Proca system reads $$V=\frac{1}{2\kappa }R^{\alpha \beta }\eta _{\alpha \beta }\frac{1}{2}\left(dA{}_{}{}^{}dA+m^2A{}_{}{}^{}A\right),$$ (1) where $`\kappa `$ is the gravitational constant ($`\kappa =\mathrm{}^2`$) and $`m`$ is the rest mass of the vector field $`A`$. The corresponding field equations arise from the independent variation of the action with respect to the coframe and the Proca one-forms, and read: $`d{}_{}{}^{}dA+m^2{}_{}{}^{}A`$ $`=`$ $`0,`$ (2) $`{\displaystyle \frac{1}{2}}R^{\beta \gamma }\eta _{\alpha \beta \gamma }`$ $`=`$ $`\kappa \mathrm{\Sigma }_\alpha .`$ (3) Here the canonical energy-momentum three-form of the massive vector field $$\mathrm{\Sigma }_\alpha =\frac{1}{2}\{(e_\alpha dA){}_{}{}^{}dA(e_\alpha {}_{}{}^{}dA)dA+m^2[(e_\alpha A)^{}A+(e_\alpha ^{}A)A]\},$$ (4) represents the usual source of the gravitational field. It is worthwhile to recall the relationship of (1)-(3) to the MAG theory. As we have mentioned already, the same physical system arises in MAG as an effective system in which the effective covector Proca field is (in the notations of our previous paper ) $$A=\sqrt{z_4}k_0\varphi .$$ (5) Here $`\varphi `$ determines the three nontrivial post-Riemannian pieces of nonmetricity and torsion (the triplet of one-forms) $`{}_{}{}^{(1)}T_{}^{\alpha }={}_{}{}^{(3)}T_{}^{\alpha }=0,{}_{}{}^{(1)}Q_{\alpha \beta }^{}={}_{}{}^{(2)}Q_{\alpha \beta }^{}=0,`$ (6) $`Q=k_0\varphi ,\mathrm{\Lambda }=k_1\varphi ,T=k_2\varphi .`$ (7) The effective mass $`m^2`$ of the vector particle and the constants $`k_0,k_1,k_2`$ are constructed from the original coupling constants of the MAG Lagrangian which contains all possible quadratic invariants of the torsion and nonmetricity (11 terms) together with the linear Hilbert type term (multiplied by the constant $`\kappa `$) and the Weyl segmental curvature quadratic term (multiplied by the constant $`z_4`$). See and for more details (note, however, that in the present paper we assume that the cosmological constant is zero). ## III Spherically symmetric static case In terms of the local time and space coordinates $`(\tau ,r,\theta ,\varphi )`$, the general spherically symmetric ansatz for the coframe can be written as $$\vartheta ^{\widehat{0}}=fd\tau ,\vartheta ^{\widehat{1}}=\frac{g}{f}dr,\vartheta ^{\widehat{2}}=rd\theta ,\vartheta ^{\widehat{3}}=r\mathrm{sin}\theta d\varphi .$$ (8) The geometrical meaning of the function $`g(r)`$ becomes evident when one computes the volume four-form $$\eta =\vartheta ^{\widehat{0}}\vartheta ^{\widehat{1}}\vartheta ^{\widehat{2}}\vartheta ^{\widehat{3}}=gr^2\mathrm{sin}\theta d\tau drd\theta d\varphi .$$ (9) Thus, $`g(r)`$ measures the deviation of $`\eta `$ from the standard spherically symmetric spacetime volume form. In a regular oriented spacetime domain we naturally have to assume $$0<g(r)<\mathrm{}.$$ (10) The general static spherically symmetric configuration of the coupled Einstein-Proca system is described by the three functions $`f=f(r),g=g(r)`$, and $`u=u(r)`$ which enter the spherically symmetric ansatz for the Proca field as follows $$A=\frac{u}{rf}\vartheta ^{\widehat{0}}=\frac{u}{r}d\tau .$$ (11) Substitution of (8)-(11) into the Proca field equation (2) results in $$\left\{\frac{1}{r^2}\frac{f}{g}\left[\frac{r^2}{g}\left(\frac{u}{r}\right)^{}\right]^{}\frac{m^2u}{rf}\right\}\eta _{\widehat{0}}=0,$$ (12) or, equivalently, $$u^{\prime \prime }\frac{g^{}}{g}\left(u^{}\frac{u}{r}\right)\frac{m^2g^2}{f^2}u=0.$$ (13) A direct calculation of the energy-momentum 3-form yields $`\mathrm{\Sigma }_\alpha `$ $`=`$ $`{\displaystyle \frac{1}{2r^2g^2f^2}}\left\{f^2\left(u^{}{\displaystyle \frac{u}{r}}\right)^2+m^2g^2u^2\right\}\delta _\alpha ^{\widehat{1}}\eta _{\widehat{1}}`$ (15) $`+{\displaystyle \frac{1}{2r^2g^2f^2}}\left\{f^2\left(u^{}{\displaystyle \frac{u}{r}}\right)^2+m^2g^2u^2\right\}\left(\delta _\alpha ^{\widehat{0}}\eta _{\widehat{0}}+\delta _\alpha ^{\widehat{2}}\eta _{\widehat{2}}+\delta _\alpha ^{\widehat{3}}\eta _{\widehat{3}}\right).`$ For the sake of completeness we also write down the Einstein 3-form $`{\displaystyle \frac{1}{2}}R^{\beta \gamma }\eta _{\alpha \beta \gamma }`$ $`=`$ $`{\displaystyle \frac{f^2}{r^2g^2}}\left\{\left[2r{\displaystyle \frac{f^{}}{f}}+1{\displaystyle \frac{g^2}{f^2}}\right]\delta _\alpha ^1\eta _1+\left[2r\left({\displaystyle \frac{f^{}}{f}}{\displaystyle \frac{g^{}}{g}}\right)+1{\displaystyle \frac{g^2}{f^2}}\right]\delta _\alpha ^{\widehat{0}}\eta _{\widehat{0}}\right\}`$ (17) $`+{\displaystyle \frac{f^2}{r^2g^2}}\left\{r^2{\displaystyle \frac{f^{\prime \prime }}{f}}+\left(r{\displaystyle \frac{f^{}}{f}}\right)^2r^2{\displaystyle \frac{f^{}}{f}}{\displaystyle \frac{g^{}}{g}}+2r{\displaystyle \frac{f^{}}{f}}r{\displaystyle \frac{g^{}}{g}}\right\}\left(\delta _\alpha ^{\widehat{2}}\eta _{\widehat{2}}+\delta _\alpha ^{\widehat{3}}\eta _{\widehat{3}}\right).`$ Inserting (15) and (17) into (3) we find (after some algebra) that the equations corresponding to $`\alpha =\widehat{0},\widehat{1}`$ are $`2\left(r(f^2)^{}+f^2g^2\right)`$ $`=`$ $`\kappa m^2{\displaystyle \frac{g^2}{f^2}}u^2\kappa \left(u^{}{\displaystyle \frac{u}{r}}\right)^2,`$ (18) $`r(g^2)^{}`$ $`=`$ $`\kappa m^2{\displaystyle \frac{g^4}{f^4}}u^2.`$ (19) Furthermore, it is straightforward to see that the (second order) equation corresponding to $`\alpha =\widehat{2},\widehat{3}`$ is a consequence of (18)-(19) and (13). The scalar invariant $`|\mathrm{\Sigma }|:={}_{}{}^{}(\mathrm{\Sigma }_\alpha {}_{}{}^{}\mathrm{\Sigma }_{}^{\alpha })`$ characterizes the “magnitude” of the energy-momentum of the massive vector field. Using (15), we find that $$|\mathrm{\Sigma }|={}_{}{}^{}(\mathrm{\Sigma }_\alpha {}_{}{}^{}\mathrm{\Sigma }_{}^{\alpha })=\frac{1}{r^4g^4}\{(u^{}\frac{u}{r})^4+m^2u^2\frac{g^2}{f^2}(u^{}\frac{u}{r})^2+m^4u^4\frac{g^4}{f^4}\}.$$ (20) ## IV Preliminary analysis Before we start the study of the complete system it is instructive to recall two particular cases: namely, massless vector particle in curved spacetime and massive vector particle in Minkowski spacetime. For the massless vector particle $$m=0,$$ (21) and one finds, from (19), that $`g=g_0=const`$. Consequently, the vector field equation can now be easily integrated to give the usual Coulomb solution $$u=q=constA=\frac{q}{r}d\tau .$$ (22) \[Strictly speaking, the general solution reads $`u=q+\beta r`$, but one can put $`\beta =0`$ since it contributes only an exact form to $`A`$\]. Turning to (18), we immediately recover the well known Reissner-Nordström solution $$f^2=g_0^2\left(1\frac{2M}{r}+\kappa \frac{q^2}{2g_0^2r^2}\right),$$ (23) with the integration constant $`M`$ interpreted as the mass of the gravitating source and $`q/g_0`$ as its electric charge. On the other hand, for the Minkowski spacetime the metric functions are $`f=g=1`$ and the solution of the Proca field equation (13) yields the well known Yukawa potential $$u=qe^{mr}A=\frac{qe^{mr}}{r}d\tau .$$ (24) This shows that the vector field is practically zero at distances much greater than the typical length $`r_0=1/m`$. We expect that a spherically symmetric configuration of coupled Einstein and Proca fields will combine both of the typical features of the above limiting cases. Namely, for small values of mass $`m`$ there will be a large part of space inside the sphere of radius $`r_0=1/m`$ where the function $`u`$ is to a high degree of approximation constant. In this region the exact solution will naturally be approximated by (22) and the metric will assume the familiar Reissner-Nordström form (23). However, due to its massiveness the field $`A`$ will remain, also in curved spacetime, confined to a finite spatial volume, whereas for $`r\mathrm{}`$ one expects a fast decay $`u0`$ which leaves one with pure Schwarzschild metric. The following observation will be very useful in the discussion of exact solutions. Multiplying (13) by $`u/g`$ and using the Leibniz rule one finds that $$\left[\frac{u}{g}\left(u^{}\frac{u}{r}\right)\right]^{}=\frac{1}{g}\left\{\left(u^{}\frac{u}{r}\right)^2+m^2u^2\frac{g^2}{f^2}\right\},$$ (25) where the right-hand side is positive definite in a regular spacetime region. Identity (25) represents a particular case of the general relation $$db=(dA{}_{}{}^{}dA+m^2A{}_{}{}^{}A),$$ (26) where $`b:=A{}_{}{}^{}dA`$. The latter identity holds true for all solutions of the Proca field equation (2). ## V Dimensionless systems The general system is nonlinear and apparently cannot be integrated analytically. Consequently, we will present the results of numerical integration in the remainder of this paper. Before we begin the numerical analysis, we introduce a new dimensionless radial variable $$\rho :=\frac{r}{\sqrt{\kappa }},$$ (27) which allows us to rewrite the system (18), (19), and (13) in the form $`2\left(\rho {\displaystyle \frac{dF}{d\rho }}+FG\right)`$ $`=`$ $`K{\displaystyle \frac{G}{F}}u^2\left({\displaystyle \frac{du}{d\rho }}{\displaystyle \frac{u}{\rho }}\right)^2,`$ (28) $`\rho {\displaystyle \frac{dG}{d\rho }}`$ $`=`$ $`K{\displaystyle \frac{G^2}{F^2}}u^2,`$ (29) $`{\displaystyle \frac{d^2u}{d\rho ^2}}`$ $`=`$ $`K{\displaystyle \frac{G}{F}}u+{\displaystyle \frac{1}{2G}}{\displaystyle \frac{dG}{d\rho }}\left({\displaystyle \frac{du}{d\rho }}{\displaystyle \frac{u}{\rho }}\right).`$ (30) Here we have introduced a dimensionless constant $$K:=\kappa m^2,$$ (31) and defined the functions $$F:=f^2,G:=g^2.$$ (32) Evidently, the dimensionless radial coordinate measures distance from the origin in units of the Planck length ($`\kappa =\mathrm{}^2`$). At the same time, the parameter $`\sqrt{K}`$, being the ratio of the Planck length to the Compton length of the vector particle, characterizes the size of the domain where the influence of the Proca field on the spacetime geometry is significant. One can, alternatively, study a different dimensionless system after defining the scaled metric functions $$\stackrel{~}{F}:=F/K,\stackrel{~}{G}:=G/K,$$ (33) and introducing a new radial coordinate $$\xi :=\sqrt{K}\rho =mr=\frac{r}{r_0}.$$ (34) The system (28)-(30) then reads $`2\left(\xi {\displaystyle \frac{d\stackrel{~}{F}}{d\xi }}+\stackrel{~}{F}\stackrel{~}{G}\right)`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{G}}{\stackrel{~}{F}}}u^2\left({\displaystyle \frac{du}{d\xi }}{\displaystyle \frac{u}{\xi }}\right)^2,`$ (35) $`\xi {\displaystyle \frac{d\stackrel{~}{G}}{d\xi }}`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{G}^2}{\stackrel{~}{F}^2}}u^2,`$ (36) $`{\displaystyle \frac{d^2u}{d\xi ^2}}`$ $`=`$ $`{\displaystyle \frac{\stackrel{~}{G}}{\stackrel{~}{F}}}u+{\displaystyle \frac{1}{2\stackrel{~}{G}}}{\displaystyle \frac{d\stackrel{~}{G}}{d\xi }}\left({\displaystyle \frac{du}{d\xi }}{\displaystyle \frac{u}{\xi }}\right).`$ (37) In this form the equations no longer contain a free parameter (such as $`K`$) and the new dimensionless coordinate measures distance in units of the characteristic (“Compton wavelength”) scale $`r_0`$. It is convenient to use both dimensionless systems. The advantage of (35)-(37) lies in the absence of $`K`$, whereas the equations (28)-(30) are more transparent from the physical point of view when one considers limits of small and big mass $`m`$. ## VI Conditions at the origin and at infinity Before one can start the numerical integration, an appropriate set of initial conditions must be specified. Unfortunately, the solution cannot be represented by analytic power series expansion for $`u,F,G`$ at the origin in view of the apparent singularity at $`\rho =0`$. Instead, one can verify that for small values of $`\rho `$, irrespective of the value of $`K`$, there is an approximate solution of the form $`u`$ $``$ $`q+b\rho ,`$ (38) $`F`$ $``$ $`{\displaystyle \frac{1}{2}}{\displaystyle \frac{q^2}{\rho ^2}},`$ (39) $`G`$ $``$ $`{\displaystyle \frac{1}{\frac{1}{c}\frac{K}{q^2}\rho ^4}},`$ (40) where $`q,b,c`$ are parameters which determine the initial conditions in the neighborhood of the origin $`\rho =0`$. At infinity, $`\rho \mathrm{}`$, following the physical discussion in Sect. IV, we expect approximate behavior of the form $`u`$ $``$ $`u_0\mathrm{exp}(\sqrt{K}\rho ),`$ (41) $`F`$ $``$ $`g_0^2\left(1{\displaystyle \frac{2M}{\rho }}\right),`$ (42) $`G`$ $``$ $`g_0^2,`$ (43) where $`u_0,g_0`$ are constants. The condition (41) means that the massive vector field is nontrivial only inside a sphere of a finite radius $`1/\sqrt{K}`$ (“Yukawa-type” behavior). On the other hand, the conditions (42)-(43) specify purely Schwarzschild asymptotic metric. A more precise form of the limit (43) is easily obtained after substituting (41) and (42) into (28)-(30): $$Gg_0^2\left(1+Ku_0^2\mathrm{Ei}(2\sqrt{K}\rho )\right),$$ (44) where $`\mathrm{Ei}(x)=\underset{\mathrm{}}{\overset{x}{}}\frac{e^t}{t}𝑑t`$ is the integral exponential function. It is worthwhile to recall that asymptotically, for $`x\mathrm{}`$, one has $`\mathrm{Ei}(x)e^x/x`$. We will use the asymptotic conditions (41)-(43), (44) in the numerical analysis of the problem under consideration. ## VII Absence of solutions with horizons In this section we show that the spherically symmetric Einstein-Proca system does not admit asymptotically flat solutions with horizons. The absence of black holes for a massive vector field was first demonstrated by Bekenstein . Let us consider an arbitrary regular solution $`u(r)`$ which vanishes at two points $`r_1`$ and $`r_2>r_1`$: $`u(r_1)=u(r_2)=0`$. Then $`u(r)=0`$ for all $`r_1rr_2`$. Indeed, integrating the identity (25) from $`r_1`$ to $`r_2`$, one finds that $$\underset{r_1}{\overset{r_2}{}}\frac{1}{g}\left\{\left(u^{}\frac{u}{r}\right)^2+m^2u^2\frac{g^2}{f^2}\right\}𝑑r=\frac{u(r_2)}{g(r_2)}\left(u^{}(r_2)\frac{u(r_2)}{r_2}\right)\frac{u(r_1)}{g(r_1)}\left(u^{}(r_1)\frac{u(r_1)}{r_1}\right)=0.$$ (45) Since the integrand is positive definite, the vanishing of the integral leads to the above conclusion. Consequently, a nontrivial solution $`u(r)`$ which vanishes asymptotically at $`r_2=\mathrm{}`$ (thus satisfying the condition (41)) cannot have zeros at any finite $`r_1`$ (since then the solution would be trivial: $`u(r)=0`$ for $`rr_1`$). This leads to the absence of the black hole type solutions of the system (28)-(30). In order to see this, let us recall that a black hole necessarily possesses a horizon. Quite generally, on a spacetime manifold $``$ a horizon is defined as a hypersurface $`S:=\{x^i|\sigma (x^i)=0\}`$ such that: (i) the normal vector $`n_i:=_i\sigma `$ is null $$n_in^i|_S=0,$$ (46) and (ii) $`S`$ is not an essential singularity. The latter means that all the curvature invariants as well as the volume 4–form $`\eta `$ are nonsingular on the horizon. In particular, the regularity of $`\eta `$ follows from the condition (10) on the function $`g`$. For a spherically symmetric gravitational field configuration, horizon $`S`$ is evidently a sphere $`\sigma =r=r_h`$. Normal vector is then $`n_i=\delta _i^1`$. Substituting (8) into (46), one obtains $$\frac{f^2(r_h)}{g^2(r_h)}=\frac{F(r_h)}{G(r_h)}=0.$$ (47) Since $`G(r_h)`$ is finite in view of (10), we find that $`F`$ must vanish on the horizon $`S`$ $$F(r_h)=0.$$ (48) The last equation formally defines the position of a horizon in the general spherically symmetric spacetime (8). Now recall the second requirement: a hypersurface $`S`$ must be free of physical singularities in order to be a horizon. Clearly, the energy-momentum invariant scalar (20) is regular at $`r=r_h`$ if and only if $$u(r_h)=0.$$ (49) Furthermore, if (49) did not hold then the quadratic curvature invariants (obtained by using the Einstein field equations (3) in the definition of $`|\mathrm{\Sigma }|`$) would diverge at $`r_h`$ because of the last term in (20) and (48). Now we are in a position to conclude that there are no solutions with a horizon and a nontrivial massive vector field. Indeed, assume the contrary is true. Then outside a horizon $`S`$ the function $`u`$ is necessarily given by $`u(r)=0,r_hr\mathrm{}`$, because $`u`$ vanishes at infinity (41) and at the horizon (49). Consequently, outside $`S`$, the system (35)-(36) has the usual Schwarzschild solution $`G=1,F=1r_h/r`$. Integrating (35)-(37) from $`r_h`$ to $`0`$ with the initial conditions (48) and (49), we find $`u(r)=0`$ everywhere. Bekenstein’s original proof was based on the assumption that the three-form $`b=b^\alpha \eta _\alpha `$ defined in (26) is bounded on the horizon. It is easy to see that $`b^\alpha =\frac{u}{r^2fg}\left(u^{}\frac{u}{r}\right)\delta _{\widehat{1}}^\alpha `$ diverges on the horizon (48) unless $`u`$ vanishes. Thus, for a massive vector field, the form $`b`$ is not only bounded but, in fact, trivial on the horizon. ## VIII Numerical solutions After fixing the value of the parameter $`K`$ to the square of the ratio of the Planck length to the Compton wavelength of the vector field, one can start numerical integration at an arbitrarily small $`\rho `$ with the initial conditions defined by (38)-(40). One is free to choose any initial value for the “boson charge” function $`u(0)=q`$ ($`0`$, otherwise $`u`$ is trivial everywhere). Solutions with the correct asymptotic behavior (41)-(43) exist only for fixed values of the parameters $`b=u^{}(0),c=G(0)`$. Technically, the numerical integration can start at a point arbitrarily close to the origin for every chosen values of $`K`$ and $`q`$. In order to obtain the asymptotic behavior (41)-(43), a fine tuning of $`b`$ and $`c`$ is required which can be achieved similarly to the construction of the Bartnik-McKinnon solutions or of the Abrikosov-Nielsen-Olesen vortices (see, e.g. and references therein). Alternatively, one can start the numerical integration at a sufficiently large radius with the initial conditions taken from (41)-(43) for arbitrary values of $`K,M,u_0`$. As a cross-check, we have used both integration schemes. The resulting approximate solutions turned out to be completely consistent with each other. Particular solutions for various values of $`K`$, $`q`$ and $`M`$ are described in Tables I-III. The graphical form of the solutions is presented in Figures 1,2, and 3. In these figures, the numerical solutions are depicted for $`K=1`$ and $`M=0.1`$ (dotted lines), $`M=0.5`$ and $`M=1.5`$ (broken lines), and $`M=2`$ (solid lines). In all cases we put $`g_0=1`$ which is always possible to achieve by the redefinition of the time coordinate $`\tau g_0\tau `$. As one can see, the relation between $`K`$ (formal rest mass of the vector field) and $`M`$ (asymptotic total mass of the solution) plays a decisive role. At the same time, the value of the boson charge $`q`$ at the origin is also important. In agreement with the results of Bekenstein et al and with the preliminary analysis of Sect. VII, all the numerical solutions obtained by us are without horizons. They possess a true physical singularity at the origin which provides us with an example of a naked spacetime singularity. Stability of these solutions against small perturbations will be studied separately. Recalling that the effective Proca field emerges naturally in the general metric-affine models, we thus conclude that the presence of the post-Riemannian geometric objects prevents, in general, a formation of a black hole in MAG theory. Only in the special case when the MAG coupling constants are such that the effective mass vanishes, $`m^2=0`$, the black holes can be formed . ###### Acknowledgements. This work was supported by a grant from the University of Newcastle which made it possible for YNO to visit Australia. We are grateful to Robin Tucker and Tekin Dereli for useful comments, and to Marc Toussaint for a discussion of the results obtained in this paper.
warning/0004/nucl-th0004028.html
ar5iv
text
# Parity violation in reactions with B- and Li-nuclei using polarized thermal neutrons ## Abstract Parity violation in the reactions <sup>6</sup>Li(n,$`\alpha `$)<sup>3</sup>H, <sup>10</sup>B(n,$`\alpha `$)<sup>7</sup>Li and <sup>10</sup>B(n,$`\alpha `$)<sup>7</sup>Li$`{}_{}{}^{}`$ $`\gamma +^7`$Li(g.s) is discussed. It appears that the effects (neutral currents contribution in particular) are measurable due to the large values of the neutron cross sections. This holds promise for the theoretical extraction of the constants of the weak NN potential because of the relative simplicity of the investigated systems. <sup>1</sup>Petersburg Nuclear Physics Institute, Gatchina 188350, Russia <sup>2</sup>Joint Institute for Nuclear Research, Dubna 141980, Russia <sup>3</sup>Moscow State University, Moscow 119899, Russia The fact that the list of measured parity violation (PV) processes in nuclei accessible for extracting the weak NN interaction constants is short sends investigators in search for new regions. Moreover, the current data on PV in nuclei are, for the most part, in the region of heavy or middle (A$`>20`$) atomic weight nuclei. The possibility of extracting the constants of the parity violating NN Hamiltonian using such data is the subject of sharp controversy. Almost the same is true for most experimental results in the light nuclei region (A$``$ 14-20). Only a few measured examples are suitable here for extracting the constants of the P-odd part of the NN-potential. The theoretical situation seems to be better for few-nucleon systems (A$`4`$) and lightest nuclei (4$`<`$ A$`10`$). However, in these cases another problem arises. This problem is that almost all measured examples of nuclear P-odd processes for A$`>`$4 are connected with the existence of parity mixing doublets (PMD) . At the same time, the spectra of the majority of lightest nuclei do not contain such doublets. As a result, P-odd effects in these nuclei are not enhanced. In addition, the theory of P-odd nuclear effects in the absence of PMD is developed for few special cases only. The idea to use the discussed reactions to measure the PV effect together with some estimates are reported in Ref. . Requirements for the necessary experimental precision turn out to be very strict. So, the sole subject of intensive investigations in the 4$`<`$A$`10`$ region was the parity forbidden $`\alpha `$-decay of the 3.56 MeV level in the <sup>6</sup>Li nucleus. However, PMD is also absent in it and the effect turned out to be extremely small according to various calculations . In the present work, we demonstrate the results of both experimental and theoretical efforts to do the investigation of parity violation in the region of lightest nuclei. Particular attention has been given to the discussion of P-odd effect in $`\gamma `$-transition in <sup>7</sup>Li-nucleus where there is a definite possibility to extract PV constants. A joint PNPI-JINR group carried out experimental research of the P-odd asymmetries $`W_{PV}^\alpha (\theta )1+a_{PV}^\alpha (𝐬_n𝐤_\alpha )`$ and $`W_{PV}^\gamma (\theta )1+a_{PV}^\gamma (𝐬_n𝐤_\gamma )`$ in the reactions <sup>10</sup>B(n,$`\alpha _{0,1}`$)<sup>7</sup>Li and <sup>6</sup>Li(n,$`\alpha `$)<sup>3</sup>H. The measurements were performed in the vertical beam of polarized thermal neutrons from the water-water reactor (WWR-M) in PNPI. The neutron beam had dimensions of 100x30 mm<sup>2</sup> with an intensity of $`310^{10}`$ neutrons/s. The neutron polarization averaged over the beam cross section amounted to P=0.8. In all discussed reactions, we observed comparatively large left-right asymmetries of the form $`W_{LR}1+a_{LR}(𝐬_n[𝐤_\alpha 𝐤_n]),(a_{LR}10^4)`$, where s<sub>n</sub> , k<sub>n</sub> and k<sub>α</sub> are, respectively, the neutron spin, the neutron momentum, and the momentum of the detected charged particle. For this reason, the proportional chamber was designed to ensure the fulfillment of the geometric condition s$`{}_{n}{}^{}`$ k$`{}_{\alpha }{}^{}`$ k<sub>n</sub>. The detector used in this study consisted of 24 double ionization chambers positioned in the cylindrical duralumin vessel. The possible contribution of the left-right asymmetry to the P-odd effect does not exceed $`10^8`$. The relative amplitude of the fluctuations of the incident neutron flux due to temporal changes in the working parameters of the reactor is about $`10^3`$ to $`10^2`$. The actual signal variations caused by the nuclear reaction are at least 30 times smaller ($`10^4`$ to $`10^5`$). Therefore, it is necessary to compensate the reactor-power fluctuations. In the forward and backward proportional chambers, the signs of the sought effects are opposite to each other. This fact is used to compensate the reactor-power fluctuations. In it is possible to read more about the scheme of measuring of PV effects for charged particle channels. An analogous scheme for measuring the $`\gamma `$-quanta asymmetry differs by the detector set. Na(Tl) crystals 150 mm long by 150 mm wide with photoelectron multipliers served as detectors. They are housed on both sides of the sample 70 mm from the neutron beam axis. The integral current method is used. A natural boron target is placed in a <sup>6</sup>LiF box 10 mm thick. The current signal is taken from the seventh dynode where the linear dependence between the values of the current and the $`\gamma `$-quanta flux occurs. Special care is also taken to suppress the magnetic field induced noise. Low energy $`\gamma `$-quanta, E$`{}_{\gamma }{}^{}500`$ keV, result in an abrupt (about an order of magnitude) increase of the inherent noise of the detectors and as a result, a increase of the instrumental error of the effect under discussion 5 to 10 times that of the statistical error. We will discuss in short the peculiarities and results of each experiment. Since the specific energy loss of the $`\alpha `$-particle in the reaction <sup>6</sup>Li(n,$`\alpha )^3`$H is higher than of the triton, no precise adjustment of the gas pressure is required. We use the chamber in the ionization regime without gas enhancement in the linear range. The total experimental time was about 40 days. Taking the background measurements into account the following experimental result for $`a_{PV}^\alpha `$ is obtained : $`a_{PV}^\alpha =(6.44\pm 5.50)10^8`$. The reaction <sup>10</sup>B(n,$`\alpha _{0,1})^7`$Li was studied in the proportional chamber with an insensitive gas region . Two 8.5-day series of measurements were carried out in the ionization regime: one at an argon pressure of 0.3 atm (the effect was measured for the mixture of two lines) and the other at a pressure of 0.78 atm (where according to calculations, only the higher energy $`\alpha `$-group was detected). The resulting experimental values of the P-odd asymmetry are as follows : $`a_{PV}^{\alpha _1}=(2.5\pm 1.6)10^7;a_{PV}^{\alpha _0}=(3.4\pm 6.7)10^7`$. A relatively novel subject is the measurement of the P-odd asymmetry $`W_{PV}^\gamma (\theta )1+a_{PV}^\gamma (𝐬_n𝐤_\gamma )`$ of the gamma transition from the first excited state of <sup>7</sup>Li$`(1/2^{},0.478`$ MeV), populated according to the scheme: <sup>10</sup>B(n,$`\alpha _1)^7`$Li$`{}_{}{}^{}(1/2^{},0.478`$ MeV) $``$ M1(E1) $`^7`$Li$`(3/2^{}`$, g.s.). In this case the $`(n,\alpha _1)`$ reaction is used for the creation of a polarized source of $`\gamma `$-quanta. Taking into account the above mentioned instrumental noises only tentative measurements were carried out. Two two-week series resulted in the value $`a_{PV}^\gamma =(6.8\pm 3.7)10^7`$, which is reasonable considering the upper limit of the effect under investigation because of discussed instrumental errors. Certainly, the accuracy of this result must be improved for the extraction of even the upper limits of the PV constants. However, the prospect of the increasing of the accuracy of the experiment seems to be good. Further instrumental improvements are the replacement of the photoelectron multipliers by semiconductor silicon photodiodes. The latter were explored when the circular polarization of $`npd\gamma `$ was measured . The application of them in this experiment permitted the investigators to compensate the fluctuations of the reactor almost completely. In the discussed case, the Hamamatsu photodiodes with the ”dark” current of about $`10^9`$A at room temperature seems to be optimal. In this case, the contribution of the instrumental noise to the total signal is expected to be negligible. Under these conditions it is possible to increase the estimated precision of the measurement by one order of magnitude. Let us directly consider the theoretical aspects of discussed problem. In most cases spectra of lightest nuclei do not contain close PMD and even far distant ones. So, the P-odd properties of these levels are associated with the admixture of continuum wave functions of nucleon and cluster channels or the wave functions of broad resonances being, in fact, the pieces of the same continuums. To develop the approach, allowing one to extract the PV constants in this case (notably for cluster-cluster channels), is an original theoretical problem which is a rather general and challenging one. This problem has been partially solved for the first time in Ref. , where the width of the P-odd decay of the $`0^+`$; T=1 level of <sup>6</sup>Li to the $`\alpha +d`$ continuum is calculated. In contrast to the discussed case this case is a typical ”one narrow resonance – one channel” problem , where the sole wave function of the continuum at the resonance energy should be known to calculate the decay width. The discussed processes are essentially different. The first two ones are typical nuclear reactions characterized not only by resonances but direct mechanism also. Naturally off-shell effects are essential here and consequently complete sets of continuum wave functions are required to describe PV in these cases. Competition between resonance and direct mechanisms mentioned earlier makes the problem of extracting of PV constants very complicated and we do not discuss it in the present work. The third PV process is not a nuclear reaction but it is a $`\gamma `$-transition and consequently, the theoretical interpretation of it is not so complicated. Therefore, it is possible to develop a rather well justified approach to the process. Let us discuss the P-odd asymmetry of the angular distribution of $`\gamma `$-quanta emitted by the polarized sample of first excited state of <sup>7</sup>Li ($`1/2^{}`$, E$`{}_{}{}^{}=0.478`$ MeV) as an example of rather general approach. The spectrum of the <sup>7</sup>Li nucleus (see Ref. ) does not contain any $`1/2^+`$ or $`3/2^+`$ discrete levels or narrow resonances (the parity mixing partners of the initial $`1/2^{}`$ and final $`3/2^{}`$ states). The lowest continuum of this 7N-system is an $`\alpha +t`$ continuum beginning at E$`{}_{0}{}^{\alpha +t}=2.47`$ MeV. The thresholds of other channels lie much higher namely E$`{}_{0}{}^{{}_{}{}^{6}Li+n}=7.25`$ MeV, E$`{}_{0}{}^{{}_{}{}^{6}He+p}=9.98`$ MeV. Such situation is typical for the discussed nuclear area and so cluster channels are usually the most important. The existence of broad ($`\mathrm{\Gamma }3`$ MeV) resonances in the neutron channel is the subject of controversy. Anyway, these resonances, even if they really exist, are the pieces of the related continuums possessing a slightly higher spectral density. In addition they are removed far away from the levels of our interest. So, it is rather obvious that the most essential contribution to the effect under discussion appears due to the P-odd coupling of the discrete levels $`1/2^{}`$ and $`3/2^{}`$ with the proper $`\alpha +t`$ continuum states of positive parity. So, the expression for the P-odd asymmetry has the form: $`a_{PV}^\gamma =a_{PV}^\gamma (excited)+a_{PV}^\gamma (ground)`$, which contains the contributions of the opposite parity components to both the initial and final states. The formulas for both contributions are in fact the same. In comparison with the PMD case the formalism of PV in the continuum contains the integration over an infinite energy range of the $`\alpha +t`$ continuum. For example, the expression for the wave function of the excited state of the <sup>7</sup>Li nucleus $`1/2^{}`$ has the form: $$\mathrm{\Psi }_{(1/2)}=\mathrm{\Psi }_{(1/2)^{}}+\underset{E_0^{\alpha +t}}{\overset{\mathrm{}}{}}\frac{\mathrm{\Psi }_{\alpha +t(1/2)^+}^E|V^{PV}|\mathrm{\Psi }_{(1/2)^{}}}{E_{(1/2)^{}}E}|\mathrm{\Psi }_{\alpha +t(1/2)^+}^E𝑑E.$$ (1) Here, $`V^{PV}`$ is the PV potential, $`\mathrm{\Psi }_{\alpha +t(1/2)^+}^E`$ is the wave function of the related positive parity continuum. The contribution of this state to the discussed effect can be written as: $$a_{PV}^\gamma (ex.)=A_F^J\underset{E_0^{\alpha t}}{\overset{\mathrm{}}{}}\frac{\mathrm{\Psi }_{(3/2)^{}}|E1|\mathrm{\Psi }_{\alpha +t(1/2)^+}^E}{\mathrm{\Psi }_{(3/2)^{}}|M1|\mathrm{\Psi }_{(1/2)^{}}}\frac{\mathrm{\Psi }_{\alpha +t(1/2)^+}^E|V^{PV}|\mathrm{\Psi }_{(1/2)^{}}}{E_{(1/2)^{}}E}dE,$$ (2) where the amplitudes of the regular M1 and irregular E1 $`\gamma `$-transitions and spin-orbital factor of reaction $`A_F^J`$ are also the factors in the expression for the PV effect. The next step is the key to the problem. The matter is that it is possible to extract the PV constants using a microscopic two-nucleon expression for the potential $`V^{PV}`$ in Exp.(2) only. So, it is necessary to write the microscopic 7N wave functions not only for bound states of <sup>7</sup>Li but also for the $`\alpha +t`$ continuum (or arbitrarily other nucleon-nucleus or cluster-nucleus continuum for the general consideration) in the matrix element of this potential. Dealing with low-laying bound states of lightest nuclei creates no serious problems. At the same time, to write the microscopic wave function of the $`\alpha +t`$ channel in the form suitable to continue operating with the microscopic two-nucleon operator $`V^{PV}`$, we use the method of the shell-model expansion of the multicluster wave function . As the guide for it the orthogonality condition model (OCM) is used. This model is the simplified version of resonating group model (RGM). The wave function of OCM has the form: $$\mathrm{\Psi }_{\alpha +t(J)^+}^E=\widehat{A}\left\{\mathrm{\Psi }_\alpha \mathrm{\Psi }_t\widehat{N}^{\frac{1}{2}}\mathrm{\Phi }_l^E(𝝆)\right\}.$$ (3) Here, $`\mathrm{\Psi }_\alpha \mathrm{\Psi }_t`$ are the internal wave functions of the alpha-particle and the triton, respectively, $`\mathrm{\Phi }_l^E(𝝆)`$ is the function of relative motion in the $`\alpha +t`$ continuum, $`\widehat{A}`$ is the antisymmetrizer and the operator $`\widehat{N}`$ is the exchange kernel of RGM. As it is shown in one can expand the continuum wave function $`\mathrm{\Phi }_l^E(𝝆)`$ in terms of the oscillator wave functions $`\mathrm{\Phi }_{nl}(𝝆)`$. Due to this, for the two-body continuum wave functions one can obtain: $$\mathrm{\Psi }_{\alpha +t(J)^+}^E=\underset{n}{}C_n(E)\epsilon _n^{1/2}\widehat{A}\{\mathrm{\Psi }_\alpha \mathrm{\Psi }_t\mathrm{\Phi }_{nl}(𝝆)\},$$ (4) where $`C_n(E)`$ are the coefficients of the oscillator expansion of the function $`\mathrm{\Phi }_l^E(𝝆)`$ and $`\epsilon _n`$ are the eigenvalues of the exchange kernel. Further transformation of the continuum wave functions $`\mathrm{\Psi }_{\alpha +t(J)^+}^E`$ is the conversion from the seven-nucleon two-cluster oscillator wave functions contained in the right-hand side of Exp. (4) to the translationally-invariant shell model (TISM) wave functions. It is important to note that the first term of Exp. (4) which does not vanish by the antisymmetrizer $`(n=4)`$ coincides with the respective TISM wave function. But it is this function that is the sole component, contributing to the matrix element of the E1 transition because the functions of the low lying levels of <sup>7</sup>Li are characterized by the principal quantum number $`n=3`$ and the operator of the E1 transition changes that number by unity. Inserting the expressions for the wave functions into the formula (2) and taking into account the above mentioned circumstances we obtain: $$a_{PV}^\gamma (ex.)=A_F^J\underset{n}{}\frac{\mathrm{\Psi }_{(3/2)^{}}|E1|\mathrm{\Psi }_{4,(1/2)^+}\mathrm{\Psi }_{n,(1/2)^+}|V^{PV}|\mathrm{\Psi }_{(1/2)^{}}}{\mathrm{\Psi }_{(3/2)^{}}|M1|\mathrm{\Psi }_{(1/2)^{}}}\underset{E_0^{\alpha t}}{\overset{\mathrm{}}{}}\frac{C_n^{}(E)C_4(E)}{E_{(1/2)^{}}E}dE.$$ (5) Here we do not present the cumbersome transformation of the wave function $`\mathrm{\Psi }_{n,(1/2)^+}=\epsilon _n^{1/2}\widehat{A}\{\mathrm{\Psi }_\alpha \mathrm{\Psi }_t\mathrm{\Phi }_{nl}(𝝆)\}`$ to the linear combination of TISM wave functions for brevity. The wave function $`\mathrm{\Phi }_l^E(𝝆)`$ is the solution of the Schrodinger equation containing the nuclear and Coulomb terms. The Woods-Saxon form of the cluster-cluster interaction with the parameters $`V_0=87`$ MeV, $`\rho _0=1.8`$ fm and $`a=0.7`$ fm proposed in to fit the phase shifts of $`\alpha t`$ scattering is used. So, the scheme which is deduced is in fact an universal theoretical approach to the solution of the problem of the P-odd mixing of a discrete level and arbitrary two body (nucleon-nucleus or cluster-nucleus) continuum. For the particular case discussed, our main goal was to show that the scale of the PV effect here is not small in comparison with the experimental precision. In the light of this the precision of the used model seems to be quite sufficient. The final formula for the P-odd asymmetry of the transition under investigation, written in terms of PV constants, takes the form: $$a_{PV}^\gamma =0.078h_\pi +0.028h_\rho ^0+0.010h_\rho ^1+0.015h_\omega ^0+0.014h_\omega ^1.$$ (6) It takes the value $`a_{PV}^\gamma =4.1710^8`$ if one uses the Dubovik-Zenkin version of PV-constants and $`a_{PV}^\gamma =7.2410^8`$ if one uses DDH ”best values” . The contribution of neutral currents to the discussed process is equal to: $`a_{NC}^\gamma =1.2810^8`$ (30.6% ) for PV-constants from and $`a_{NC}^\gamma =3.7610^8`$ (51.9%) for ”best values” from . So predicted values of the effect turn out to be measurable by the discussed method . To demonstrate qualitatively the scale of continuum PV effects it is convenient to define the value of $``$ E<sup>eff</sup> as the equivalent energy distance between a level and its nonexisting PV partner producing the effect of continuum mixing. The use of $`n=4`$ term matrix elements brings the results: $``$ E$`{}_{}{}^{eff}=`$17.8 MeV and 21.2 MeV for excited $`1/2^{}`$ and ground $`3/2^{}`$ states, respectively. The effect of continuum mixing is equal to that of a discrete level spaced an approximately $`\mathrm{}\omega `$ energy distance apart from the top of the potential barrier although for convergency of the results it is necessary to take into account the influence of wide range of continuum wave functions up to the energy value close to 300 MeV. So the effect of the continuum with a rather close threshold and a low barrier has the scale similar to that of a nonenhanced process. The effect of continuum turns out to be tolerant to manifold variations of the input namely the model of cluster-cluster interaction ( if realistic versions of it are used), cluster wave functions etc. This property appears, of course, if any narrow resonance is absent in a continuum. In this case the integral with respect to E in (2) is a smooth function of the upper limit. That’s why the related values of $``$ E<sup>eff</sup> are notably independent on the model of interaction. In this respect, the effect under discussion is even more preferable for extracting the PV constants than the PV effect caused by PMD being more tolerant to the nuclear peculiarities. Naturally that is true if PMD are absent in the spectra of clusters. The versatility of the method of shell model expansion, used in the present work, allows one to incorporate rather easily more precise models of clusters and cluster-cluster interaction. For example it is possible to take into account another channels, to use one- or multi-channel RGM in place of OCM. The extension of the approach to the PV effect in direct reactions is also not hopeless. So the discussed measurements deserve serious theoretical support. In conclusion let us note that the present work demonstrates the possibilities of modern experiments to measure P-odd effects in above mentioned nonenhanced nuclear processes i.e. to achieve the precision of the order of several $`110^8`$ using the peculiarities of lightest nuclei. Good prospects for theory in the discussed area are also demonstrated. The result of calculation of the P-odd asymmetry of the 478-keV $`\gamma `$-transition in the polarized <sup>7</sup>Li nucleus demonstrates that the above mentioned precision is satisfactory for the extracting of the constants of the PV NN potential. And what is more, the contribution of neutral currents is large enough to be measured. It is shown that the existence of PMD in the discussed systems is not a strict condition for a successful investigation of PV effects. So, new prospects for the study of PV constants in the reported area of nuclear masses, in general, and in the discussed reactions, in particular, are seen. The $`\gamma `$-transition between two lowest levels in a polarized <sup>7</sup>Li nucleus is believed to be the most promising for these purposes. Finally let us conclude that extended experimental investigations of the last-named process using a high-intensity thermal neutron source seems us to be very desirable, so the present work may be considered as a preliminary proposal for them. The work supported by RFBR grants $`N^{\underset{¯}{o}}`$ 00-02-16707 (experimental part) and $`N^{\underset{¯}{o}}`$ 00-02-16683 (theoretical part).
warning/0004/nlin0004034.html
ar5iv
text
# New Nonlocal Charges in SUSY Integrable Models ## 1 Introduction: Integrable models appear naturally in the study of strings in the matrix model approach. Thus, while the KdV hierarchy is obtained in the double scaling limit of the one matrix model , the supersymmetric matrix models lead to a particular supersymmetric version of the KdV hierarchy known as the $`N=1`$ supersymmetric KdV-B hierarchy . In simple terms, if $`u`$ denotes the dynamical variable of the KdV equation $$u_t=u_{xxx}+6uu_x,$$ where the subscripts represent differentiation with respect to the corresponding variables, then, the $`N=1`$ supersymmetric KdV-B hierarchy is given by $$\mathrm{\Phi }_t=\mathrm{\Phi }_{xxx}+3(D(D\mathrm{\Phi })^2).$$ Here, $`\mathrm{\Phi }(x,\theta )=\psi (x)+\theta u(x)`$ represents the dynamical variable which is a $`N=1`$ fermionic superfield with $`\theta `$ denoting the Grassmann coordinate and $$D=\frac{}{\theta }+\theta \frac{}{x}.$$ There are, of course, other supersymmetrizations of the KdV hierarchy that are integrable , but it is this particular supersymmetrization that manifests in the study of string theories. Therefore, in this letter, we undertake a systematic study of the properties of such a supersymmetrization. In particular, we show, in section 2, that this particular method of supersymmetrization can be applied to any integrable model, although the original study involved the bosonic KdV hierarchy. In section 3, we bring out some general properties of such models, such as the Hamiltonians, Hamiltonian structures, recursion operators etc. In these models, there arise local conserved charges which have opposite Grassmann parity relative to the Hamiltonians of the system. The origin of such “exotic” charges is explained in section 4, where we identify that such local charges belong to the hierarchy of an infinite set of non-local charges. In fact, we show that, in such models, there exist, at least, two infinite sets of non-local charges and may be more. Explicitly, in the $`N=2`$ supersymmetric KdV-B hierarchy, we show that there exist three infinite sets of non-local charges and present a method for constructing them. In section 5, we present briefly an alternate description for such a supersymmetrization which allows the construction of $`B`$-extensions of systems such as the NLS and the AKNS hierarchies. A brief conclusion is presented in section 6. We used the symbolic computer language Reduce and the special package in all calculations presented in this letter. ## 2 Model: Let us consider a general integrable model of the form $$\varphi _t=(A[\varphi ])_x,$$ (1) where the subscripts refer to differentiation with respect to the corresponding variables. Here, $`\varphi `$ is a general dynamical variable. It can be a purely bosonic function of $`x`$ alone, in which case, the equation will represent the dynamics of a bosonic integrable system ($`N=0`$ supersymmetry). Alternately, $`\varphi `$ may represent a superfield (bosonic or fermionic) depending on $`x`$ as well as $`N`$ fermionic coordinates, $`\theta _i,i=1,2,\mathrm{},N`$, in which case, the dynamical equation will describe an integrable model with $`N`$-extended supersymmetry. Let us denote the covariant derivatives with respect to the $`N`$ fermionic coordinates by $$D_i=\frac{}{\theta _i}+\theta _i\frac{}{x},i=1,2,\mathrm{},N,$$ (2) which satisfy $$[D_i,D_j]_+=\delta _{ij}\frac{}{x}=\delta _{ij}.$$ We can now introduce a new superfield, ($`\varphi `$ and $`\stackrel{~}{\varphi }`$ are superfields depending on the original $`N`$ fermionic coordinates) $$G=G(x,\theta _1,\mathrm{},\theta _{N+1})=\stackrel{~}{\varphi }+\theta _{N+1}\varphi ,$$ (3) which depends on one extra fermionic coordinate and has opposite Grassmann parity relative to $`\varphi `$ (the original dynamical variable), and define the dynamical equation $$(D_{N+1}G)_t=(A[(D_{N+1}G)])_x,$$ (4) which would represent an integrable system with $`(N+1)`$-extended supersymmetry. This, therefore, describes the generalization of Beckers’ extension to (extended) supersymmetric models. (Basically, the original $`\varphi `$ equation remains unchanged under this extension since $`(D_{N+1}G)|_{\theta _{N+1}=0}=\varphi `$.) Thus, for example, with $`\varphi =u(x)`$ and $`A[u]=u_{xx}+3u^2`$, we have the bosonic KdV equation ($`N=0`$ supersymmetry) while, with $`G=\mathrm{\Phi }(x,\theta )`$, where $`\mathrm{\Phi }`$ is a fermionic superfield, the equation $`(D\mathrm{\Phi })_t`$ $`=`$ $`\left((D\mathrm{\Phi }_{xx})+3(D\mathrm{\Phi })^2\right)_x,`$ $`\mathrm{or},\mathrm{\Phi }_t`$ $`=`$ $`\mathrm{\Phi }_{xxx}+3D\left((D\mathrm{\Phi })^2\right),`$ (5) represents the $`N=1`$ supersymmetric KdV-B equation . Similarly, for $`\mathrm{\Phi }(x,\theta _1)`$ a fermionic superfield and $`A[\mathrm{\Phi }]=(\mathrm{\Phi }_{xx}+3\mathrm{\Phi }(D_1\mathrm{\Phi }))`$, the dynamical equation represents the $`N=1`$ supersymmetric KdV equation , while, with $`G(x,\theta _1,\theta _2)`$ a bosonic superfield, the equation $`(D_2G)_t`$ $`=`$ $`\left((D_2G_{xx})+3(D_2G)(D_1D_2G)\right)_x,`$ $`\mathrm{or},G_t`$ $`=`$ $`G_{xxx}3D_2\left((D_2G)(D_1D_2G)\right),`$ (6) would give rise to an $`N=2`$ extended supersymmetric KdV equation of the $`B`$-type. Similarly, if $`\varphi `$ represents an $`N=2`$ superfield and the dynamical equation gives the $`N=2`$ supersymmetric KdV equation , then, the $`G`$ equation would correspond to the $`N=3`$ supersymetric KdV-B equation and so on. While this procedure is quite general, in this letter, we would study the specific model in eq. (6) and bring out properties of this model which are nonetheless common to all such models. We also note here that, as described above, this extension, when applied twice to a given equation, would seem to lead to a nonlocal dynamical equation and, therefore, is not useful. Similarly, if the right hand side of eq. (1) is not a total space derivative, this method will also appear to fail. We would come back to this point in section 5, where we would describe an alternate, but equivalent generalization, which can be applied to any given equation and as many times, without introducing non-locality. ## 3 Hamiltonians and Hamiltonian Structures: Let us next look at the general model of eq. (1). We note that if $$H_n^{(N)}=𝑑x𝑑\theta _1\mathrm{}𝑑\theta _Nh_n^{(N)}[\varphi ],n=1,2,\mathrm{},$$ (7) represent the Hamiltonians of the original model, then, $$H_n^{(N+1)}=𝑑x𝑑\theta _1\mathrm{}𝑑\theta _{N+1}h_n^{(N)}[(D_{N+1}G)],n=1,2,\mathrm{},$$ (8) would correspond to the Hamiltonians of the extended $`B`$-model . These are conserved local quantities which would be invariant under the extended supersymmetry and we note that, since the integration, in the second case, is over an additional fermionic variable relative to the definition of the original charges, the Hamiltonians of the new system would have an opposite Grassmann parity compared to those of the original system. It is also not hard to see that the Hamiltonian densities can be written as $$h_n^{(N)}[(D_{N+1}G)]=h_n^{(N)}[\varphi ]+\theta _{N+1}\stackrel{~}{h}_n^{(N+1)}[\varphi ,\stackrel{~}{\varphi }],$$ (9) so that each of the two parts of the Hamiltonians, namely, the $`\theta _{N+1}`$ independent term as well as the linear term in $`\theta _{N+1}`$, will be independently conserved. However, the $`\theta _{N+1}`$ independent term in the density would give a conserved charge (when integrated over appropriate coordinates) which is invariant only under the lower, $`N`$-extended supersymmetry. The Hamiltonian structures of the two systems are also related in a simple manner. Suppose $`𝒟^{(N)}[\varphi ]`$ represents the Hamiltonian structure of the original system so that we can write $$\varphi _t=𝒟^{(N)}[\varphi ]\frac{\delta H_n^{(N)}[\varphi ]}{\delta \varphi }.$$ (10) Then, it follows that, we can write $`(D_{N+1}G)_t`$ $`=`$ $`𝒟^{(N)}[(D_{N+1}G)]{\displaystyle \frac{\delta H_n^{(N+1)}}{\delta (D_{N+1}G)}},`$ $`\mathrm{or},G_t`$ $`=`$ $`𝒟^{(N+1)}[G]{\displaystyle \frac{\delta H_n^{(N+1)}}{\delta G}},`$ (11) where $$𝒟^{(N+1)}[G]=D_{N+1}^1𝒟[(D_{N+1}G)]D_{N+1}^1.$$ (12) We note here that the new Hamiltonian structure would have an opposite Grassmann parity from the old one, simply because of the delta functions involving fermionic coordinates. (Such Hamiltonian structures are known as anti-brackets or Buttin brackets .) This is consistent with the change of the Grassmann parity for the Hamiltonians that we have already noted. The recursion operators for the two systems are similarly related and, without going into details, we simply note here that $$R^{(N+1)}[G]=D_{N+1}^1R^{(N)}[(D_{N+1}G)]D_{N+1}.$$ (13) The Lax description for the two systems are also simply related. For example, we know that the Lax description for the KdV hierarchy is given in terms of the Lax operator of the form $$L=^2+u,$$ where $``$ represents $`\frac{}{x}`$. It follows, then, that the Lax operator $$L=^2+(D\mathrm{\Phi }),$$ (14) would describe the $`N=1`$ supersymmetric KdV-B hierarchy through the same normal Lax representation, $$\frac{L}{t}=[L,(L^{3/2})_+]$$ . Similarly, the $`N=1`$ supersymmetric KdV equation can be described either by a standard representation with the Lax operator $$L=^2+D_1\mathrm{\Phi },$$ or by a Lax operator $$L=+D_1^1\mathrm{\Phi },$$ with the non-standard Lax representation $$\frac{L}{t}=[L,(L^3)_1].$$ Correspondingly, the $`N=2`$ supersymmetric $`B`$-extension of this system can also have a standard as well as a non-standard representation through the Lax operators $`L^{\mathrm{std}}`$ $`=`$ $`^2+D_1(D_2G),`$ $`L^{\mathrm{nstd}}`$ $`=`$ $`+D_1^1(D_2G).`$ (15) As opposed to these Lax operators it is also possible to define the Lax operator on the $`N=2`$ superspace. Indeed the method of supercomplexification provides a much more general procedure for obtaining the $`B`$-extensions and applied to this model, it provides the Lax operator $$L^{\mathrm{sc}}=+D_1^1(D_2G)D_1^1G_xD_2^1,$$ which would describe the system with a non-standard Lax representation. The conserved Hamiltonians of the system can be obtained from the super residues of any of these three Lax operators. Thus, for example, the Hamiltonians of the system can be written in terms of the non-standard Lax operator as $$H_n=𝑑x𝑑\theta _1𝑑\theta _2sRes(L^{\mathrm{nstd}})^{2n1}=𝑑x𝑑\theta _1𝑑\theta _2h_n,$$ (16) and the first two nontrivial charges of the series have the forms $`H_3`$ $`=`$ $`{\displaystyle 𝑑x𝑑\theta _1𝑑\theta _2G(D_1G_x)},`$ $`H_5`$ $`=`$ $`{\displaystyle 𝑑x𝑑\theta _1𝑑\theta _2\left[G(D_1G_{xxx})+4G(D_1G_x)(D_1D_2G)\right]}.`$ It is worth noting here that the $`N=2`$ supersymmetric $`B`$-extension (eq. (6)) has yet another Lax representation . Namely, consider the Lax operator $$L=D_1+^1D_2GGD_2^1.$$ (17) Then, it is straight forward to check that the non-standard Lax equation $$\frac{L}{t}=[L,(L^6)_1],$$ (18) gives the $`N=2`$ equation of eq. (6). However, this Lax operator, surprisingly, does not yield any of the conserved charges of the system. This is indeed a puzzling feature which deserves further study. ## 4 Non-local Charges Let us now concentrate on the $`N=2`$ model of the previous section (eq. (6)) for concreteness, although the features we are going to discuss are quite general. We note that although the Hamiltonians for this system are fermionic, as we have discussed, there are also the following bosonic charges which can be explicitly checked to be conserved, namely, $`\stackrel{~}{H}_1`$ $`=`$ $`{\displaystyle 𝑑x𝑑\theta _1𝑑\theta _2G},`$ $`\stackrel{~}{H}_2`$ $`=`$ $`{\displaystyle 𝑑x𝑑\theta 𝑑\theta _2G^2},`$ $`\stackrel{~}{H}_3`$ $`=`$ $`{\displaystyle 𝑑x𝑑\theta _1𝑑\theta _2\left(\frac{1}{3}G^3G(D_1D_2G)\right)}.`$ (19) At first sight, the existence of such charges with opposite Grassmann parity would seem surprising and, in fact, the existence of such a charge was already noted earlier in connection with the $`N=1`$ supersymmetric KdV-B system and was termed “exotic” . In what follows, we would try to clarify the origin of such charges and show that, in such systems, there are, at least, two infinite sets of conserved non-local charges (and may be more) and that these “exotic” charges are part of an infinite hierarchy of non-local charges. To begin with, let us recall that supersymmetric integrable systems, in general, possess non-local charges . However, in the study of dispersionless limits of $`B`$-extended models, it was already noted that there are two sets of conserved, non-local charges present. In fact, a little bit of analysis shows that the $`B`$-extensions of integrable models will always have at least two infinite sets of non-local conserved charges. For example, in the $`N=2`$ supersymmetric KdV-B hierarchy, let us note that the charges $$\overline{H}_n=𝑑x𝑑\theta _1𝑑\theta _2\left(D_2^1sRes(L^{\mathrm{nstd}})^{2n1}\right)=𝑑x𝑑\theta _1𝑑\theta _2(D_2^1h_n),$$ (20) will be conserved simply because these correspond to the conserved charges of the original system and the original equations are unmodified by this extension. Such a hierarchy of charges will always be present. (In the spirit of eq. (9), these charges would be obtained from the $`\theta _2`$ independent part of the densities.) They are manifestly non-local and, consequently, are not invariant under the $`N=2`$ extended supersymmetry. Let us also note that, by definition, $$𝑑x𝑑\theta _1𝑑\theta _2\left(^1D_1D_2sRes(L^{nstd})^{2n1}\right)=0.$$ There is also a second set of non-local charges which one can construct. Namely, let us evaluate the square root of the Lax operator for the non-standard representation. Conventionally, the super residues of the odd powers of the square root of the Lax operator gives rise to conserved non-local charges in supersymmetric integrable models. As we will now show, the system under study presents a novel feature and, consequently, leads to new charges. Let us note that since both $`D_1`$ and $`D_2`$ satisfy $$D_1^2==D_2^2,$$ it follows that the general form of the square root can be determined to have the form $`(L^{\mathrm{nstd}})^{1/2}`$ $`=`$ $`\alpha D_1+\beta D_2+2\alpha (D_2^1G)(\alpha (^1D_1D_2G)+\beta G)D_1^1`$ (21) $`+\left(\alpha (D_2G)\beta (D_1G)\beta (D_2^1G^2)\right)^1`$ $`+{\displaystyle \frac{1}{2}}\left(\alpha (^1D_1D_2G)^2+\beta (^1D_1D_2G^2)\right)D_1^3+\mathrm{},`$ where the constant parameters $`\alpha `$ and $`\beta `$ are constrained to satisfy $`\alpha ^2+\beta ^2=1`$. While non-local charges have been constructed earlier from square (and quartic) roots , here we have the novel feature that there is a one parameter family of square roots of the Lax operator. We think this feature would exist in extended supersymmetric models with $`N>1`$. In fact, let us note that for $`\alpha =1`$ and $`\beta =0`$, this square root coincides with what has been calculated earlier . But, this is, in fact, the more general form of the square root with a richer structure. From the structure of this general square root, let us note that we can construct conserved charges by taking the “$`sRes`$” of odd powers of the square root and would, in general, give non-local charges. In fact, let us note that the first few of these charges have the forms ($`dz=dxd\theta _1d\theta _2`$) $`{\displaystyle 𝑑zsRes(L^{\mathrm{nstd}})^{1/2}}`$ $`=`$ $`\beta {\displaystyle 𝑑zG},`$ $`{\displaystyle 𝑑zsRes(L^{\mathrm{nstd}})^{3/2}}`$ $`=`$ $`{\displaystyle \frac{3\alpha }{2}}{\displaystyle 𝑑zG^2},`$ $`{\displaystyle 𝑑zsRes(L^{\mathrm{nstd}})^{5/2}}`$ $`=`$ $`{\displaystyle }dz[2\beta ({\displaystyle \frac{1}{3}}G^3G(D_1D_2G))+\beta \left(D_2^1(G_x(D_1G))\right)`$ (22) $`{\displaystyle \frac{\alpha }{2}}({\displaystyle \frac{1}{3}}(D_1^1D_2G)^3+2(D_1^1((D_2G)(D_1D_2G))))].`$ Thus, we see that the “$`sRes`$” of the odd powers of the square root leads to conserved charges which are a combination of new non-local charges (the first few of which are local) as well as old ones of the form eq. (20). In fact, if we neglect the old non-local charges in these expressions, we see that the one parameter family of charges really leads to two distinct sets of conserved charges. Thus, for example, when $`\alpha =1`$ (and, therefore, $`\beta =0`$), the non-local charges coincide with what has been obtained earlier . However, when $`\beta =1`$, we have a new set of non-local conserved charges for the system. Thus, we conclude that, in this $`N=2`$ supersymmetric model, we have, in fact, three sets of conserved non-local charges. Furthermore, we now recognize that the three “exotic” charges belong to this hierarchy of non-local charges and can only be obtained if we take the general square root. (In other words, the first few members of the non-local hierarchy of charges is really local, even though higher order ones are truly non-local. We have explicitly verified with REDUCE that there are no more local “exotic” (bosonic) conserved charges present.) To complete the story of the “exotic” charges, let us look at the simpler system of $`N=1`$ supersymmetric KdV-B hierarchy. Here the Lax operator, as we have seen in eq. (14), has the form $$L=^2+(D\mathrm{\Phi }).$$ The fermionic Hamiltonians of the system are given by $$H_n=𝑑x𝑑\theta Res(L^{(2n1)/2})=𝑑x𝑑\theta h_n,n=1,2,\mathrm{},$$ (23) and the first set of non-local charges are given by $$\overline{H}_n=𝑑x𝑑\theta (D^1Res(L^{(2n1)/2}))=𝑑x𝑑\theta (D^1h_n).$$ (24) Since, in this case, we have only one fermionic coordinate, the quartic root is without any arbitrary parameter and the residues of the odd powers of it give rise to a linear combination of new non-local conserved charges and the ones in eq. (24). Ignoring these old charges, we can write the new set of non-local charges to be coming from $$\stackrel{~}{H}_n=𝑑x𝑑\theta Res(L^{(2n1)/4}),$$ (25) These are bosonic charges and explicitly, the first few of them have the form $`\stackrel{~}{H}_1`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle 𝑑x𝑑\theta \mathrm{\Phi }},`$ $`\stackrel{~}{H}_2`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle 𝑑x𝑑\theta (D^2\mathrm{\Phi })}=0,`$ $`\stackrel{~}{H}_3`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle 𝑑x𝑑\theta \mathrm{\Phi }(D\mathrm{\Phi })},`$ $`\stackrel{~}{H}_4`$ $`=`$ $`{\displaystyle \frac{3}{8}}{\displaystyle 𝑑x𝑑\theta \mathrm{\Phi }(D^3\mathrm{\Phi })},`$ $`\stackrel{~}{H}_5`$ $`=`$ $`{\displaystyle \frac{1}{8}}{\displaystyle 𝑑x𝑑\theta \mathrm{\Phi }\left((D^5\mathrm{\Phi })+2(D\mathrm{\Phi })^2\right)}.`$ (26) Of these, only $`\stackrel{~}{H}_3`$ was found earlier and termed “exotic” . We see that it belongs to a hierarchy of non-local charges, the first four of which are, in fact, local. (We suspect that, in this particular case, this new set of charges is indeed local. This follows from an analysis of the structure of the charges in the dispersionless limit . However, this is not a general feature.) ## 5 Alternate description: As we had noted earlier, the conventional $`B`$-extension cannot be applied to equations where the time evolution of the dynamical variable is not a space derivative. Furthermore, even when the $`B`$-extension exists, it gives non-local equations if applied more than once. In this section, we will describe very briefly, an alternate extension which does not suffer from this problem. Let us consider an integrable system of the form $$\varphi _t=B[\varphi ].$$ (27) If we now define a superfield $$\overline{G}(x,\theta _1,\mathrm{},\theta _{N+1})=\varphi +\theta _{N+1}\overline{\varphi },$$ (28) then, we note that the new superfield depends on one extra fermionic coordinate and has the same Grassmann parity as the original variable $`\varphi `$. If we now define a dynamical system described by $$\overline{G}_t=B[\overline{G}].$$ (29) then, this system would be integrable. We note that the $`\theta _{N+1}`$ independent part of this equation would correspond to eq. (27), the original equation, so that it provides an alternate description of the $`B`$-extension. In fact, when we can write $`B[\varphi ]=(A[\varphi ])_x`$, the two descriptions would be equivalent and will map into each other under $$\overline{G}(D_{N+1}G).$$ (30) All the discussions of the earlier sections can be carried out in this framework as well. However, the advantage of such a description may lie in the fact that the $`B`$-extended equation, in eq. (29), is exactly like the original equation independent of whether the right hand side is a total derivative or not. Consequently, we can apply $`B`$-extension to any given equation and as many times without running into the problems of non-locality. (In simple terms, the new variables in the alternate representation are more local than the older ones.) As a result, systems, such as the NLS and the AKNS hierarchies, which were thought not to have a local $`B`$-extension (in the standard approach) , can actually have one in this alternate description. ## 6 Conclusion In this letter, we have studied systematically the general features of $`B`$-extension of any given integrable system. We have brought out general features such as the Hamiltonians, Hamiltonian structures and recursion operators. We have clarified the origin of “exotic” charges in such models and have identified them as belonging to an infinite set of non-local charges. We have shown that, in such models, there naturally exist, at least, two infinite sets of non-local charges and, for $`N>1`$ supersymmetry, even more. We have explicitly shown that the $`N=2`$ supersymmetric KdV-B hierarchy has three sets of non-local conserved quantities and have discussed their construction starting from the Lax operator. ## Acknowledgments A.D. acknowledges support in part by the U.S. Dept. of Energy Grant DE-FG 02-91ER40685 while Z.P. is supported in part by the Polish KBN Grant 2 P0 3B 136 16.
warning/0004/math0004100.html
ar5iv
text
# On the Relation Between Pommaret and Janet BasesDedicated to the memory of Alyosha Zharkov. ## 1 Introduction Pommaret and Janet bases may be cited as typical representatives of involutive bases . Involutive bases are Gröbner bases, though, generally redundant, and involutive methods provide an alternative approach to computation of Gröbner bases. In so doing, polynomial Pommaret bases which were first introduced in have become a research subject in commutative algebra. They can be considered as generalized left Gröbner bases in the commutative ring with respect to non-commutative grading . Pommaret bases of homogeneous ideals in generic position coincide with their reduced Gröbner bases . The use of these bases makes more accessible the structural information of zero-dimensional ideals . Pommaret bases provide an algorithmic tool for determining combinatorial decompositions of polynomial modules and for computations in the syzygy modules . Linear differential Pommaret bases form the main tool in formal theory of linear partial differential equations whereas linear differential Janet bases form an algorithmic tool in Lie symmetry analysis of nonlinear differential equations . Unlike reduced Gröbner bases, Pommaret and Janet bases along with any other involutive bases lead to explicit formulae for Hilbert functions and Hilbert polynomials . Basic properties of Pommaret and Janet bases are determined by the underlying involutive divisions . Non-noetherity of Pommaret division is responsible for non-existence of finite Pommaret bases for some polynomial (linear differential) ideals of positive (differential) dimension. On the other hand, any polynomial ideal as well as any linear differential ideal has a finite Janet basis due to the noetherity of Janet division. The two divisions differ greatly by their definition: unlike Janet divisibility, Pommaret divisibility do not depend on the leading terms of generators. Given an ideal and an admissible monomial ordering, or ranking in the differential case, its monic Pommaret basis is unique whereas there are infinitely many different monic Janet bases and among them only the minimal Janet basis is uniquely defined . However, in spite of the above differences, Pommaret and Janet bases are closely related, and Zharkov was the first to observe this fact in the last paper of his life. He argued that if a polynomial ideal has a finite Pommaret basis and a Janet basis which is Pommaret autoreduced, then they have identical monic forms (c.f. ). Zharkov put also forward an algorithm for construction of Janet bases by sequential treatment of Janet nonmultiplicative prolongations followed by Pommaret autoreduction. The goal of this paper is to study the relation between polynomial Pommaret and Janet bases in more details and to improve the Zharkov algorithm. Our analysis is based on the properties of Janet and Pommaret divisions and involutive algorithms studied in . This paper is organized as follows. The next section sketches some definitions, notations and conventions which are used in the sequel. Section 3 deals with analysis of the relationships between polynomial Pommaret and Janet bases. In particular, we prove that if an ideal has a finite Pommaret basis, then it is a minimal Janet basis. We describe here an algorithm of the combined Pommaret and Janet autoreduction which, given a Janet basis, converts it into another Janet basis. Since a minimal Janet basis is both Janet and Pommaret autoreduced, the existence of a finite Pommaret basis is equivalent to Pommaret-Janet autoreducibility of any non-minimal Janet basis into a minimal one. In Section 4 we describe an algorithm for computation of polynomial Janet bases which is an improved version of the Zharkov algorithm . One of the improvements is the use of Pommaret-Janet autoreduction rather then the pure Pommaret autoreduction. Another improvement is incorporation of the involutive analogue of Buchberger’s chain criterion . Section 5 contains generalization of the results of Sections 3 and 4 to linear differential ideals. The generalization is based on paper where general involutive methods and algorithms of papers are extended from commutative to differential algebra. ## 2 Basic Definitions and Notations Let be the set of nonnegative integers, and $`\text{𝕄}=\{x_1^{d_1}\mathrm{}x_n^{d_n}|d_i\text{}\}`$ be a set of monomials in the polynomial ring $`\text{}=K[x_1,\mathrm{},x_n]`$ over a field $`K`$ of characteristic zero. By $`deg(u)`$ and $`deg_i(u)`$ we denote the total degree of $`u\text{𝕄}`$ and the degree of variable $`x_i`$ in $`u`$, respectively. If monomial $`u`$ divides monomial $`v`$ we shall write $`u|v`$. Throughout the paper we restrict ourselves with admissible monomial orderings $``$ which are compatible with $$x_1x_2\mathrm{}x_n.$$ (1) The leading monomial of the polynomial $`f\text{}`$ with respect to $``$ will be denoted by $`lm(f)`$. If $`F\text{}`$ is a polynomial set, then by $`lm(F)`$ we denote the leading monomial set for $`F`$, and $`Id(F)`$ will denote the ideal in $`R`$ generated by $`F`$. The initial ideal of an ideal $`I\text{}`$ with respect to the monomial ordering $``$ will be denoted by $`in_{}(I)`$. The support of a polynomial $`f`$, that is, the set of monomials occurring in $`f`$ with nonzero coefficients will be denoted by $`supp(f)`$. For the least common multiple of two monomials $`u,v\text{𝕄}`$ we shall use the conventional notation $`lcm(u,v)`$. ###### Definition 2.1. For a monomial $`u=x_1^{d_1}\mathrm{}x_k^{d_k}`$ with $`d_k>0`$ the variables $`x_j,jk`$ are considered as Pommaret multiplicative or $`P`$multiplicative and the other variables as $`P`$nonmultiplicative. For $`u=1`$ all the variables are $`P`$multiplicative. ###### Definition 2.2. Let $`U\text{𝕄}`$ be a finite monomial set. For each $`1in`$ divide $`U`$ into groups labeled by non-negative integers $`d_1,\mathrm{},d_i`$: $$[d_1,\mathrm{},d_i]=\{uU|d_j=deg_j(u),1ji\}.$$ A variable $`x_i`$ is called Janet multiplicative or $`J`$multiplicative for $`uU`$ if $`i=1`$ and $`deg_1(u)=\mathrm{max}\{deg_1(v)|vU\}`$, or if $`i>1`$, $`u[d_1,\mathrm{},d_{i1}]`$ and $`deg_i(u)=\mathrm{max}\{deg_i(v)|v[d_1,\mathrm{},d_{i1}]\}`$. If a variable is not $`J`$multiplicative for $`uU`$, it is called $`J`$nonmultiplicative for $`u`$. For a polynomial $`f\text{}`$ the Pommaret separation of variables into multiplicative and nonmultiplicative is done in accordance with Definition 2.1 where $`u=lm(f)`$. Analogously, for an element $`fF`$ in a finite polynomial set $`F\text{}`$ the Janet multiplicative and nonmultiplicative variables are determined by Definition 2.2 with $`u=lm(f)U=lm(F)`$. We denote by $`M_P(f)`$, $`NM_P(f)`$ and by $`M_J(f,F)`$, $`NM_J(f,F)`$, respectively, the sets of $`P`$multiplicative, $`P`$nonmultiplicative and $`J`$multiplicative, $`J`$nonmultiplicative variables for $`f`$. A set of monomials in $`P`$multiplicative variables for $`u`$ and $`J`$multiplicative variables for $`uU`$ will be denoted by $`P(u)`$ and $`J(u,U)`$, respectively. ###### Remark 2.3. The monomial sets $`P(u)`$ and $`J(u,U)`$ for any $`u,U`$ such as $`uU`$ satisfy the following axioms | (a) | If $`wL(u,U)`$ and $`v|w`$, then $`vL(u,U)`$. | | --- | --- | | (b) | If $`u,vU`$ and $`uL(u,U)vL(v,U)\mathrm{}`$, then | | | $`uvL(v,U)`$ or $`vuL(u,U)`$. | | (c) | If $`vU`$ and $`vuL(u,U)`$, then $`L(v,U)L(u,U)`$. | | (d) | If $`VU`$, then $`L(u,U)L(u,V)`$ for all $`uV`$. | if one takes either $`P(u)`$ or $`J(u,U)`$ as $`L(u,U)`$. The axioms characterize an involutive monomial division, a concept invented in . Every monomial set $`L(u,U)`$ satisfying the axioms generates an appropriate separation of variables into $`(L)`$multiplicative and $`(L)`$nonmultiplicative. As this takes place, an element $`uU`$ is an $`L`$divisor of a monomial $`w\text{𝕄}`$ if $`w/uL(u,U)`$. In this case $`w`$ is $`L`$multiple of $`u`$. Using the axioms, a number of new divisions was constructed which may be also used for algorithmic computation of involutive bases. All the next definitions in this section are those in specified to Pommaret and Janet divisions. ###### Definition 2.4. Given a finite monomial set $`U`$, its cone $`C(U)`$, $`P`$cone $`C_P(U)`$ and $`J`$cone $`C_J(U)`$ are the following monomial sets $$C(U)=_{uU}u\text{𝕄},C_P(U)=_{uU}uP(u),C_J(U)=_{uU}uJ(u,U).$$ ###### Definition 2.5. Given an admissible ordering $``$, a polynomial set $`F\text{}`$ is called Pommaret autoreduced or $`P`$autoreduced if every $`fF`$ has no terms $`P`$multiple of an element in $`lm(F)lm(f)`$. Similarly, a finite polynomial set $`F`$ is Janet autoreduced or $`J`$autoreduced if each term in every $`fF`$ has no $`J`$divisors among $`lm(F)lm(f)`$. A finite set $`F`$ will be called Pommaret-Janet autoreduced or $`PJ`$autoreduced if it is both $`P`$autoreduced and $`J`$autoreduced. ###### Remark 2.6. From Definition 2.2 it follows that any finite set $`U`$ of distinct monomials is $`J`$autoreduced $$(u,vU)(uv)[uJ(u,U)vJ(v,U)=\mathrm{}].$$ ###### Definition 2.7. Given an admissible ordering $``$ and a polynomial set $`F\text{}`$, a polynomial $`h\text{}`$ is said to be in $`P`$normal form modulo $`F`$ if every term in $`h`$ has no $`P`$ divisors in $`lm(F)`$. Similarly, if all the terms in $`h`$ have no $`J`$divisors among the leading terms of a finite polynomial set $`F`$, then $`h`$ is in $`J`$normal form modulo $`F`$. A general involutive normal form algorithm is described in , and an involutive normal form of a polynomial modulo any involutively autoreduced set is uniquely defined. We denote by $`NF_P(f,F)`$ and $`NF_J(f,F)`$, respectively, $`P`$normal and $`J`$normal form of polynomial $`f`$ modulo $`F`$. Pommaret or Janet autoreduction of a finite polynomial set $`F`$ may be performed similar to the conventional autoreduction . If $`H`$ is obtained from $`F`$ by the conventional, or $`J`$autoreduction we shall write $`H=Autoreduce(F)`$ or $`H=Autoreduce_J(F)`$, respectively. ###### Definition 2.8. A $`Pautoreduced`$ set $`F`$ is called a Pommaret basis ($`P`$basis) of the ideal $`Id(F)`$ generated by $`F`$ if $$(fF)(xNM_P(f))[NF_P(fx,F)=0].$$ (2) Similarly, a $`J`$autoreduced set $`F`$ is called a Janet basis ($`J`$basis) of $`Id(F)`$ if $$(fF)(xNM_J(f,F))[NF_J(fx,F)=0].$$ (3) In accordance with Definition 2.7 the nonmultiplicative prolongation $`fx`$ with the vanishing $`P`$ or $`J`$normal form modulo polynomial set $`F=\{f_1,\mathrm{},f_m\}`$ can be rewritten as $$fx=\underset{i=1}{\overset{m}{}}f_ih_i,h_i\text{},lm(f)xlm(f_ih_i)(1im)$$ where $`supp(h_i)P(f_i)`$ or $`supp(h_i)J(f_i,F)`$, respectively, for every polynomial product $`f_ih_i`$. Let $`G_P`$ and $`G_J`$ be Pommaret and Janet bases of an ideal $`I`$, respectively. Then, from Definition 3 it follows that $$hI\text{iff}NF_P(h,G_P)=0\text{and}hI\text{iff}NF_J(h,G_J)=0.$$ (4) This implies, in particular, the equalities $$C_P(lm(G_P))=C(lm(G_P)),C_J(lm(G_J))=C(lm(G_J)).$$ (5) It is immediate from (5) that $`lm(G_P)`$ and $`lm(G_J)`$ are $`P`$and $`J`$bases of the initial ideal $`in_{}(I)`$. ###### Corollary 2.9. If for a $`P`$autoreduced set $`G_P`$ the equality (5) of its cone and $`P`$cone holds and $`lm(G_P)`$ is a basis of the initial ideal $`in\left(Id(G_P)\right)\}`$, then $`G_P`$ is a $`P`$ basis of $`Id(G_P)`$. Analogous statement holds for a $`J`$basis. Whereas monic Pommaret bases much like to reduced Gröbner bases are unique, every ideal, by property (2.6), has infinitely many monic Janet bases. Among them there is the unique $`J`$basis defined as follows. ###### Definition 2.10. A Janet basis $`G`$ of ideal $`Id(G)`$ is called minimal if for any other $`J`$basis $`F`$ of the ideal the inclusion $`lm(G)lm(F)`$ holds. ###### Remark 2.11. Every zero-dimensional polynomial ideal has a finite Pommaret basis, and for a positive dimensional ideal the existence of finite Pommaret basis can be always achieved by means of an appropriate linear transformation of variables . ## 3 Relation Between Polynomial Pommaret and Janet Bases Given a finite monomial set $`U`$, Definitions 2.1 and 2.2 generally give different separations of variables for elements in $`U`$. ###### Example 3.1. $`U=\{x_1x_2,x_2x_3,x_3^2\}`$. | monomial | Pommaret | | Janet | | | --- | --- | --- | --- | --- | | | $`M_P`$ | $`NM_P`$ | $`M_J`$ | $`NM_J`$ | | $`x_1x_2`$ | $`x_2,x_3`$ | $`x_1`$ | $`x_1,x_2,x_3`$ | $``$ | | $`x_2x_3`$ | $`x_3`$ | $`x_1,x_2`$ | $`x_2,x_3`$ | $`x_1`$ | | $`x_3^2`$ | $`x_3`$ | $`x_1,x_2`$ | $`x_3`$ | $`x_1,x_2`$ | Here is, however, an important relation between Pommaret and Janet separations: ###### Proposition 3.2. (see also ). If a finite monomial set $`U`$ is $`P`$autoreduced, then for any $`uU`$ the following inclusions hold $$M_P(u,U)M_J(u,U),NM_J(u,U)NM_P(u,U).$$ For $`U`$ in Example 3.1 the minimal Janet basis $`U_J`$ and the Pommaret basis $`U_P`$ of the monomial ideal $`Id(U)`$ are $$U_J=U\{x_1x_3^2\},U_P=U_J_{i=2}^{\mathrm{}}\{x_1^ix_2\}_{j=2}^{\mathrm{}}\{x_1^jx_3^2\}_{k=2}^{\mathrm{}}\{x_2^kx_3\}.$$ (6) Note, that $`U_P`$ is infinite and $`U_JU_P`$. Below we show that the inclusion $`G_JG_P`$ holds for any minimal Janet basis $`G_J`$ and Pommaret basis $`G_P`$ if both of them are monic and generate the same ideal. Furthermore, the proper inclusion $`G_JG_P`$ holds iff $`P`$ is infinite. The following algorithm, given a finite polynomial set $`F\text{}`$ and an admissible ordering $``$, performs $`PJ`$autoreduction of $`F`$ and outputs a $`PJ`$-autoreduced set $`H`$. In this case we shall write $`H=Autoreduce_{PJ}F`$. Since the involutive $`P`$ and $`J`$reductions which are performed in the course of the algorithm form subsets of the conventional reductions , the algorithm terminates. Furthermore, the while-loop generates the $`P`$autoreduced monomial set $`lm(H)`$. In accordance with Remark 2.6 the Janet autoreduction in line 12 does not affect the leading terms, and, hence, the output polynomial set is both Pommaret and Janet autoreduced. | Algorithm: Pommaret-JanetAutoreduction($`F,`$) | | --- | | Input: $`F\text{}`$, a finite set; $``$, an admissible ordering | | Output: $`H=Autoreduce_{PJ}(F)`$ | | begin | | $`G:=\mathrm{}`$; $`H:=F`$ 1 repeat 2 $`\stackrel{~}{H}:=H`$ 3 while exist $`hH`$ such that 4 $`lm(h)C_P\left(lm(H\{h\})\right)`$ do 5 $`H:=H\{h\}`$; $`G:=G\{h\}`$ 6 for each $`gG`$ do 7 $`G\{g\}`$; $`f:=NF_J(g,H)`$ 8 if $`f0`$ then 9 $`H:=H\{f\}`$ 10 until $`H=\stackrel{~}{H}`$ 11 $`H:=Autoreduce_J(H)`$ 12 end | ###### Proposition 3.3. If algorithm Pommaret-JanetAutoreduction takes a Janet basis $`F`$ as an input its output $`H`$ is also a Janet basis of the same ideal, and $`HF`$. ###### Proof 3.4. Let $`F`$ be a Janet basis of the ideal $`Id(F)`$ and $`H`$ be a polynomial set which computed by the algorithm. Apparently, $`Id(H)=Id(F)`$. Consider $`U=lm(F)`$. If $`U`$ is $`P`$autoreduced, then $`H`$ initiated as $`F`$ in line 1 does not change in the process of the algorithm, and, hence, $`H=F=Autoreduce_{PJ}(F)`$. Otherwise, consider the output polynomial sets $`H`$. Denote $`lm(H)`$ by $`V`$. Then, by construction, $`VU`$. Consider a monomial $`tC_J(U)`$ and show that $`tC_J(V)`$. Let $`uU`$ be a $`J`$divisor of $`t`$, that is, $`tuJ(u,U)`$, and $`vV`$ be such that $`v|u`$. If $`u=v`$, by property (d) of Janet division in Remark 2.3, we are done. Let now $`uUV`$. The while-loop provides that $`uvP(v)`$. If $`v=x_1^{d_1}\mathrm{}x_k^{d_k}`$ with $`d_k>0`$ $`(1kn)`$, then from Definition 2.1 it follows that $`u=x_1^{d_1}\mathrm{}x_k^{d_k+e_k}\mathrm{}x_n^{e_n}`$. We have to prove that any variable $`x_iJ(u,U)`$ $`(1in)`$ satisfy $`x_iJ(v,V)`$. Consider two alternative cases: $`i<k`$ and $`ik`$. In the first case, by Definition 2.2, both $`u,v`$ belong to the same group $`[d_1,\mathrm{},d_{i1}]`$ of monomials in $`U`$. It follows that $`x_iJ(v,V)`$. In the second case, by Definition 2.1, $`x_iP(v)`$ and Proposition 3.2 we find again that $`x_i`$ is $`J`$multiplicative for $`v`$ as an element in $`V`$. Therefore, $`V`$ is a Janet monomial basis of $`Id(U)`$. Thus, by Corollary 2.9, $`H`$ is a $`J`$basis. In so doing, every $`J`$normal form computed in line 8 of the algorithm apparently vanishes. This implies the inclusion $`HF`$. ∎ ###### Corollary 3.5. A minimal Janet basis is Pommaret autoreduced. ###### Proof 3.6. Let $`G`$ be a minimal Janet basis. From Proposition 3.3 and Definition 2.10 it follows that $`lm(G)`$ is $`P`$autoreduced. Thus, $`G`$ is $`P`$autoreduced. ∎ It is clear that, given a Janet basis, its $`PJ`$autoreduction yields, generally, more compact basis than the pure $`P`$autoreduction. ###### Example 3.7. Consider polynomial set $$F=\{xyztxz,xyz+z^2,xzt+x^2,xy+z,zt+x\}$$ which is a Janet basis of the ideal $`Id(F)`$ with respect to the degree-reverse-lexicographical ordering $``$ such that $`xyzt`$. Given $`F`$ and $``$ as input, the algorithm Pommaret-JanetAutoreduction$`(F,)`$ outputs the minimal $`J`$basis $$G_1=\{xzt+x^2,xy+z,zt+x\}.$$ If one uses the Pommaret normal form computation in line 7 instead of that of Janet, then the algorithm leads to the output $$G_2=\{xzt+x^2,z^2t+xz,xy+z,zt+x\}=G_1\{z^2t+xz\}.$$ The following theorem is the main theoretical result of the present paper and forms a basis of an algorithm for construction of Janet and Pommaret bases described in the next section. ###### Theorem 3.8. Given an ideal $`IR`$ and an admissible monomial ordering $``$ compatible with (1), the following are equivalent: | $`(i)`$ | $`I`$ has a finite Pommaret basis. | | --- | --- | | $`(ii)`$ | A minimal Janet basis of $`I`$ is its Pommaret basis. | | $`(iii)`$ | If $`F`$ is a Janet basis of $`I`$, then $`G=Autoreduce_{JP}(F)`$ is | | | its Pommaret basis. | ###### Proof 3.9. $`(i)(iii)`$: Suppose $`G=\{g_1,\mathrm{},g_m\}`$ which, by Proposition 3.3, is also a $`J`$basis of $`I`$, is not its $`P`$basis. Our goal is to prove that a Pommaret basis of $`I`$ is an infinite polynomial set. From Corollary 2.9 it follows that $$(gG)(x_kNM_P(g))[lm(g)x_kC_P\left(lm(G)\right)].$$ (7) Among nonmultiplicative prolongations $`gx_k`$ satisfying (7) choose one with the lowest $`lm(g)x_k`$ with respect to the pure lexicographical ordering $`_{Lex}`$ generated by (1). If there are several such prolongations choose that with the lexicographically lowest $`x_k`$, that is, with the lexicographically highest $`g`$. This choice is unique since $`G`$ is $`J`$autoreduced. We claim that $`x_kM_J(g,G)`$. Assume for a contradiction that $`x_kNM_J(g,G)`$. Then from Janet involutivity conditions (3) we obtain $$(fG)[lm(g)x_k=lm(f)w|wJ(lm(f),lm(G))].$$ (8) In accordance with condition (7) $`w`$ contains $`P`$nonmultiplicative variables for $`f`$ and from (8) it follows that $`f_{Lex}g`$. If $`w=x_j`$, then $`x_jx_k`$, and both $`lm(f)x_j`$ and $`lm(g)`$ belong to the same monomial group $`[deg_1(g),\mathrm{},deg_{k1}(g)]`$ appearing in Definition 2.2. Hence, $`x_jNM_J(f,G)`$ in contradiction to (8). Therefore, $`deg(w)2`$ and (8) can be rewritten as $$lm(g)x_k=\left(lm(f)v\right)x_j$$ with $`w=vx_j`$. Show that $`lm(f)vP\left(lm(f)\right)`$. If $`vC_P\left(lm(G)\right)`$ we are done. Otherwise there is $`x_m|v`$ such that $`x_mNM_P\left(lm(f)\right)`$. Since $`lm(f)x_m_{Lex}lm(g)x_k`$, our choice of $`g`$ and $`x_k`$ implies the existence $`g_1G`$ such that $`fx_m=g_1v_1`$ and $`v/x_mP\left(lm(g_1)\right)`$. If $`v_1=v/x_mP\left(lm(g_1)\right)`$ we are done. Otherwise we select again a $`P`$nonmultiplicative variable for $`g_1`$ occurring in $`v_1`$ and rewrite the corresponding prolongation in terms of its $`P`$divisor $`lm(g_2)lm(G)`$. Continuity of Pommaret division provides termination of the rewriting process with an element in $`\stackrel{~}{g}G\{g\}`$ such that $`lm(\stackrel{~}{g})`$ is a $`P`$divisor of $`lm(f)v`$. Because $`G`$ is $`P`$autoreduced, by Proposition 3.2 $`lm(\stackrel{~}{g})`$ is also a $`J`$divisor of $`lm(f)v`$. By this means there are two different Janet divisors $`lm(f)`$ and $`lm(\stackrel{~}{g})`$ of the same monomial that contradicts Remark 2.6 and proves the claim. Let now $`H`$ be a $`P`$basis of $`Id(G)`$. Denote $`lm(g)x_k`$ by $`u`$ and show that $`ulm(H)`$. Suppose there is an element $`hHG`$, such that $`u=lm(h)v`$ with $`vP\left(lm(h)\right)`$. Then $`h_{Lex}u`$ and there is $`qG`$ satisfying $`lm(h)=lm(q)w`$ where $`wP\left(lm(q)\right)`$. Thus, there is a $`P`$nonmultiplicative prolongation $`qx_j`$ with $`x_j|w`$ such that $`lm(q)x_j_{Lex}u`$ and $`lm(q)x_jC_P\left(lm(G)\right)`$ that contradicts the above choice of $`g`$ and $`x_k`$. Now consider monomial set $`U=lm(G)\{u\}lm(H)`$. By Definition 2.2, $`x_kNM_P(u)`$ and $`ux_k`$ is obviously the lexicographically lowest $`P`$nonmultiplicative prolongation of elements in $`U`$. It is easy to see that $`ux_kC_P(U)`$. Indeed, since $`x_kM_J(g,G)`$, it follows that $`ux_kU`$, and a $`P`$divisor of $`ux_k`$ would also $`P`$divide $`lm(g)`$ that is impossible as $`G`$ is $`P`$autoreduced. Thus, we find that $`ux_klm(H)`$. By sequential repetition of this reasoning for $`ux_k^i`$ $`(i=2,3\mathrm{})`$ we deduce that every such monomial is an element in $`lm(H)`$, and, therefore, $`H`$ is infinite. $`(iii)(ii)`$: If $`G`$ is a minimal Janet basis of $`Id(G)`$, then Corollary 3.5 implies $`G=Autoreduce_{JP}(G)`$, and the above arguments show that either $`G`$ is also a $`P`$basis or the latter is infinite. $`(ii)(i)`$: This implication is obvious. ∎ ###### Corollary 3.10. Let $`G_J`$ be a monic minimal Janet basis and $`G_P`$ be a monic Pommaret basis for the same polynomial ideal and monomial ordering. Then $`G_JG_P`$ and $`G_JG_P`$ iff $`G_P`$ is infinite. ###### Proof 3.11. This follows from the above proof of Theorem 3.8. ∎ ###### Corollary 3.12. If a polynomial ideal is in generic position, then its minimal Janet basis is also a Pommaret basis. If such an ideal is homogeneous, then these bases are also reduced Gröbner bases of the ideal. ###### Proof 3.13. $`I`$ has a finite Pommaret basis . If $`I`$ is homogeneous, then its Pommaret basis is the reduced Gröbner basis . ∎ ## 4 Algorithm for Construction of Janet Bases In this section we present an algorithm for constructing $`J`$bases of polynomial ideals which is based on Theorem 3.8 and will be called JanetBasis. Whenever the ideal generated by an input polynomial set has a finite $`P`$basis for a given admissible ordering, the algorithm outputs just this basis which is also a minimal $`J`$basis. Otherwise, the $`J`$basis computed by the algorithm is not necessarily minimal as we demonstrate below by the explicit example. | Algorithm: JanetBasis($`F,`$) | | --- | | Input: $`F\text{}`$, a finite set; $``$, an admissible ordering | | Output: $`G`$, an involutive basis of the ideal $`Id(F)`$ | | begin | | $`G:=Autoreduce(F)`$ 1 $`T:=\mathrm{}`$ 2 for each $`gG`$ do 3 $`T:=T\{(g,lm(g),\mathrm{})\}`$ 4 while exist $`(g,u,D)T`$ and $`xNM_J(g,G)D`$ do 5 choose such $`(g,u,D),x`$ with the lowest $`lm(g)x`$ w.r.t. $``$ 6 $`T:=T\{(g,u,D)\}\{(g,u,D\{x\})\}`$ 7 if $`Criterion(gx,u,T)`$ is false then 8 $`h:=NF_J(gx,G)`$ 9 if $`h0`$ then 10 if $`lm(h)=lm(gx)`$ then 11 $`T:=T\{(h,u,\mathrm{})\}`$ 12 else 13 $`T:=T\{(h,lm(h),\mathrm{})\}`$ 14 $`G:=Autoreduce_{PJ}(G\{h\})`$ 15 $`Q:=T`$ 16 $`T:=\mathrm{}`$ 17 for each $`gG`$ do 18 if exist $`(f,u,D)Q`$ s.t. $`lm(f)=lm(g)`$ then 19 choose $`\stackrel{~}{g}G`$ s.t. $`ulm(\stackrel{~}{g})J(lm(\stackrel{~}{g}),lm(G))`$ 20 $`T:=T\{(g,lm(\stackrel{~}{g}),DNM_J(g,G))\}`$ 21 else 22 $`T:=T\{(g,lm(g),\mathrm{})\}`$ 23 end | $`Criterion(g,u,T)`$ is true provided that if there is $`(f,v,D)T`$ such that $`lcm(u,v)lm(g)`$ and $`lm(g)lm(f)J(lm(f),lm(G))`$. The structure of algorithm JanetBasis is very close to that of the algorithm InvolutiveBasis devised in for construction of involutive bases for arbitrary constructive involutive divisions, and, hence, applicable to Janet division. The main difference between the algorithms is the form of their intermediate autoreduction. Whereas the previous algorithm, when specified for Janet division, uses the pure Janet autoreduction which do not affect the leading terms, the below one uses the above algorithm Pommaret-JanetAutoreduction. By this reason, the algorithm InvolutiveBasis, unlike the below one, almost never outputs a minimal Janet basis or a Pommaret basis if the latter is finite. In paper we designed the algorithm MinimalInvolutiveBasis which always outputs a minimal involutive basis whenever the latter exists. As we now see from Theorem 3.8, this algorithm in the case of Janet division outputs also a Pommaret basis if it is finite. Besides, in the algorithm InvolutiveBasis the involutive autoreduction is performed whenever nonzero normal form is obtained in the while-loop. In the algorithm JanetBasis the subalgorithm Pommaret-JanetAutoreduction is caused by a nonzero $`J`$normal form $`h`$ in line 15 only if the leading term of the related prolongation is $`J`$reducible. Otherwise, since $`G`$ is always $`P`$autoreduced before its enlargement with $`h`$, $`lm(h)`$ cannot $`P`$divide, in accordance with Definition 2.1, any other element in $`lm(G)`$. Note that we indicated the intersection in line 21 to emphasize that elements in $`D`$ must be nonmultiplicative variables for the corresponding polynomial that is always understood in algorithms of papers (). Noetherity of Janet division provides termination of the algorithm JanetBasis . To show this consider the intermediate bases $`G_0=Autoreduce(F)`$ and $`G_i`$ $`(i=1,2,\mathrm{})`$ generated after the $`ith`$ iteration of the while-loop. It is clear that $$Id\left(lm(G_0)\right)Id\left(lm(G_1)\right)Id\left(lm(G_2)\right)\mathrm{}$$ (9) and this chain is stabilized after finitely many steps. Namely, the stabilization starts when the intermediate polynomial set $`G`$ becomes a (non-necessarily reduced) Gröbner basis of $`Id(F)`$. By partial involutivity of $`G`$ , the proper inclusion in chain (9) holds only when $`G`$ is enlarged by $`h`$ in line 15. In between of such proper inclusions and after the chain stabilization $`lm(G)`$ is completed with $`lm(h)=lm(gx)`$ as stands in line 11, and this completion cannot be infinite by noetherity of Janet division . Once algorithm terminates, it produces, by Proposition 3.3, a $`PJautoreduced`$ Janet basis of $`F`$ because the involutivity conditions (3) are satisfied as is checked in lines 6 and 9 where $`h=0`$. In so doing, correctness of the criterion which is verified in line 8 is proved exactly as it done in for the algorithm InvolutiveBasis. ###### Remark 4.1. In line 6 of the algorithm JanetBasis one can use any admissible ordering for selection of the current $`J`$nonmultiplicative prolongation $`gx`$ to be treated in the following lines. This selection ordering may not only be different from the main ordering $``$ but also may vary at every step when the selection is done. Correctness of this arbitrariness in the choice of selection ordering follows from the correctness of this arbitrariness for the monomial completion procedure . As mentioned above, the algorithm JanetBasis for an ideal of positive dimension may not output its minimal $`J`$basis. We demonstrate this by the following example. ###### Example 4.2. Let $`F`$ be a set $`\{x^2yz,xy^2y\}`$ generating one-dimensional ideal, and $``$ be the degree-reverse-lexicographical ordering with $`xyz`$. Then the algorithm JanetBasis$`(F,`$) outputs the following Janet basis $$G=\{x^2yz,x^2zz^3,xyyz,xzz^2,y^2zy,yz^2z\}$$ whereas the minimal Janet basis coinciding with the reduced Gröbner basis is $$\{xyyz,xzz^2,y^2zy,yz^2z\}.$$ ###### Remark 4.3. The algorithm JanetBasis is an improved version of the algorithm designed by Zharkov . The first improvement is the use of the mixed Pommaret-Janet autoreduction instead of the pure Pommaret autoreduction as Zharkov proposed. Let $`G_1`$ and $`G_2`$ be output Janet bases computed with the use of $`PJ`$ and $`P`$autoreduction, respectively. Then, Proposition 3.2 implies $`G_1G_2`$ and below we give an example when the proper inclusion holds. The second improvement is the criterion used in line 7. This criterion is an involutive analogue of the Buchberger’s chain criterion and is superior to the criterion used in . ###### Example 4.4. The following polynomial set generates three-dimensional ideal $$F=\left\{\begin{array}{c}4x_5(x_6+2a_18x_1)(a_2a_3)x_2x_3x_4+x_2+x_4\hfill \\ 4x_5(x_6+2a_18x_2)(a_2a_3)x_1x_3x_4+x_1+x_3\hfill \\ 4x_5(x_6+2a_18x_3)(a_2a_3)x_1x_2x_4+x_2+x_4\hfill \\ 4x_5(x_6+2a_18x_4)(a_2a_3)x_1x_2x_3+x_1+x_3\hfill \end{array}\right\}$$ (10) where $$a_1=x_1+x_2+x_3+x_4,a_2=x_1x_2x_3x_4,a_3=x_1x_2+x_2x_3+x_3x_4+x_4x_1.$$ For the polynomial set (10) and the degree-reverse-lexicographical ordering compatible with (1) the algorithm Janet basis outputs set $`G_1`$ with 71 polynomials. The pure Pommaret autoreduction in line 15 generates set $`G_2G_1`$ of 75 elements. Note that a minimal Janet basis contains 49 polynomials whereas the reduced Gröbner basis contains 44 polynomials. ## 5 Linear Differential Bases Let 𝕂 be a zero characteristic differential field with a finite number of mutually commuting derivation operators $`/x_1,\mathrm{},/x_n`$. Consider the differential polynomial ring $`\text{𝔻ℝ}=\text{𝕂}\{y_1,\mathrm{},y_m\}`$ with the set of differential indeterminates $`\{y_1,\mathrm{},y_m\}`$. Elements in 𝔻ℝ are differential polynomials in $`\{y_1,\mathrm{},y_m\}`$. An ideal in 𝔻ℝ generated by linear differential polynomials is called linear differential ideal . In , by exploiting the well-known algorithmic similarities between polynomial and linear differential systems and the association between monomials and the derivatives, we extended the general involutive methods and algorithms designed in for polynomial ideals to linear differential ideals. The statements and algorithms of Sect. 3 and 4 admit similar extension. As this takes place, all the above statements have proven for the polynomial case are proved by parallel arguments for the differential case. In the following table we give a short correspondence between these two cases. In particular, this correspondence allows one to rewrite the algorithms Pommaret-JanetAutoreduction and JanetBasis for linear differential bases. | Commutative Algebra | Differential Algebra | | --- | --- | | $`\text{}=K[x_1,\mathrm{},x_n]`$ | $`\text{𝔻ℝ}=\text{𝕂}\{y_1,\mathrm{},y_m\}`$ | | $`f,g,h\text{}`$ | $`f,g,h\text{𝔻ℝ}`$ | | $`F,G,H\text{}`$ | $`F,G,H\text{𝔻ℝ}`$ | | $`x^\alpha x_1^{\alpha _1}x_n^{\alpha _n}`$ | $`_\alpha y_j\frac{^{\alpha _1+\mathrm{}+\alpha _n}y_j}{x_1^{\alpha _1}\mathrm{}x_n^{\alpha _n}}[x^\alpha ]_j`$ | | $`gx`$ | $`_xg`$ | | monomial ordering | ranking | | $`lm(g)`$ | leading derivative $`ld(g)`$ | | $`lm(G)`$ | $`ld(G)=_{gG}\{ld(g)\}`$ | ## 6 Conclusion As we have seen finite Pommaret bases of polynomial and linear differential ideals are minimal Janet bases. The above proof of Theorem 3.8 shows that, given a $`P`$autoreduced Janet basis $`G`$, it is easy to verify the existence of a finite Pommaret basis. One suffices to check the condition (7). Another check which may be even easier in practice is to look at the Pommaret and Janet separation of variables for elements in $`G`$. As shown in , the existence of a finite Pommaret basis implies coincidence of both separations for every element in $`G`$. Moreover, the leading monomial structure of an infinite Pommaret basis can be read off the structure of $`lm(G)`$ (c.f. (6) for Example 3.1). Therefore, minimal Janet bases can be used in commutative and differential algebra as well as a Pommaret bases with the advantage of finiteness. For example, in the formal theory of differential equations infinity of a Pommaret basis signals on $`\delta `$-singularity of the coordinate system chosen, and the condition (7) gives the same signal for Janet bases. On the face of it, Pommaret division looks like more attractive than Janet one since its separation is easier to compute than the Janet separation. Besides, Pommaret division, unlike that of Janet, is globally defined . Hence, after enlargement of the intermediate polynomial set with an irreducible $`P`$nonmultiplicative prolongation there are no needs to recompute the Pommaret separation for other elements in the set. However, the careful implementation of both divisions do not reveal any notable advantage of Pommaret division over Janet division in construction of polynomial bases. This rather surprising fact was firstly observed by Zharkov . One of the explanations of this experimental phenomenon is given by Proposition 3.2: one must generally treat more $`P`$nonmultiplicative prolongations than $`J`$nonmultiplicative ones. In so doing, the search for an involutive divisor in the process of involutive reduction can be done similarly for both divisions as we show in our forthcoming paper . The algorithm JanetBasis presented above is now under implementation in C in parallel with the algorithm MinimalInvolutiveBasis specified for Janet division. Our first experimentation with the codes shows that sometimes the former algorithm runs faster than the latter one and needs less computer memory. For example, modular computation of the degree-reverse-lexicographic Janet basis for the Cyclic 7 example is about twice faster with the algorithm JanetBasis than with the algorithm MinimalInvolutiveBasis. Currently, the timings for computation modulo 31013 are about 5 and 10 minutes, respectively, on a Pentium Pro 333 Mhz computer. ## 7 Acknowledgements I am grateful to W.M.Seiler for useful discussions and for providing me with his recent paper . I would also like to thank D.A.Yanovich for his assistance in computing the above examples. This work was partially supported by grant INTAS-99-1222 and grants No. 98-01-00101 and 00-15-96691 from the Russian Foundation for Basic Research.
warning/0004/quant-ph0004059.html
ar5iv
text
# Sampling the canonical phase from phase-space functions ## I Introduction Studying the role of phase in quantum mechanics has a long history (for a review on the phase concepts, see, e.g., ). Its importance in today’s problems is also apparent. For example, phase in atomic systems has recently been used for storing quantum information , and phase and photon number measurements have been considered as a basis in some quantum teleportation schemes . Notwithstanding the various phase-dependent effects in quantum physics, phase itself has not been uniquely measured and its very definition as physical quantity has been subject to many disputes. Whereas for highly excited (quasi-classical) states different approaches give similar results, the various concepts differ in the phase properties of quantum states close to vacuum. Therefore the question has been arisen of what are the differences between these approaches and how relevant are they experimentally. In this paper we concentrate on the canonical phase and its relation to $`s`$-parametrized phase-space functions, with special emphasis on the measurability of its exponential moments by “weighted” averaging of measured phase space functions. The canonical phase distribution $`P(\phi )`$ of a radiation field mode (harmonic oscillator) prepared in a quantum state described by a density operator $`\widehat{\varrho }`$ is defined by $`P(\phi )=(2\pi )^1\phi |\widehat{\varrho }|\phi ,`$ (1) where the Fock state expansion of the (unnormalizable) phase states $`|\phi `$ reads $`|\phi ={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}e^{in\phi }|n.`$ (2) Even though there has been no known experimental scheme that is directly governed by $`P(\phi )`$, this distribution has very nice properties: it is non-negative, conjugated to the photon-number distribution (in the sense that a phase shifter shifts a phase distribution while a number shifter does not change it ), there exist number-phase uncertainty relations , and in comparison to other phase distributions, $`P(\phi )`$ is the most sharp one. The lack of direct experimental availability of $`P(\phi )`$ has led us to the search of schemes for sampling the canonical phase statistics from quantities that can be measured directly . In balanced homodyne detection (for a review on quantum state measurement using homodyning, see, e.g., ), the exponential moments $`\mathrm{\Psi }_k`$ of the canonical phase, $`\mathrm{\Psi }_k={\displaystyle _{2\pi }}𝑑\phi e^{ik\phi }P(\phi ),\mathrm{\Psi }_k=\mathrm{\Psi }_k^{},`$ (3) can be sampled by integrating the measured quadrature-component statistics multiplied by well-behaved kernel functions . An advantage of the method is that it applies to both the quantum regime and the classical regime in a unified way. Of course, the question has been as of whether or not it is possible to find other (and possibly better) measurement schemes suitable for sampling the exponential moments of the canonical phase. It is well known that balanced double-homodyne detection (eight-port homodyning, ) provides us with a two dimensional set of data whose statistics correspond to a $`s`$-parametrized phase-space function $`W_s(q,p)`$ with $`s`$ $``$ $`1`$ . In this scheme, the limiting case of $`s`$ $`=`$ $`1`$, which corresponds to the Husimi $`Q`$-function $`Q(q,p)`$ $`=`$ $`W_1(q,p)`$, requires perfect detection, i.e., 100% detection efficiency. Having a sampling scheme leading from a measured $`s`$-parametrized phase-space function to the exponential canonical-phase moments would be the most direct method of measuring the exponential moments of the canonical phase. Since each measurement event $`(q,p)`$ already yields a phase value Arg$`(q+ip)`$, the measured values only need to be “weighted” by the kernel functions in the averaging procedure yielding the exponential moments $`\mathrm{\Psi }_k`$. There are also measuring schemes, e.g., unbalanced homodyning, suitable for determining $`s`$-parametrized phase-space functions $`W_s(q,p)`$ with larger values of $`s`$ . However, in these schemes the functions $`W_s(q,p)`$ are not obtained in terms of the statistics of measurement events $`(q,p)`$, but they are obtained pointwise for each phase-space point $`(q,p)`$ set up in the experiment. Moreover, they are typically reconstructed from the measured data rather than measured directly. Nevertheless, it is interesting to ask the question of the prospects of phase measurement in schemes of that type. In this paper, we try to answer the questions raised above, focusing our attention to the problem of using balanced double-homodyne detection for sampling the exponential moments of the canonical phase. In Sec. II we present the kernels that relate the $`s`$-parametrized phase-space functions to the exponential phase moments, and in Sec. III we apply the results to direct sampling of the exponential phase moments in balanced double-homodyne detection. Other measurement schemes are discussed in Sec. IV. Section V addresses the problem of determining the phase distribution itself, and a conclusion is given in Sec. VI. ## II The kernel function Our task is to find the kernel function $`K_k(r;s)`$ such that the exponential moments of the canonical phase can be given by ($`k`$ $`>`$ $`0`$) $`\mathrm{\Psi }_k=\widehat{E}^k`$ (5) $`={\displaystyle _0^{2\pi }}𝑑\phi e^{ik\phi }{\displaystyle _0^{\mathrm{}}}r𝑑rW(r,\phi ;s)K_k(r;s),`$ and $`\mathrm{\Psi }_k`$ $`=`$ $`\mathrm{\Psi }_k^{}`$, where $`\widehat{E}={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}|nn+1|.`$ (6) In Eq. (5), the phase-space function $`W(r,\phi ;s)`$ is written in polar coordinates, i.e., $`W(r,\phi ;s)`$ $`=`$ $`W_s(r\mathrm{cos}\phi ,r\mathrm{sin}\phi )`$. Note that $`e^{ik\phi }K_k(r;s)`$ is the $`(s)`$-parametrized phase-space function of the operator $`\widehat{E}^k`$. We now take advantage of the expression $$\mathrm{\Psi }_k=\underset{l=0}{\overset{\mathrm{}}{}}\underset{n=0}{\overset{l}{}}\frac{(1)^{ln}}{(ln)!\sqrt{n!(l+n)!}}\widehat{a}^l\widehat{a}^{l+k},$$ (7) where the expectation value of the normally ordered correlations of the photon creation and destruction operators can be calculated by means of $`W(r,\phi ;s)`$ as $`\widehat{a}^l\widehat{a}^{l+k}=(1)^ll!\left({\displaystyle \frac{1s}{2}}\right)^l{\displaystyle _0^{2\pi }}𝑑\phi e^{ik\phi }`$ (9) $`\times {\displaystyle _0^{\mathrm{}}}rdrr^kL_l^k\left({\displaystyle \frac{2r^2}{1s}}\right)W(r,\phi ;s)`$ ($`L_l^k`$, Laguerre polynomial). Combining Eqs. (5), (7), and (9), we derive (Appendix A) $`K_k(r;s)={\displaystyle \frac{r^k2^{k+1}}{\pi ^{k/2}}}{\displaystyle _0^{\mathrm{}}}d\rho \{\rho ^{k1}\mathrm{\Omega }^{(k)}(\rho ^2)`$ (11) $`\times [1+s+(1s)e^{\rho ^2}]^{k1}\mathrm{exp}[{\displaystyle \frac{2(1e^{\rho ^2})r^2}{1+s+(1s)e^{\rho ^2}}}]\}.`$ Here, the function $`\mathrm{\Omega }^{(k)}(\rho ^2)`$ is given by $$\mathrm{\Omega }^{(k)}(\rho ^2)=e^{\rho ^2}\underset{n=0}{\overset{\mathrm{}}{}}\frac{(1)^n}{n!}A_n^{(k)}\rho ^{2n},$$ (13) where $`A_n^{(k)}={\displaystyle _0^\pi }𝑑\phi _1\mathrm{sin}^{k2}\phi _1\mathrm{}`$ (16) $`\mathrm{}{\displaystyle _0^\pi }d\phi _i\mathrm{sin}^{ki1}\phi _i\mathrm{}{\displaystyle _0^{2\pi }}d\phi _{k1}\{[\mathrm{sin}^2\phi _1`$ $`\times (1+\mathrm{sin}^2\phi _2(1+\mathrm{}(1+\mathrm{sin}^2\phi _{k1})))]^n\}.`$ It is worth noting that $`K_k(r;s)`$ is unique, which follows from the fact that $`K_k(r;s)`$ is the phase-space function of $`\widehat{E}^k`$ and from the uniqueness of phase-space representations. This is in contrast to the kernel functions that relate quantities to the quadrature-component statistics measured in balanced homodyne scheme, where certain functions can be added to the kernels without changing the result . The integral in Eq. (LABEL:KERNEL) converges for $`s`$ $`>`$ $`1`$ because $`|\mathrm{\Omega }^{(k)}(\rho ^2)|<e^{\rho ^2}V_k,`$ (17) $`V_k`$ being some constant. Plots of the kernel function for different values of $`s`$ and $`k`$ are shown in Fig. 1. We can see that $`K_k(r;s)`$ monotonically increases with $`r`$ from zero to one for $`s`$ $``$ $`0`$. If $`s`$ $`<`$ $`0`$, then $`K_k(r;s)`$ attains the maximum at a finite value of $`r`$. The position of the maximum shifts towards the origin and the value of the maximum tends to infinity as $`s`$ $``$ $`1`$. Hence the kernels that relate the exponential phase moments to the $`Q`$-function diverge at $`r`$ $`=`$ $`0`$. To be more specific, it can be shown (Appendix B) that $`K_k(r;1)r^k`$ (18) near the origin. Though the function $`K_k(r)`$ $``$ $`K_k(r;1)`$ diverges, it can be used to obtain the exponential phase moments $`\mathrm{\Psi }_k`$ from the $`Q`$-function $`Q(r,\phi )`$ $`=`$ $`W(r,\phi ,1)`$. It is not difficult to prove that Eq. (5) can be rewritten as $$\mathrm{\Psi }_k=_0^{\mathrm{}}r𝑑rQ_k(r)K_k(r),$$ (19) where $`Q_k(r)={\displaystyle _0^{2\pi }}𝑑\phi e^{ik\phi }Q(r,\phi )`$ (21) $`=\mathrm{\hspace{0.17em}2}e^{r^2}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{r^{2n+k}}{\sqrt{n!(n+k)!}}}\rho _{n+k,n}`$ ($`\rho _{n+k,n}`$ $`=`$ $`n`$ $`+`$ $`k|\widehat{\varrho }|n`$). It follows from Eq. (21) that $`Q_k(r)`$ $``$ $`r^k`$ for small $`r`$, and thus $`Q_k(r)`$ exactly compensates for the divergence of $`K_k(r)`$, Eq. (18). In other words, if the $`Q`$-function of the state is known exactly, then the integration in (19) can be performed straightforwardly, thus yielding the sought $`\mathrm{\Psi }_k`$. However, measurement of the $`Q`$-function is always associated with some error, so that the region close to the origin of the phase space needs careful consideration in praxis. ## III Canonical phase from double homodyning ### A Statistical error Let us consider balanced double-homodyne detection (Fig. 2) and first assume perfect detection. Each experimental event then gives a pair of real numbers which, after rescaling, define a point in the phase space of the signal, and the probability density of detecting the space points is equal to the $`Q`$-function of the signal state . When the $`j`$th measurement yields the phase-space point with polar coordinates $`(r_j,\phi _j)`$ and altogether $`N`$ measurements are performed, then the exponential phase moments can be estimated to be $`\mathrm{\Psi }_k^{(\mathrm{est})}={\displaystyle \frac{1}{N}}{\displaystyle \underset{j=1}{\overset{N}{}}}\mathrm{exp}(k\phi _j)K_k(r_j).`$ (22) In order to answer the question of how close is $`\mathrm{\Psi }_k^{(\mathrm{est})}`$ to the actual moment $`\mathrm{\Psi }_k`$, we calculate the mean value and dispersion of the estimate (22) over all possible measurement results. Since individual measurement outcomes are independent of each other, we can take advantage of the summation rule for mean values and dispersions of independent quantities. Thus, for the real part of $`\mathrm{\Psi }_k^{(\mathrm{est})}`$ we get the mean value $`E\left(\mathrm{Re}\mathrm{\Psi }_k^{(\mathrm{est})}\right)={\displaystyle \frac{1}{N}}{\displaystyle \underset{j=1}{\overset{N}{}}}E\left[\mathrm{cos}(k\phi _j)K_k(r_j)\right]`$ (25) $`={\displaystyle \frac{1}{N}}{\displaystyle \underset{j=1}{\overset{N}{}}}{\displaystyle _{2\pi }}𝑑\phi _j{\displaystyle _0^{\mathrm{}}}r_j𝑑r_j\mathrm{cos}(k\phi _j)K_k(r_j)Q(r_j,\phi _j)`$ $`={\displaystyle \frac{1}{N}}N\mathrm{Re}\mathrm{\Psi }_k=\mathrm{Re}\mathrm{\Psi }_k,`$ as it should be, and the dispersion $`D\left(\mathrm{Re}\mathrm{\Psi }_k^{(\mathrm{est})}\right)={\displaystyle \frac{1}{N^2}}{\displaystyle \underset{j=1}{\overset{N}{}}}D\left[\mathrm{cos}(k\phi _j)K_k(r_j)\right]`$ (28) $`={\displaystyle \frac{1}{N}}\{{\displaystyle _{2\pi }}d\phi {\displaystyle _0^{\mathrm{}}}rdr\mathrm{cos}^2(k\phi )K_k^2(r)Q(r,\phi )`$ $`\left[{\displaystyle _{2\pi }}d\phi {\displaystyle _0^{\mathrm{}}}rdr\mathrm{cos}(k\phi )K_k(r)Q(r,\phi )\right]^2\}.`$ Similar expressions hold for the imaginary part of $`\mathrm{\Psi }_k^{(\mathrm{est})}`$. Since cos$`{}_{}{}^{2}(k\phi )`$ $`=`$ $`1/2`$ $`+`$ cos$`(2k\phi )/2`$, after performing the angular integration in the first term on the right-hand side of Eq. (28), the radial part contains the product $`Q_0(r)K_k^2(r)`$ so that this integral over the divergent kernel can become infinite. Let $`n_0`$ be the number of photons at which the Fock expansion of the state starts. Taking into account Eq. (21), we see that the integrand behaves as $``$ $`r^{2(n_0k)+1}`$ for small $`r`$. Thus, the exponential phase moments $`\mathrm{\Psi }_k`$ can be directly sampled from the double-homodyne data, provided that $`k`$ $`<`$ $`n_0`$ $`+`$ $`1`$, because in this case the dispersion of the estimation is bounded. In the opposite case of $`k`$ $``$ $`n_0`$ $`+`$ $`1`$, the statistical fluctuation diverges so that the exponential phase moments cannot be sampled without a proper regularization of the kernels. Note that for states that contain the vacuum, regularization of the kernels is necessary for all exponential phase moments. ### B Kernel regularization and sampling algorithm Since the main part of the statistical error arises from data close to the origin, it is natural to modify the procedure by omitting the data falling inside a small circle $`r`$ $`<`$ $`r_0`$ (see Fig. 3). Of course, such a deliberate data filtering introduces into the measurement a state-dependent systematic error. Nevertheless, the statistical error is reduced and the total error may be acceptable. Replacing $`K_k(r)`$ by the regularized function $`K_k^{}(r)`$ according to $$K_k^{}(r)=\theta (rr_0)K_k(r)$$ (29) \[$`\theta (x)`$, Heaviside step function\], the systematic error of the $`k`$th moment can be given by $`\sigma _k^{(\mathrm{sys})}`$ $`=`$ $`{\displaystyle _0^{2\pi }}𝑑\phi e^{ik\phi }{\displaystyle _0^{r_0}}r𝑑rQ(r,\phi )K_k(r)`$ (30) $`=`$ $`{\displaystyle _0^{r_0}}r𝑑rQ_k(r)K_k(r).`$ (31) A measure of the total error is then the sum of the statistical and systematic errors, $`\mathrm{Re}\sigma _k^{(\mathrm{tot})}`$ $`=`$ $`\left|\mathrm{Re}\sigma _k^{(\mathrm{sys})}\right|+\left[D\left(\mathrm{Re}\mathrm{\Psi }_k^{(\mathrm{est})}\right)\right]^{1/2},`$ (32) and $`\mathrm{Im}\sigma _k^{(\mathrm{tot})}`$ accordingly. From the example in Fig. 4 it is seen that the statistical error decreases with increasing radius $`r_0`$ \[Fig. 4(a)\], whereas the systematic error increases with the radius \[Fig. 4(b)\]. The total error has thus a minimum at a certain radius \[Fig. 4(c)\], which can be regarded as the optimal radius for regularization. Unfortunately, the determination of the systematic error requires knowledge of the state. Nevertheless, an upper bound of the systematic error can be estimated, without any a priori knowledge of the measured state. Assuming $`r`$ $`<`$ $`1`$, we may write $`\left|Q_k(r)\right|=2{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{r^{2n+k}e^{r^2}}{\sqrt{n!(n+k)!}}}\left|\rho _{n+k,n}\right|`$ (35) $`2{\displaystyle \frac{r^ke^{r^2}}{\sqrt{k!}}}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\sqrt{\rho _{n,n}\rho _{n+k,n+k}}`$ $`2{\displaystyle \frac{r^ke^{r^2}}{\sqrt{k!}}}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\rho _{n,n}+\rho _{n+k,n+k}}{2}}2{\displaystyle \frac{r^ke^{r^2}}{\sqrt{k!}}},`$ where we have used the inequality $`|\rho _{m,n}|^2`$ $``$ $`\rho _{mm}\rho _{nn}`$ implied by positive definiteness of $`\rho `$. Hence, an upper bound of the systematic error can be estimated. Using Eq. (31), we find that $$|\sigma _k^{(\mathrm{sys})}|\frac{2}{\sqrt{k!}}_0^{r_0}𝑑rr^{k+1}e^{r^2}K_k(r).$$ (36) A typical state for which $`|\sigma _k^{(\mathrm{sys})}|`$ is of the order of magnitude of upper-bound value is $`|\psi _k`$ $`=`$ $`(|0`$ $`+`$ $`|k)/\sqrt{2}`$. For this state, $`Q_k(r)`$ $`=`$ $`r^k\mathrm{exp}(r^2)/\sqrt{k!}`$, which yields one half of the upper bound value. Taking into account that $`K_k(r)`$ $``$ $`1/r^k`$ for $`r`$ $``$ $`1`$, we find from the inequality (36) that the upper bound of $`|\sigma _k^{(\mathrm{sys})}|`$ increases quadratically with $`r_0`$. The dependence on $`r_0`$ of the upper bound of the $`|\sigma _k^{(\mathrm{sys})}|`$ is shown in Fig. 5. Notice that the systematic error is smaller for higher $`k`$. The state-independent upper bound of the systematic error and the estimated statistical error can now be used to determine the upper bound of the total error. Its minimum then determines an appropriate regularization radius $`r_0`$. A possible algorithm for optimized data processing is the following one. In the zeroth step, sampling of the desired exponential phase moments from all $`N`$ measurement events is performed. Since also data with very small $`r`$ may contribute to the result, the statistical error can be very large. In contrast to standard sampling technique, where there is no need for data storage, here the data within a certain small circle are stored. The radius of the circle should be slightly larger than the expected regularization radius. The regularized kernel function (29) is now used, with $`r_0`$ being increased step by step, so that in the $`n`$th step $`n`$ events closest to the origin are covered by $`r_0`$. In each step statistical and systematic errors are estimated. The value of $`r_0`$ for which the total error is minimized is used for calculation of the final result. Let us mention that the detrimental effect of divergent kernels $`K_k(r)`$ at $`r`$ $`=`$ $`0`$ (in connection with nonzero $`Q`$-function) resembles the experiment in , where the statistics of sine and cosine phases are obtained from low-efficiency double homodyning. In the experiment, data giving rise to divergences are disregarded, which is criticized in from the argument that the disregarded data represent an extra noise in the statistics. In our case, we disregard data leading to high statistical error and include the resulting systematic error into the sampling scheme. ### C Total error and number of measurements Let us assume that a particular phase moment $`\mathrm{\Psi }_k`$ is desired to be determined with a prescribed total precision $`\sigma _k^{(\mathrm{tot})}`$. What is the necessary number of measurement events $`N`$? If there were no need for regularization and the precision were limited only by (finite) statistical fluctuation, then $`N`$ $``$ $`(\sigma _k^{(\mathrm{tot})})^2`$. When the vacuum contributes to the state to be measured and a regularization radius $`r_0`$ is introduced, then the total error reads $`\sigma _1^{(\mathrm{tot})}`$ $`=`$ $`A_1(\mathrm{ln}r_0)^{1/2}N^{1/2}+B_1r_0^2,`$ (37) $`\sigma _k^{(\mathrm{tot})}`$ $`=`$ $`A_kr_0^{1k}N^{1/2}+B_kr_0^2,k2,`$ (38) where $`A_k`$ and $`B_k`$ are constants. The optimal regularization radius $`r_0^{(\mathrm{opt})}`$, which minimizes the total error (38) depends on $`N`$ as $$r_0^{(\mathrm{opt})}N^{1/[2(1+k)]}.$$ (39) From this expression and Eq. (38) we find that $`\sigma ^{(\mathrm{tot})}N^{1/(1+k)},`$ (40) i.e., $`N\left(\sigma _k^{(\mathrm{tot})}\right)^{(1+k)}`$ (41) ($`k`$ $``$ $`2`$). The case $`k`$ $`=`$ $`1`$ needs separate consideration, because of the logarithm, which does not provide us with a simple analytical expression. Obviously, $`N`$ increases faster than ($`\sigma _1^{(\mathrm{tot})})^2`$ with decreasing error. Thus, we can see that in the limit of small total error ordinary homodyning (which does not require regularization) is better suitable for sampling exponential phase moments than the double homodyning, because it requires a smaller amount of data to achieve the same precision. ### D Computer simulation To demonstrate the feasibility of the method, we have performed Monte Carlo simulations of double-homodyne detection of the $`Q`$-function for sampling the exponential phase moments of a coherent state. Results are shown in Fig. 6 and Table I. From Fig. 6 and Table I it is seen that the sampled exponential phase moments are in good agreement with the exact ones. Note the strong increase of the error with the index $`k`$ of the moment (for a detailed error analysis, see Fig. 4). Further, a comparison between the dashed and undashed bars in Fig. 6 clearly shows the difference between the concept of canonical phase and the phase concept based on the radially integrated $`Q`$-function. In order to compare double homodyning with ordinary homodyning, we have also simulated homodyne detection of the quadrature-component statistics for sampling the exponential phase moments of the same coherent state as in the simulated double-homodyne experiment, using the method in . The results are presented in Table II. Comparing Tables I and II, we see that (for equal total numbers of events) the error in ordinary homodyning is indeed smaller than in double homodyning. The difference between the errors observed in the two schemes increases with increasing index $`k`$ of the moment. Note that for $`k`$ $`=`$ $`4`$ the error in the double-homodyne scheme is ten times larger than in the ordinary homodyne measurement. ### E Imperfect photodetection In a real experiment, the overall detection efficiency $`\eta `$ would be always smaller than $`100\%`$, but it can be very high, e.g., $`\eta `$ $`=`$ $`99\%`$. Nonperfect detection introduces additional noise into the sampling scheme and gives rise to an additional systematic error, which cannot be diminished by increasing the number of measurements. The effect of nonperfect detection is that the exponential phase moments of a “smoothed” quantum state are sampled rather than those of the true one. Since the additional noise is Gaussian, the phase-space function that is actually recorded is not the $`Q`$-function but the function $`W_{12\eta ^1}(q,p)`$. The sampling of $`\mathrm{\Psi }_k`$ from this function by means of the kernel function $`K_k(r;1)`$ is equivalent to sampling of $`\mathrm{\Psi }_k`$ from the $`Q`$-function by means of the kernel function $`K_k[r;1+2(\eta ^11)]`$. For a given quantum state, the systematic error can thus be given by $$\mathrm{\Delta }_\eta \mathrm{\Psi }_k=_0^{\mathrm{}}r𝑑rQ_k(r)\left[K_k(r;1)K_k(r;3+2\eta ^1)\right].$$ (42) Its result is the underestimation of the magnitude of the moment. Since the difference of the kernel functions is essentially nonzero only around the origin $`r`$ $`=`$ $`0`$, the systematic error $`\mathrm{\Delta }_\eta \mathrm{\Psi }_k`$ will be highest for states close to vacuum. After a proper regularization, one can use Eq. (42) to get a reasonable estimation of the error by substituting the measured statistics for the unknown $`Q`$-function. ## IV Other measurements of the phase-space functions The phase moments $`\mathrm{\Psi }_k`$ can be also obtained from quasidistributions reconstructed in unbalanced homodyning , or cavity measurements . However, these methods do not yield $`W_s(q,p)`$ as statistics of events $`(q,p)`$, but the functions $`W_s(q,p)`$ are determined pointwise. The restriction in practice to a selected finite number of points necessarily results in a systematic error, because the integration over the phase space is replaced by summation over a finite number of points selected by the experimentalist. Having determined $`W_s(q,p)`$, the exponential phase moments $`\mathrm{\Psi }_k`$ can then be reconstructed from $`W_s(q,p)`$ on the basis of Eq. (5). Since the kernel function $`K_k(r;s)`$ is well behaved for $`s`$ $`>`$ $`1`$, no problems with divergences arise here. In unbalanced homodyning displaced Fock-state distributions $`p(n,\alpha )`$ are measured . It can be expected that the cumbersome way of reconstructing $`\mathrm{\Psi }_k`$ from $`p(n,\alpha )`$ via $`W_s(q,p)`$ may be avoided and the reconstruction can be performed directly from the measured data. This could be done in a similar way as in the reconstruction of the density matrix in the Fock basis . A similar approach can be used for different physical systems: statistics of the displaced Fock states of vibrating trap ions has been obtained in state-reconstruction experiments , and schemes based on displaced Fock statistics of the cavity fields have been suggested . In particular, the scheme of directly yields the Wigner function, from which the exponential phase moments can be obtained in a very straightforward way. Even though many interesting problems are related to these schemes, we will not deal with them in this paper in any more detail. ## V Phase distribution In order to answer the question of the possibility of direct sampling of the phase distribution itself, we have first to answer the question of the existence of kernels $`F(r,\phi \psi ;s)`$ such that $$P(\phi )=_0^{2\pi }𝑑\psi _0^{\mathrm{}}r𝑑rW(r,\psi ;s)F(r,\phi \psi ;s).$$ (43) Obviously, $`F(r,\psi \phi ;s)`$ is the ($`s`$)-parametrized phase-space function of the phase state $`|\phi `$ in Eq. (2). In it is shown that this function can be given by $$F(r,\phi ;s)=\underset{m,n}{}B_{m,n}(r,s)e^{i(mn)(\phi )},$$ (44) where $`B_{mn}(r,s)=\sqrt{{\displaystyle \frac{n!}{m!}}}r^{mn}\left({\displaystyle \frac{s1}{2}}\right)^n\left({\displaystyle \frac{2}{1+s}}\right)^{m+1}`$ (46) $`\times L_n^{mn}\left({\displaystyle \frac{4r^2}{1s^2}}\right)\mathrm{exp}\left({\displaystyle \frac{2r^2}{1+s}}\right)`$ for $`m`$ $``$ $`n`$, and $`B_{nm}`$ $`=`$ $`B_{mn}`$. The series (44) only converges for $`s`$ $`>`$ $`0`$ (for the limiting case $`s`$ $`=`$ $`0`$, see also ). Let us express $`F(r,\phi ;s)`$ for $`s`$ $`>`$ $`0`$ in terms of $`K(r;s)`$. From Eq. (3) it follows that $$P(\phi )=\frac{1}{2\pi }\underset{k=\mathrm{}}{\overset{\mathrm{}}{}}\mathrm{\Psi }_ke^{ik\phi }.$$ (47) Combining Eqs. (47) and (5) and recalling Eq. (43), we may write $`F(r,\phi \psi ;s)`$ in the form $$F(r,\phi ;s)=\frac{1}{2\pi }\left[1+2\underset{k=1}{\overset{\mathrm{}}{}}K_k(r;s)\mathrm{cos}(k\phi )\right].$$ (48) When $`r`$ $``$ $`\mathrm{}`$ then $`K_k(r;s)`$ $``$ $`1`$, and thus $`F(r,\phi ,s)`$ $``$ $`\delta (\phi )`$. Regrouping the terms in Eqs. (A1) and (A2) (for $`s`$ $`>`$ $`0`$) and using a summation formula for Laguerre polynomials , we can rewrite $`K_k(r;s)`$ as $`K_k(r;s)=r^k\left({\displaystyle \frac{2}{s+1}}\right)^{k+1}e^{2r^2/(1+s)}`$ (50) $`\times {\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(1)^n}{\sqrt{(n+1)\mathrm{}(n+k)}}}\left({\displaystyle \frac{1s}{1+s}}\right)^nL_n^k\left({\displaystyle \frac{4r^2}{1s^2}}\right),`$ which is suitable for computing $`F(r,\phi ;s)`$. An example is displayed in Fig. 7. The fact that for a large class of states $`W(r,\phi ;s)`$ does not exist as a regular function for $`s`$ $`>`$ $`0`$ limits the applicability of Eq. (43). Nevertheless, there exist states for which $`W(r,\phi ;s)`$ for $`s`$ $`>`$ $`0`$ is a regular function which can be sampled using unbalanced homodyning. However, for $`s`$ $`>`$ $`0`$ the statistical error of the sampled $`W(r,\phi ;s)`$ increases with $`r`$ , so that application of Eq. (43) requires special regularization. ## VI Summary and Conclusion The main results can be summarized as follows. $`(i)`$ There exist well-behaved kernels for sampling the exponential moments of the canonical phase from $`s`$-parametrized phase-space functions for $`s`$ $`>`$ $`1`$. For $`s`$ $`=`$ $`1`$ the kernels diverge in the origin, and for $`s`$ $`<`$ $`1`$ the kernels do not exist as regular functions. $`(ii)`$ Even though for $`s`$ $`=`$ $`1`$ the kernels diverge, their integral with the $`Q`$-function is finite, so that they may be used for inferring the exponential phase moments from the exact $`Q`$-function. However, the kernel divergence may cause divergent errors of some moments for some states if fluctuating experimental data are used. $`(iii)`$ Finite errors can be obtained if regularized kernels are used. Since regularization introduces a systematic error, an optimization procedure should be used in order to minimize the sum of the statistical and systematic errors. $`(iv)`$ The fact that the canonical phase moments can be sampled in double homodyning has an interesting interpretation. Each measurement event yields a unique phase value, but these values must be taken with different weights in dependence on the distance from the origin of the phase space. This is in contrast to the ordinary (four-port) homodyning, where a single measurement does not provide us with a phase value. $`(v)`$ Even if optimally regularized kernels are used, the amount of data necessary for realizing a desired precision is larger than in standard sampling. This is a disadvantage of the double-homodyne scheme in comparison with ordinary homodyning. $`(vi)`$ In double homodyning, correct results require perfect detection, because there is no simple possibility of compensation for detection losses, which cause an additional systematic error. This is another disadvantage of double homodyning compared to ordinary homodyning where a compensation of imperfect detection is possible for efficiencies down to $`\eta `$ $`>`$ $`0.5`$. $`(vii)`$ Thus, in reply to the question posed in the Introduction, it does not seem that phase-space measurements based on double-homodyning are closer to canonical-phase measurement than quadrature-component measurements based on ordinary homodyning. $`(viii)`$ In contrast to ordinary homodyning however, the sampling functions in double homodyning are uniquely defined. This follows from the fact that they are actually phase-space representations of quantum mechanical operators. $`(ix)`$ The exponential phase moments can also be inferred from the data recorded in other schemes such as unbalanced homodyning, in which $`s`$-parametrized phase-space functions are reconstructed pointwise. $`(x)`$ Kernels for sampling the distribution of the canonical phase exist as regular functions only for $`s`$ $`>`$ 0. Even though for some states the corresponding phase space functions exist and can be measured using unbalanced homodyning, the behavior of the statistical error would require a special regularization of the scheme to be applicable. ###### Acknowledgements. We thank J. Peřina for stimulating discussions and acknowledge discussions with J. Clausen. J.F. acknowledges support of the US-Israel Binational Science Foundation. This work was supported by the Deutsche Forschungsgemeinschaft. ## A Sampling kernels and the function $`\mathrm{\Omega }`$ Substituting Eq. (9) into Eq. (7) and comparing with Eq. (5), we can express $`K_k(r;s)`$ as $$K_k(r;s)=r^k\underset{l=0}{\overset{\mathrm{}}{}}\left(\frac{1s}{2}\right)^lC_l^{(k)}L_l^k\left(\frac{2r^2}{1s}\right),$$ (A1) where the coefficients $`C_l^{(k)}={\displaystyle \underset{n=0}{\overset{l}{}}}{\displaystyle \frac{(1)^n}{\sqrt{(n+1)\mathrm{}(n+k)}}}\left({\displaystyle \genfrac{}{}{0pt}{}{l}{n}}\right),`$ (A2) can be rewritten as $$C_l^{(k)}=\frac{1}{\pi ^{k/2}}_{\mathrm{}}^{\mathrm{}}𝑑t_1e^{t_1^2}\mathrm{}_{\mathrm{}}^{\mathrm{}}𝑑t_ke^{kt_k^2}z_k^l,$$ (A3) with $$z_k=1e^{\rho _k^2},\rho _k^2=\underset{j=1}{\overset{k}{}}t_j^2.$$ (A4) Substituting this expression into Eq. (A1) and using the summation rule $`{\displaystyle \underset{l=0}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{1s}{2}}z_k\right)^lL_l^k\left({\displaystyle \frac{2r^2}{1s}}\right)`$ (A6) $`=\left(1z_k{\displaystyle \frac{1s}{2}}\right)^{k1}\mathrm{exp}\left[{\displaystyle \frac{z_kr^2}{z_k(1s)/21}}\right],`$ we arrive at $`K_k(r;s)={\displaystyle \frac{r^k2^{k+1}}{\pi ^{k/2}}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}dt_1\{e^{t_1^2}\mathrm{}{\displaystyle _{\mathrm{}}^{\mathrm{}}}dt_ke^{kt_k^2}`$ (A8) $`\times [1+s+(1s)e^{\rho _k^2}]^{k1}\mathrm{exp}[{\displaystyle \frac{2(1e^{\rho _k^2})r^2}{1+s+(1s)e^{\rho _k^2}}}]\}`$ Note that the series in Eq. (A6) is only convergent for $`|z_k(1s)/2|<1`$. We have $`z_k`$ $``$ $`1`$, thus $$|(1s)/2|<1s>1$$ (A10) must hold so that the $`Q`$-function ($`s`$ $`=`$ $`1)`$ represents limiting case for sampling the phase moments. The multiple integration in Eq. (LABEL:K) can be conveniently performed in hyperspherical coordinates. For this purpose we introduce the function $`\mathrm{\Omega }^{(k)}(\rho ^2)`$ , $`\mathrm{\Omega }^{(k)}(\rho ^2)={\displaystyle _0^\pi }d\phi _1[\mathrm{sin}^{k2}\phi _1\mathrm{}{\displaystyle _0^\pi }d\phi _i\mathrm{sin}^{ki1}\phi _i`$ (A12) $`\times \mathrm{}{\displaystyle _0^{2\pi }}d\phi _{k1}\mathrm{exp}({\displaystyle \underset{l=1}{\overset{k}{}}}lt_l^2)],`$ where $`t_i=\rho \mathrm{cos}\phi _i{\displaystyle \underset{j=1}{\overset{i1}{}}}\mathrm{sin}\phi _j\mathrm{if}i<k,`$ (A13) and $`t_k=\rho {\displaystyle \underset{j=1}{\overset{k1}{}}}\mathrm{sin}\phi _j,`$ (A14) with $`\rho `$ $`=`$ $`\rho _k`$. The exponent can be expressed in hyperspherical coordinates as $`{\displaystyle \underset{l=1}{\overset{k}{}}}lt_l^k=\rho ^2`$ (A16) $`+\rho ^2\mathrm{sin}^2\phi _1(1+\mathrm{sin}^2\phi _2(1+\mathrm{}(1+\mathrm{sin}^2\phi _{k1}))).`$ Inserting this expression into Eq. (A12) and expanding the exponential function into a Taylor series, we arrive at Eq. (13) together with Eq. (16). A recurrence formula for the coefficients $`A_n^{(k)}`$ in Eq. (16) can be readily obtained: $$A_n^{(k)}=B_{2n+k2}\underset{l=0}{\overset{n}{}}\left(\genfrac{}{}{0pt}{}{n}{l}\right)A_l^{(k1)},k3$$ (A17) where $$B_j=_0^\pi 𝑑\phi \mathrm{sin}^j\phi =\sqrt{\pi }\frac{\mathrm{\Gamma }(\frac{j+1}{2})}{\mathrm{\Gamma }(\frac{j+2}{2})},(j0).$$ (A18) Starting from $`A_n^{(2)}`$ $`=`$ $`2B_{2n}`$, the formulas (A17) and (13) allow for fast and accurate numerical determination of the functions $`\mathrm{\Omega }^{(k)}(\rho ^2)`$ even for high $`k`$. ## B Asymptotics of $`\mathrm{\Omega }^{(\mathrm{k})}(\rho ^2)`$ and divergence of kernels $`K_\mathrm{k}(r)`$ In order to analyze the divergence of the kernels $`K_k(r)`$, we must first know the asymptotic behavior of the functions $`\mathrm{\Omega }^{(k)}(\rho ^2)`$ for large $`\rho `$. We start from the integral representation (A12) and write the exponent (A16) as $$\underset{l=1}{\overset{k}{}}lt_l^k=\rho ^2+\rho ^2\mathrm{sin}^2\phi _1\mathrm{\Phi }(\phi _2,\mathrm{},\phi _{k1}),$$ (B1) with $$\mathrm{\Phi }(\phi _2,\mathrm{},\phi _{k1})=1+\mathrm{sin}^2\phi _2(1+\mathrm{sin}^2\phi _3(1+\mathrm{})).$$ (B2) Note that $`\mathrm{\Phi }`$ $``$ $`1`$. We insert Eq. (B1) into Eq. (A12) and integrate over $`\phi _1`$. Assuming $`k2`$, the relevant integral is $`I`$ $`=`$ $`{\displaystyle _0^\pi }𝑑\phi _1\mathrm{sin}^{k2}\phi _1e^{\rho ^2\mathrm{\Phi }\mathrm{sin}^2\phi _1}`$ (B3) $`=`$ $`2{\displaystyle _0^{\frac{\pi }{2}}}𝑑\phi _1\mathrm{sin}^{k2}\phi _1e^{\rho ^2\mathrm{\Phi }\mathrm{sin}^2\phi _1}.`$ (B4) Assuming $`\rho ^2`$ $``$ $`1`$, we may write $`\mathrm{sin}\phi _1`$ $``$ $`\phi _1`$, because the integrand is essentially nonzero only for $`\phi _1`$ $``$ $`1`$. From the same argument, we can extend the integration from $`(0,\pi /2)`$ to $`(0,\mathrm{})`$, $$I2_0^{\mathrm{}}𝑑\phi _1\phi _1^{k2}e^{\rho ^2\mathrm{\Phi }\phi _1^2}=\mathrm{\Gamma }\left(\frac{k1}{2}\right)(\rho ^2\mathrm{\Phi })^{\frac{1k}{2}}.$$ (B5) To finish the calculation of $`\mathrm{\Omega }`$, one has to integrate $`\mathrm{\Phi }^{(1k)/2}`$ over the remaining angles $`\phi _2,\mathrm{},\phi _{k1}`$ (with appropriate measure). We eventually find the asymptotic behavior $$\mathrm{\Omega }^{(k)}(\rho ^2)C_k\rho ^{1k}e^{\rho ^2}.$$ (B6) The factor $`\mathrm{exp}(\rho ^2)`$ comes from the first $`\rho ^2`$ in Eq. (B1). Taking into account that $`\mathrm{\Omega }^{(1)}(\rho ^2)`$ $`=`$ $`2\mathrm{exp}(\rho ^2)`$, we see that the asymptotic behavior (B6) holds for all $`k1`$. To investigate the divergence of $`K_k(r)`$ at $`r0`$, we make use of the integral representation (LABEL:KERNEL), which is rewritten here as $`K_k(r)={\displaystyle \frac{r^k}{\pi ^{k/2}}}{\displaystyle _0^{\mathrm{}}}𝑑\rho \rho ^{k1}\mathrm{\Omega }^{(k)}(\rho ^2)`$ (B8) $`\times e^{(k+1)\rho ^2}\mathrm{exp}\left[(e^{\rho ^2}1)r^2\right].`$ For small $`r`$, the dominant contribution to this integral comes from large $`\rho `$. We can replace $`\mathrm{\Omega }`$ by the asymptotic formula (B6) and absorb the prefactors into $`C_k`$, $$K_k(r)C_kr^k_0^{\mathrm{}}𝑑\rho e^{k\rho ^2}\mathrm{exp}\left[(e^{\rho ^2}1)r^2\right].$$ (B9) Change of the variable according to $$t=(e^{\rho ^2}1)r^2$$ (B10) yields $$K_k(r)\frac{1}{2r^k}_0^{\mathrm{}}𝑑t\left(\mathrm{ln}\frac{t+r^2}{r^2}\right)^{1/2}(t+r^2)^{k1}e^t.$$ (B11) The integration region of (B11) can be divided in two parts, $`t`$ $`<R^2`$ and $`t`$ $`>R^2`$, with $`R`$ $``$ $`r`$. For small $`r`$, the dominant contribution stems from the latter part where the approximation $$\mathrm{ln}\frac{t+r^2}{r^2}\mathrm{ln}\frac{1}{r^2}$$ (B12) can be used, and we find that $$K_k(r)\frac{(2\mathrm{ln}r)^{1/2}}{2r^k}_0^{\mathrm{}}𝑑t(t+r^2)^{k1}e^t.$$ (B13) The integral is finite for $`r`$ $``$ $`0`$, and thus $$K_k(r)\frac{r^k}{(\mathrm{ln}r)^{1/2}}.$$ (B14) The logarithm singularity in the denominator is very weak in comparison to the polynomial one in the numerator. Only the polynomial divergence is relevant in Eq. (5), because it determines whether the integral is convergent or not. Thus we need not consider the logarithmic part, so that Eq. (B14) simplifies to $$K_k(r)r^k.$$ (B15)
warning/0004/cond-mat0004176.html
ar5iv
text
# Critical behavior of the 𝑆=1/2 Heisenberg ferromagnet: A Handscomb quantum Monte Carlo study ## I introduction The slowing down of the relaxation to thermal equilibrium is an important physical phenomenon associated to the build up of long-range correlations at the critical point of spin systems. According to the finite size scaling hypothesis, the critical relaxation time scales with the system size as $`\tau L^z`$, with the dynamical critical exponent $`z`$ governing the rate of convergence towards thermal equilibrium. The value of $`z`$ depends on the particular equation of motion of the order parameter and of the conservation laws that apply to the spin system. In particular, the dynamic scaling theory predicts that for the classical isotropic Heisenberg model with conserved order parameter $`z=d\beta /\nu `$ ($`z2.5`$ for $`d=3`$), which is consistent with both experiments and numerical simulations. For a recent review see Ref. Monte Carlo simulation is a powerful tool to study the equilibrium properties of a spin system through some stochastic relaxational dynamics. It defines a Markovian process in which the associated stochastic model evolves in the phase space according to certain transition probabilities. The transition rates between two distinct spin configurations are imposed to satisfy the detailed balance condition in order to lead the system to its equilibrium state distribution. The dynamical evolution of the stochastic model can be thought of resulting from the interactions among its many degrees of freedom. However, the evolution of the model-system under the Monte Carlo dynamics do not need to be consistent with any motion equation. In this context, the critical exponent $`z`$ characterizes therefore the rate at which a particular set of stochastic dynamical rules generates uncorrelated spin configurations. A general feature of Monte Carlo simulations of classical spin models is that cluster dynamics are usually much more effective in overcoming the critical slowing down than those which are based on single spin-flip procedures. For example, the cluster Monte Carlo algorithm introduced by Swendsen-Wang is known to have a very fast convergence to equilibrium at the critical point in contrast to the slower convergence of local dynamics such as Metropolis and heat-bath. For the isotropic classical Heisenberg model it has been found that $`z2`$ for Metropolis and $`z0`$ for the single cluster Monte Carlo dynamics. The traditional spin-flip Monte Carlo algorithms as applied to classical spin systems can not be directly extended to quantum spin models. The problem resides in the fact that the Hamiltonian is not, in general, diagonal in the spin configuration basis. A quantum Monte Carlo method was first introduced by Handscomb for the calculation of the thermodynamical properties of quantum Heisenberg ferromagnets. The main difference of this technique in comparison to traditional Monte Carlo algorithms is that the sample space is not related to any kind of physical phase space. It is the diagrammatic series of the partition function that is calculated by means of the Monte Carlo method. The Handscomb method has been successfully used to compute the thermodynamical properties of the Heisenberg $`S=1/2`$ ferromagnet and extended to a form applicable to a series of quantum spin models. Another algorithm commonly used in Monte Carlo simulations of quantum spin models is based on the use of the generalized Trotter formula to map the quantum system onto a classical system with an additional imaginary time dimension. The $`S=1/2`$ Heisenberg model has been extensively studied within this line. Recently, a decoupled cell method for quantum Monte Carlo based on the Suzuki-Trotter approach has been used to compute the critical dynamical exponent $`z`$ of the $`S=1/2`$ Heisenberg model on the simple cubic lattice. It was found that $`z2`$ which is quite similar to the value obtained from simulations of the classical Heisenberg model under Metropolis dynamics. In the present work, we are going to investigate both the static and dynamic critical properties of the $`S=1/2`$ Heisenberg ferromagnet on the simple cubic lattice by means of the Handscomb dynamics. We will employ a phenomenological renormalization group to obtain precise estimates of the critical temperature and some static critical exponents. By employing a so called moving block bootstrap (MBB) technique, we are going to calculate the equilibrium relaxation time for the energy and susceptibility at criticality. The critical time-displaced equilibrium correlation function will be computed as well. We will employ a finite size scaling analysis of the equilibrium relaxation time to obtain the critical dynamical exponent $`z`$ associated to the Handscomb dynamics and we contrast it with the results from the quantum decoupled cell method and the results from distinct Monte Carlo simulations of the classical Heisenberg model. ## II The Handscomb Monte Carlo method Let us briefly draw the main ideas of the Handscomb method. Consider the Hamiltonian of a quantum spin system to be given by $$H=\underset{i}{\overset{N_0}{}}H_i,[H_i,H_j]0$$ (1) The canonical average of a physical observable $`A`$ can be expanded in the form $$A=\frac{Tr[Aexp(\beta H)]}{Tr[exp(\beta H)]}=\underset{r}{}\underset{C_r}{}A(C_r)p(C_r)$$ (2) where $`\beta =1/k_BT`$, $`_{C_r}`$ denotes a summation over all ordered sets of indices $`C_r\{i_1,i_2,\mathrm{},i_r\}`$ (Mayer diagrams) and $$\begin{array}{c}A(C_r)\frac{Tr[AH_{i_1}\mathrm{}.H_{ir}]}{Tr[H_{i_1}\mathrm{}.H_{i_r}]}\hfill \\ p(C_r)\frac{\frac{(\beta )^r}{r!}Tr[H_{i_1}\mathrm{}.H_{i_r}]}{_r_{C_r}\frac{(\beta )^r}{r!}Tr[H_{i_1}\mathrm{}.H_{i_r}]}\hfill \end{array}$$ (3) Once $`p(C_r)0`$, it can be considered as a probability distribution and the canonical averages can be written as $`A=A(C_r)_{p(C_r)}`$. This is the case of the Heisenberg $`S=1/2`$ ferromagnet. The Hamiltonian can be represented in terms of transposition operators as $$H=JE_{i,j}$$ (4) so that the relevant trace to be computed is that of a permutation operator $$TrP(C_r)Tr[E_{(i,j)_1}E_{(i,j)_2}\mathrm{}E_{(i,j)_r}]=2^{k(C_r)}$$ (5) where $`k(C_r)`$ is the number of cycles in the irreducible representation of the permutation $`P(C_r)`$. It is straightforward to show that any physical observable can be expressed in terms of the diagram structure. For instance, the internal energy is related to the average number of transposition operators in the diagrams and the susceptibility to the average size of the cycles in the diagram’s irreducible representation . The Handscomb Monte Carlo method organizes a random walk in the space of the diagrams $`C_r`$ which has $`p(C_r)`$ as the limit distribution. The dynamics suggested by Handscomb consists of three types of steps: (i) Step forward, chosen with probability $`f_r`$, which tries to include a randomly chosen bond to the right of the permutation operator; (ii) Step backwards, chosen with probability $`1f_r`$, which tries to remove a bond from the left of $`P(C_r)`$; (iii) cyclic transposition, chosen when step backwards is rejected, which moves a bond from the left to the right. The transition probabilities for performing each movement on the space of Mayer’s diagrams are chosen in order to satisfy the detailed balance condition. After a single step of the Handscomb Monte Carlo dynamics, the irreducible representation of the sequency $`C_r`$ can have its cycle structure changed considerably. When two sites belong to distinct cycles, the insertion of the corresponding bond results in the coalescence of the two cycles of permutations. On the other hand, i.e., when the sites belong to the same cycle, the insertion breaks the cycle in two new ones. The same process occurs when a bond is removed from the sequence. Therefore, entire sets of sites can have their status changed during a single Monte Carlo step and, in this sense, the Handscomb dynamics is similar to the classical Monte Carlo cluster algorithms. ## III Finite size scaling for the susceptibility and order parameter The susceptibility per spin of the quantum $`S=1/2`$ Heisenberg model is written in terms of the cyclic structure of the irreducible representation of $`C_r`$ as $$\beta \chi =\frac{1}{N}\underset{j=1}{\overset{k(C_r)}{}}a_j^2_P$$ (6) where $`a_j`$ is the length of the $`j`$-th cycle of the permutation $`P(C_r)`$ and $`\mathrm{}_P`$ denotes an average with respect to the $`C_r`$-space probability distribution. In figure 1, we show our results for the susceptibility from lattices of $`L^3`$ spins with $`L=16,24`$ and, $`32`$. In these simulations $`150L^3`$ Monte Carlo steps (insertion, removal or bond permutation) were enough to let the system evolve to an equilibrium diagram configuration starting from an initial diagram containing no transpositions. After equilibrium was reached, we averaged over $`2\times 10^4`$ distinct diagrams, $`10^3`$ MCS apart. These results were averaged over $`10`$ distinct realizations of the numerical experiment. The susceptibility exhibits a critical behavior around $`k_BT_c/J1.68`$ in agreement with previous Monte Carlo estimates. For $`T>T_c`$, $`\chi `$ is only weakly dependent on the system size; whereas it is nearly proportional to $`L^3`$ at low temperatures. Notice that $`\chi `$ equals the magnetization second moment for temperatures below $`T_c`$ once the magnetization is strictly zero within the Handscomb prescription. In order to obtain a precise estimate of the critical temperature, we implemented a phenomenological renormalization group analysis of the data from finite size lattices as introduced by Nightingale. The basic assumption is that near the transition the susceptibility of a finite lattice of linear size $`L`$ scales as $$\chi (T,L)=L^{\gamma /\nu }f_\pm (tL^{1/\nu })$$ (7) where $`t=|(TT_c)/T_c|`$ and ($`\pm `$) stands for distinct scaling functions above and below $`T_c`$. The renormalization of temperature is defined by the following transformation relating lattices of two different sizes, $`L`$ and $`L^{}`$ $$\chi (T,L)=(L/L^{})^{\gamma /\nu }\chi (T^{},L^{})$$ (8) with the fixed point giving $`T_c`$. Then a set of auxiliary functions is introduced as $$g_\chi (T,L,L^{})=\frac{\mathrm{ln}[\chi (T,L)/\chi (T,L^{})]}{\mathrm{ln}(L/L^{})}$$ (9) and these intercept as a function of temperature at a common point from which we can directly measure $`T_c`$ and $`\gamma /\nu =g_\chi (T_c,L,L^{})`$. In figure 2 we plot the auxiliary functions $`g_\chi (T,L,L^{})`$ for typical renormalizations. Using all possible renormalizations with lattice sizes $`L=16,24,32,40`$ and $`48`$, we estimate $`k_BT_c/J=1.677\pm 0.001`$ and $`\gamma /\nu =1.98\pm 0.01`$. These values are one order of magnitude more accurate than the previous Monte Carlo estimates from simulations on small lattices ($`L10`$) which reported $`k_BT_c/J=1.68\pm 0.01`$. An even more accurate value for the critical temperature can be found by employing a renormalization study of critical quantities which are known to depict smaller fluctuations near the critical point such as the magnetization itself. Unfortunately, as we mention before, the magnetization is exactly zero for all temperatures due to an intrinsic symmetry of the Handscomb dynamics. However, we can explore the cycle structure of the Mayer diagrams to introduce a graph quantity which display the same critical behavior of the order parameter. In the simulations of classical spin models, such a quantity is the size of the largest cluster of spins which are in the same state. This might suggest that the largest cycle within a diagram in the context of Handscomb MC, may exhibit the same scaling behavior as the magnetization. Therefore, we will introduce a graph order parameter as the average size of the largest cycle of permutations. In figure 3, we plot the average size of the largest cycle (normalized by the total number of sites) as a function of temperature from simulations on lattices with $`L=16,24`$ and, $`48`$. From this figure, one can see that the average size of the largest cycle depicts an overall behavior similar to the one expected for an order parameter, and it will be considered as a true order parameter so forth. It also indicates a phase transition around $`k_BT_c/J1.68`$. A renormalization analysis performed on the order parameter data is shown in figure 4. From these data we found $`k_BT_c/J=1.6778\pm 0.0002`$ and $`\beta /\nu =0.512\pm 0.002`$. From the best of our knowledge, the presently reported values for $`k_BT_c/J`$, $`\gamma /\nu `$ and $`\beta /\nu `$ are the most accurate Monte Carlo estimates up to date for the quantum 3D Heisenberg ferromagnet. Our quoted value for $`T_c`$ is in complete agreement with the most accurate high-temperature series study which yielded $`J/k_BT_c=0.5960(5)`$ . The critical exponents are in excellent agreement with the best estimates for the classical Heisenberg ferromagnet. ## IV Critical Relaxation of the Spin 1/2 Heisenberg Model The critical relaxation within the Handscomb prescription can be investigated by computing some equilibrium time-displaced correlation functions $`C(t)`$ at the Curie temperature. We look at the equilibrium relaxation time $`\tau `$ which is expected to depict a power-law increase with the system size $`L`$ whose exponent characterizes the critical relaxation process. In particular, it governs the size dependence of the rate at which uncorrelated configurations are generated during the Monte Carlo temporal evolution in phase space. The fast growth of the relaxation time is referred to as the critical slowing down which may be governed by several relaxation modes. One generally is interested in the slower relaxation modes, i. e., the longest relaxation times. Therefore, it is safer to work with the integrated correlation time given by $$\tau _{int}=_0^{\mathrm{}}C(t)𝑑t.$$ (10) In order to estimate $`\tau _{int}`$, we perform very long MC simulation on $`L^3`$ simple cubic lattices, with $`L=16,20,24,28,32,36,40,44,48`$, at the previously calculated critical temperature $`k_BT_c/J=1.6778`$. The simulation started from a diagram containing no transposition and we observed that typically $`150L^3`$ configurations were needed to bring the system to equilibrium. So we discarded the appropriate number of configurations for equilibration, after which we recorded the susceptibility and energy every $`\delta t=2000`$ MCS, generating long equilibrium time series of $`10^6`$ measurements each. The time-displaced correlation functions were obtained by $`C_q(t)=\mathrm{\Delta }_q(t)/\mathrm{\Delta }_q(0)`$, where $`\mathrm{\Delta }_q(t)`$ is the autocovariance function given by $$\mathrm{\Delta }_q(t)=\frac{1}{nt}\underset{i=1}{\overset{nt}{}}(q_iq)(q_{i+t}q),$$ (11) $`n`$ is the length of the time series, and $`q`$ represents the physical property one is interested in. Here, we computed $`C_\chi (t)`$ and $`C_E(t)`$, the correlation function of the susceptibility and energy respectively. Typical equilibrium traces of the susceptibility and energy are shown in figure 5, where the microscopic time scale used equals to $`2000`$ MCS. From these we can infer that the number of MCS needed to generate two diagram configurations with uncorrelated susceptibilities is much smaller than the one required to generate uncorrelated energies. In practice, $`\tau _{int}`$ was estimated by $$\tau _{int}=\underset{t=0}{}C_q(t)$$ (12) and the sum was cut off at the first negative value of $`C(t)`$. Despite of our long runs we were not able to get reliable estimates of $`\tau _{int}`$ by integrating $`C(t)`$ for the largest lattices simulated. It is well known that $`C(t)`$ fluctuates wildly for large $`t`$, hampering the convergence of its integral. On the other hand, in the context of MC simulations the error associated to a given quantity can be written as $$\sigma ^2=\sigma _0^2\left(1+\frac{2\tau }{\delta t}\right)$$ (13) where $`\sigma _0`$ is the standard deviation treating all data as they were statistically independent and $`\sigma `$ is the actual statistical uncertainty. This variance inflation correctly takes into account the correlations of the MC data. It is not a simple matter to access the actual error in a finite time series of correlated data. Here we employed the moving block bootstrap (MBB) method which exploits resampling techniques. Within the MBB scheme a block of observations is defined by its length and by its starting point in the series. For instance, $`Q_i=\{q_i,q_{i+1},\mathrm{},q_{i+l}\}`$ defines the i$`th`$ block of $`l`$ observations. A MBB sample is then obtained by: (i) randomly drawing with replacement from the set of all possible overlapping blocks of size $`l`$; (ii) concatenating the selected blocks forming a replicated series. Each set of replicated data obtained in this way yields one estimate for the sample mean. The drawing is repeated many times and the block size dependent error is approximated by the standard deviation of the bootstrap generated mean values. It can be shown that in the case of arithmetic mean, $`\sigma ^2`$ can be calculated exactly without resampling. For a series with $`n`$ observations, $`q_t`$, and $`k`$ blocks of size $`l`$, $`\sigma ^2`$ is given by $$\sigma ^2=\frac{1}{kn}\underset{j=0}{\overset{n1}{}}\left[\frac{1}{l}\underset{t=1}{\overset{l}{}}(q_{j+t}q)\right]^2.$$ (14) The behavior of the ratio $`\sigma ^2/\sigma _0^2`$ is illustrated in figure 7 for the susceptibility. The error increases with the block size until it becomes roughly size independent for block length large enough. The maximum value reached by error corresponds to the actual standard error of the mean. The underlying idea of the MBB method is that if the block length is large enough, observations belonging to different blocks are nearly independent, while the correlation present in observations forming each block is retained. Having an estimate to $`\sigma ^2/\sigma _0^2`$, Eq. 13 can be employed to extract $`\tau _{int}`$. The above outlined procedure was applied for the data of the susceptibility and energy of all lattices. Good agreement was achieved between the estimates of $`\tau _{int}`$ obtained from MBB and by applying directly Eq. 12 for small lattices. The computed equilibrium relaxation times from both, susceptibility and energy, are plotted in figure 8 as obtained from lattices of size $`L=16,24,28,\mathrm{},48`$. Notice that, although $`\tau _{int}`$ is quite smaller for the susceptibility, both exhibit the same power-law size dependence. A linear best fit for the energy and susceptibility data yields $`3.0\pm 0.1`$ for the regression coefficient. Therefore, the equilibrium relaxation time scales as $`\tau _{int}L^{3.0\pm 0.1}`$. This means that, within the Handscomb dynamics, the number of Monte Carlo steps per site required to generate uncorrelated diagram configurations at criticality is roughly size independent. ## V conclusions In summary, we performed Monte Carlo simulations of the $`S=1/2`$ Heisenberg ferromagnet on the simple cubic lattice to investigate the critical relaxation of the Handscomb quantum Monte Carlo method which samples the space of permutation operators appearing in the series expansion of the partition function. Precise estimates of the critical temperature and exponents $`\gamma /\nu `$ and $`\beta /\nu `$ were obtained from a phenomenological renormalization group analysis of data from the susceptibility and order parameter. At the critical temperature we measured the equilibrium relaxation time from the time-displaced correlation functions of the susceptibility and energy (small lattices only). For the largest lattices ($`L32`$) $`\tau _{int}`$ was estimated through the moving block bootstrap technique. From either susceptibility or energy we obtained that, at criticality, the number of Monte Carlo steps (sampled permutation sequences) required to generate uncorrelated equilibrium diagram configurations scales with the system’s volume. Some care must be taken when estimating the efficiency of the Handscomb method and comparing it with other Monte Carlo prescriptions. Firstly, the phase space sampled within the Handscomb method is not related to any physical space. Therefore, there is no direct relation between the time scales of the Handscomb and the traditional spin-flip dynamics. However, a crude estimate can be drawn by considering that during an elementary Monte Carlo step of the Handscomb dynamics the sites belonging to a particular cycle of permutations have their status updated. The average number of sites involved in a single Monte Carlo step is then proportional to $`\frac{1}{N}a_i^2\chi L^{\gamma /\nu }`$. Within this reasoning, the average fraction of sites updated in a MCS scales as $`L^{\gamma /\nu }/L^d`$. Therefore, a time scale which would correspond to a lattice sweep in spin-flip dynamics would be $`\tau _0L^{d\gamma /\nu }`$. In units of this time scale the relaxation time scales as $`\tau _{int}\tau _0L^z`$, with $`z=2\pm 0.1`$, which is quite similar to the value of $`z`$ found for the decoupled cell Quantum Monte Carlo and the classical Metropolis dynamics. Although the Handscomb dynamics depicts some characteristics of the classical spin-flip cluster dynamics it has not a similar effect on dealing with the critical slowing down. It is relevant to mention here that the present Handscomb prescription, which inserts or removes transposition operators at the extremes of the permutation sequence, is the one that provide the simplest algorithm to control the dynamics in the permutation phase space. A natural generalization is to insert and remove operators at random locations within the sequence. This would drastically change its cycle structure with all sites being able to have their status updated on a single step. It would be valuable to estimate the efficiency of such relaxational dynamics at criticality as well as that of other generalizations of the Handscomb prescription as applied to antiferromagnet and large spin models. ## VI acknowledgments We are indebted to D.P. Landau for his suggestions and critical reading of the manuscript. This work was partially supported by CNPq and CAPES (Brazilian research agencies). MLL would like to thank the hospitality of the Physics Department at Universidade Federal de Pernambuco during the Summer School 2000 where this work was partially developed. ## FIGURE CAPTIONS Fig.1 - The susceptibility per spin as a function of temperature for $`L=16,24`$ and $`32`$ (from below). Due to an intrinsic symmetry of the Handscomb dynamics, the susceptibility equals the magnetization second moment below $`T_c`$. The errors are much smaller than the size of the symbols. Fig.2 - The auxiliary functions $`g_\chi (T,L,L^{})`$ for the scaling of susceptibility data. The renormalizations were performed from $`L=24`$ to $`L^{}=16`$ (circles); $`L=32`$ to $`L^{}=16`$ (squares); $`L=40`$ to $`L^{}=16`$ (diamonds) and from $`L=40`$ to $`L^{}=24`$ (triangles). Typical error bars are shown. The solid lines are the results from renormalizations of the best fits of our original susceptibility data. These have a common point from which we estimate $`T_c=1.677\pm 0.001`$ and $`\gamma /\nu =1.98\pm 0.01`$. Fig.3 - The average size of the largest cycle (normalized the the total number of sites) $`\psi `$ as a function of temperature for $`L=16,24`$ and $`48`$. At high temperatures all cycles are small indicating no long range order and $`\mathrm{\Psi }`$ vanishes. With lowering $`T`$, the onset of the ferromagnetic order makes itself felt around $`k_BT_c/J1.68`$, and $`\psi `$ start to grow until saturation. At criticality, $`\psi `$ shows power law size dependence. Fig.4 - The auxiliary functions $`g_\psi (T,L,L^{})`$ for the scaling of the order parameter data. The renormalizations were performed from $`L=24`$ to $`L^{}=16`$ (circles); $`L=40`$ to $`L^{}=16`$ (squares); $`L=48`$ to $`L^{}=16`$ (triangles up) and from $`L=48`$ to $`L^{}=24`$ (triangles down). Typical error bars are shown. The solid lines are the results from renormalizations of the best fit of our original order parameter data. From the interception of these functions computed for all possible renormalizations with lattice sizes $`L=16,24,32,40`$ and $`48`$ we estimate $`k_BT_C/J=1.6778\pm 0.0002`$ and $`\beta /\nu =0.512\pm 0.002`$. Fig.5 - An equilibrium trace of the susceptibility $`\chi `$ and energy $`E`$ at criticality. Local quantities, as energy, are more time correlated than non-local ones due to the cluster nature of the Handscomb Monte Carlo. Fig.6 - The time-displaced equilibrium correlation function of the susceptibility and energy at criticality for several lattice sizes. Fig.7 - Moving block bootstrap estimates of the standard errors of the susceptibility as a function of the block length $`l`$ at criticality for several lattice sizes. The lines are guides to the eye. Fig.8 - The equilibrium relaxation time versus linear size $`L`$ for susceptibility and energy. The error in our estimates of $`\tau _{int}`$ is around 2%. Though, $`\tau _{int}`$ is much smaller for the susceptibility, both quantities scale the same way. The microscopic time scale used is $`2000`$ MCS.
warning/0004/cond-mat0004117.html
ar5iv
text
# Coherent Atomic Oscillations and Resonances between Coupled Bose-Einstein Condensates with Time-Dependent Trapping Potential ## I Introduction It has recently been shown that there exists a macroscopic quantum phase difference in processes connected with atomic waves. Namely, this effect is observed between two tunnel-coupled Bose-Einstein condensates -. The coupling occurs due to a trap potential which has the form of a double-well potential. The theoretical investigation shows that here is possibly an interesting phenomenon of periodic oscillations of the atomic population between condensates and quantum self-trapping of population in the dependence of the relative phase between condensates -. Analogous phenomena which have been studied are the ac Josephson effect and the periodic exchange of power and switching of electromagnetic waves between cores in nonlinear optical couplers . In particular, there is a direct analogy between the tunneling phenomena in two prolongated Bose-Einstein condensates and two tunnel-coupled single mode optical fibers. In this optical analogy, the role of the chemical potential is played the propagation constant, and to the nonlinear interaction between atoms is analogous to the Kerr nonlinearity of optical media. The tunnnel-coupling arising from the overlaps of the electromagnetic fields outside of dielectric cilinders (fibers) exactly corresponds to the tunnel coupling between two Bose-Einstein condensates due to the overlaps of the wavefunctions. In this context , it is natural to investigate the influence of a time varying coupling on the quantum coherent tunneling process. In the nonlinear optical coupler analogy this corresponds to the variation in the longitudinal direction coupling . As was shown recently, the variation in time of the trap potential can lead to resonant oscillations of the Bose-Einstein condensate . The numerical solution of the Gross-Pitaevsky equation shows that the rapid variation in time of the trap potential can lead to a bifurcation of the effective (averaged ) form of the trap potential and consequently to splitting of the BEC . From what we have said above, we could wait for new phenomena in the process of quantum coherent atomic tunneling (QCAT) process too. Recalling the analogy with Josephson effects in coupled junctions, we note that in the last case it is very difficult to produce such kind of effects, the difficulty coming from the fact that it is hard to implement the time varying overlap properties of junctions. Thus, in this article we will study the atomic tunneling between two tunnel-coupled BEC’s in a double-well time dependent trap. Namely, in Section II we will study the nonlinear resonance phenomena in the interference and, in particular, the stationary regimes, with damping taked into account. The phase and population damping are considered as well. The phase locked states are interesting because the suppression in these states of the fluctuations of the relative phase between two condensates. The regions of regular and chaotic oscillations of the relative atomic population are calculated in Section III, using the Melnikov function approach. In Section IV the macroscopic quantum interference in two overlapped condensates with rapidly varying traps is investigated. We derive the system of averaged equations for coupled modes and define in Section V new regimes of parameters for the macroscopic quamtum self-trapping (MQST). ## II Formulation of problem. Slowly varying in time trap The problem of Bose-Einstein Condensates in a two double-well time dependent trap can be described by the following model $`ih{\displaystyle \frac{\psi _1}{t}}=(E_1(t)+\alpha _1|\psi _1|^2)\psi _1K(t)\psi _2`$ (1) $`ih{\displaystyle \frac{\psi _2}{t}}=(E_2(t)+\alpha _2\psi _2|^2)\psi _2K(t)\psi _1,`$ (2) where the parameters $`E_i,\alpha _i,K(t)`$ are defined by the overlaps integrals of the time dependent Gross-Pitaevsky eigenfunctions as given in . $`\alpha _i`$ are parameters of nonlinear interactions between atoms and $`\alpha _ig_0=4\pi h^2a/m`$, $`a`$ is the atomic scattering length. The attractive case can be obtained analogously, using symmetries of the equation This system of equations is valid in the approximation of a weak link between condensates. Comparison with the numerical solution of GPE shows that the system (2) is a reasonable approximation for $`z0.50.6.`$ For the strongly overlapped condensates the system should be modified. For a periodically varying $`K(t)=K_0+K_1\mathrm{sin}(\mathrm{\Omega }t)`$ we have limitations on the parameters $`K_1`$ and $`\mathrm{\Omega }`$ for which the two-mode approximation is valid. The small deformation of bound states requires $`K_1K_0`$. To avoid the resonances of the modulations with normal modes of the spherical well we should require that the difference of energy between the ground state and the first normal mode be larger than the modulation energy, i.e. $`\delta E\mathrm{}\mathrm{\Omega }`$. Taking into account that $`\delta E\mathrm{}\omega _0`$, , where $`\omega _0`$ is the harmonic oscillator frequency, we find the condition $`\mathrm{\Omega }\omega _0`$. Taking into account that $`\omega _0\omega _L=2K_0`$ – the frequency of the linear oscillations of the atomic population, defined by the tunneling frequency, we conclude that by a proper choice of parameters we can have the resonant case $`\mathrm{\Omega }\omega _L`$ or the rapidly varying modulation $`\mathrm{\Omega }\omega _L`$ . Introducing new variables $`u_i=\sqrt{N_i}\mathrm{exp}(i\theta _i),z=(N_1N_2)/N_T,N_T=N_1+N_2,\psi =\theta _1\theta _2`$, where $`N_i`$, and $`\theta _i`$ are respectively the number of atoms and phases in i-th trap, we get the following system $`z_t=2K(t)\sqrt{1z^2}sin\mathrm{\Phi }\eta \mathrm{\Phi }_t,`$ (3) $`\mathrm{\Phi }_t=\nu \mathrm{\Lambda }z+{\displaystyle \frac{2K(t)z}{\sqrt{1z^2}}}\mathrm{cos}\mathrm{\Phi }+\mathrm{\Delta }E(t),`$ (4) where $`\mathrm{\Lambda }=(\alpha _1+\alpha _2)N_T,`$ and $`\nu =\pm 1`$ for the positive and negative atomic scattering length respectively. In what follows, we will consider the case $`\nu =1`$, ($`a_s>0`$), if not stated otherwise. We have included in Eq(3) the damping term $`\eta \varphi _t(t)`$. It appears if we take into account a noncoherent dissipative current of normal-state atoms, proportional to the chemical potential difference $`\mathrm{\Delta }\mu `$. For the other type of overlapping condensate there might be another type of damping. For example, for the two interacting condensates with different hyperfine levels in a single harmonic trap the damping has the form $`\eta z(t)`$. The Hamiltonian of the unperturbed system ($`K(t)=`$ const,$`\mathrm{\Delta }E=`$ const, $`\eta =0`$) is $$H=\frac{\mathrm{\Lambda }z^2}{2}2K\sqrt{1z^2}\mathrm{cos}(\mathrm{\Phi })+\mathrm{\Delta }Ez.$$ (5) Let us consider the case of periodic modulations of the tunnel coupling parameter, when $`K(t)=K_0+K_1\mathrm{sin}\mathrm{\Omega }t`$ and the time-periodic energy difference is $`\mathrm{\Delta }E(t)=\mathrm{\Delta }E+\mathrm{\Delta }E_1\mathrm{sin}(\mathrm{\Omega }t)`$. It is useful to introduce the parameter $`\mathrm{\Delta }=\mathrm{\Phi }\mathrm{\Omega }t`$, which is a slow-varying function of the time in a comparison with the period $`\mathrm{\Omega }`$. The averaged system of equations can be found by averaging over the period of fast oscillations $`2\pi /\mathrm{\Omega }`$. Doing so, we get the system $`\overline{z}_t=K_1\mathrm{cos}\overline{\mathrm{\Delta }}\sqrt{1\overline{z}^2}\eta \mathrm{\Delta }_t\eta \mathrm{\Omega },`$ (6) $`\overline{\mathrm{\Delta }}_t=\mathrm{\Omega }+\mathrm{\Delta }E+\mathrm{\Lambda }z{\displaystyle \frac{K_1z}{\sqrt{1\overline{z}^2}}}\mathrm{sin}\overline{\mathrm{\Delta }}.`$ (7) Let us now look for the fixed points of Eq.(6). Introduce first the notation $`\alpha =\mathrm{\Omega }\mathrm{\Delta }E`$. From the system (6) we find the following equation for the fixed points $`z_c=\sqrt{1{\displaystyle \frac{\eta ^2\mathrm{\Omega }^2}{K_1^2\mathrm{cos}\mathrm{\Delta }_{c}^{}{}_{}{}^{2}}}},`$ (8) $`\alpha =z_c(\lambda K_1^2{\displaystyle \frac{\mathrm{sin}2\mathrm{\Delta }_c}{2\eta \mathrm{\Omega }}}).`$ (9) For small $`z_c,z_c^2<<1`$ we can find the explicit solution $`\mathrm{\Delta }_c=\mathrm{arccos}({\displaystyle \frac{\eta \mathrm{\Omega }}{K_1}}),`$ (10) $`z_c{\displaystyle \frac{\mathrm{\Omega }\mathrm{\Delta }E}{\mathrm{\Lambda }\sqrt{K_1^2\eta ^2\mathrm{\Omega }^2}}}.`$ (11) The existence of a stationary solution is connected with the circumstance that the damping of the oscillation in quantum tunneling is compensated by the periodic variation of the trap. In reference the influence of the damping on the macroscopic quantum interference has been investigated numerically and degradation of the interference due to damping has been analyzed. The predicted stationary solutions, where these effects are compensated by the variable driving, are indeed phase - locked solutions. Thus we can think that the fluctuations of the relative phase will be suppressed in such states. It is an important prediction since BEC’s have phase fluctuations and it is difficult to maintain the constant relative phase of condensates. This phenomenon can be useful to develop the phase standard for the Bose-Einstein condensate- a problem attracting considerable attention recently . We can estimate the value of fixed points for an experimental situation. For example, when $`\mathrm{\Omega }=0.7,\eta =0.1,K_1=0.2,\mathrm{\Lambda }=2.5`$, we obtain the fixed points $`z_{1c}=0.2599,z_{2c}=0.3037`$. Stability analysis shows that only the first point is stable and corresponds to the global attractor. The results of the numerical simulations of the initial system for the relative population $`z(t)`$ are presented in Fig.1. It shows that the agreement between theory and numerical simulations is good. The same system of equations also describes the nonlinear Josephson type oscillations in the relative population of a driven , two-component Bose-Einstein condensate. The two component condensates where two different hyperfine states can be populated and confined in the same trap. A weak driving field couples two internal states and as result the oscillations in the relative population will occur . In this case the damping term has the form $`\mathrm{\Gamma }=\eta z`$. Repeating the above mentioned procedure we obtain the averaged system $`\overline{z}_t=K_1\sqrt{1\overline{z}^2}\mathrm{cos}\overline{\mathrm{\Delta }}\eta \overline{z},`$ (12) $`\overline{\mathrm{\Delta }}_t=\mathrm{\Lambda }\overline{z}+\mathrm{\Delta }E\mathrm{\Omega }{\displaystyle \frac{K_1\overline{z}}{\sqrt{1\overline{z}^2}}}\mathrm{sin}\overline{\mathrm{\Delta }}.`$ (13) In the model , $`\mathrm{\Delta }E`$ corrsponds to $`\mu _2\mu _1\kappa `$, where $`\mu _i`$ are chemical potentials and $`\kappa `$ is frequency detuning $`\kappa =\omega _d\omega _0`$, $`\omega _0`$ is the separation in frequency of two hyperfine states, $`\omega _0`$ is the frequency of drive. The fixed points are: $`\overline{\mathrm{\Delta }}_c=\mathrm{arccos}\left({\displaystyle \frac{\eta z_c}{K_1\sqrt{1z_c^2}}}\right),`$ (14) $`\mathrm{\Lambda }z_c\alpha {\displaystyle \frac{K_1z_c}{\sqrt{1z_c^2}}}(1{\displaystyle \frac{\eta ^2z_c^2}{K_1^2(1z_c^2)}})^{1/2}.`$ (15) Here $`\alpha =\mathrm{\Omega }\mathrm{\Delta }E`$. For the case of small $`z_c^2<<1`$ we have the estimate for $`z_c`$ $$z_c=\frac{\mathrm{\Delta }E\mathrm{\Omega }}{\mathrm{\Lambda }K_1}+\frac{(\mathrm{\Delta }E\mathrm{\Omega })^3(K_1^2\eta ^2)}{2K_1(\mathrm{\Lambda }K_1)^4}.$$ (16) The typical values of parameters are:$`\mathrm{\Omega }=0.7,\mu _2\mu _1\kappa =\mathrm{\Delta }E=1,\eta =0.1,K_1=0.2,\mathrm{\Lambda }=4`$, and the fixed points are $`z_{3c}=0.111,z_{4c}=0.13`$. The analysis shows that the first fixed point is stable and second is unstable. In Fig.2. we present the plot of solution of the full system for the damping term $`\eta z`$ with the above mentioned set of parameters. All solutions decay to the first fixed point. ## III Chaotic oscillations in relative atomic population In this section we consider the interference in the system which in some limit is mathematically equivalent to the motion of a particle in a double-well potential under parametric periodic perturbations. The particle’ motion is of course unharmonic and the nonlinear resonances between oscillations of a particle and oscillations of the trap potential are possible. One of consequences of this is the possible appearence of chaotic dynamics in such systems. We will now study the conditions for the appearence of the chaotic dynamics in atomic tunneling. To proceed it will be useful first to consider some aspects of the unperturbed system, that is, when the coupling is constant. In this case the system (3) is equivalent to the quartic (Duffing) oscillator . We can write the equation for $`z(t)`$ as $$z_{tt}=\frac{V}{z},$$ (17) where the potential $`V(z)`$ is given by $$V(z)=z^2(a+bz^2),a=2K_0^2\frac{\mathrm{\Lambda }H}{2},b=\frac{\mathrm{\Lambda }^2}{8}.$$ (18) and $`H`$ is the Hamiltonian. The total energy of the effective particle is $$E_0=\frac{z_t^2}{2}+V(z)=2K^2H^2/2.$$ When $`E_0>0`$ we have the oscillating regime with $`<z(t)>=0`$ (nonlinear Rabi oscillations). This case corresponds to the periodic flux of atoms from one BEC to other. When $`E_0<0`$ the motion of effective particle is confined in the one of the wells. This case corresponds to the localization of atomic population in one of the condensates - the so called macroscopic quantum self-trapping (MQST). The case $`E_0=0`$, i.e. $`H=2K`$, corresponds to the separatrix solution separating these two regimes. It is interesting to investigate the dynamics of atomic population, when the inital parameters $`z(0),\mathrm{\Phi }(0)`$ are close to the separatrix of the unperturbed systsem. From the general theory of nonlinear driven oscillations we can expect the appearence of chaotic macroscopic tunneling phenomena. The separatrix solutions for the right hand side well(condensate) are: $`z_s(t)=\sqrt{{\displaystyle \frac{a}{b}}}\text{sech}(\sqrt{2a}t),`$ (19) $`\mathrm{sin}^2(\mathrm{\Phi }_s)={\displaystyle \frac{a^2\text{sech}^2(\sqrt{2a}t)\mathrm{tanh}^2(\sqrt{2a}t)}{2K_0^2(ba\text{sech}^2(\sqrt{2a}t))}},`$ (20) where $`H=H_s=2K`$. The motion near the separatrix is very sensitive to the change of the initial condition and a stochastic layer appeas, leading to the chaotic dynamics of the solutions of the system (3). The Melnikov function method is effective to find the regions of chaotic behavior . The system (3) may written in the form $`𝐫_t=𝐟+ϵ𝐠`$ and has a hyperbolic fixed point. According to this method, we need to calculate the Melnikov function $`M(t_0)`$ $$M(t_0)=_{\mathrm{}}^{\mathrm{}}(f_{2s}g_{1s}(t,t_0)f_{1s}g_{2s}(t,t_0))𝑑t.$$ (21) As is well known, the existence of zeros of $`M(t_0))`$ indicates the intersection of separatrices and the existence of homoclinical chaos. Substituting the expressions for $`f_i,g_i`$ from the system (3) and calculating the integrals, we find the final expressions for the Melnikov function. We shall give the results for two possible damping terms. ### A The case of the damping in the form $`\eta \mathrm{\Phi }_t`$ The integrations in (21) are cumbersome, and will be omitted. The final result for the Melnikov function is: $$M_1(t_0)=K_1F_1\mathrm{cos}(\mathrm{\Omega }t_0)F_2\eta ,$$ (22) where $`F_1={\displaystyle \frac{\pi \mathrm{\Lambda }\mathrm{\Omega }^2}{4K_0b\mathrm{sinh}(\frac{\pi \mathrm{\Omega }}{2\sqrt{2a}})}},`$ (23) $`F_2={\displaystyle \frac{2\mathrm{\Lambda }\sqrt{2a}}{\sqrt{aba^2}}}(\mathrm{\Lambda }2H_s)\text{arctan}(\sqrt{{\displaystyle \frac{a}{ba}}}){\displaystyle \frac{\sqrt{2a}}{16(ba)b^2}}G_1,`$ (24) where $`G_1=[2b^2(2H_s\mathrm{\Lambda })(bH_s4a\mathrm{\Lambda }+3b\mathrm{\Lambda })]{\displaystyle \frac{\text{arctan}(a/\sqrt{aba^2})}{\sqrt{aba^2}}}`$ (25) $`2\left[4a^2\mathrm{\Lambda }^25ab\mathrm{\Lambda }^2+2b^2(4H_s^24H_s\mathrm{\Lambda }+3\mathrm{\Lambda }^2)\right].`$ (26) The simple zero of $`M(t_0)`$ is absent if the condition $$\eta >\frac{F_1}{F_2},$$ (27) is satisfied. In Fig.3 we plot this criterion in the $`(K_1,\mathrm{\Omega })`$ plane for $`\eta =0.1`$ . Regions below the curve correspond to regular oscillations and above to chaotic oscillations of $`z(t)`$. The criterion gives a good lower bound on the region of chaos in this plane. We verified this lower bound in ($`K_1,\mathrm{\Omega }`$) plane by numerical simulations and found a good agreement with the formulae (27). When the damping coefficient is zero we have the expression for the width of stochastic layer near the separatrix as $$D=\frac{\pi \mathrm{\Lambda }K_1\mathrm{\Omega }^2}{4K_0b\mathrm{sinh}(\frac{\pi \mathrm{\Omega }}{2\sqrt{2a}})}.$$ (28) In Fig.4 we plot the width of the stochastic layer as a function of the frequency for $`K_1=0.2,\mathrm{\Lambda }=9.9,z(0)=0.6,\mathrm{\Phi }(0)=0`$. The maximum of the width is given by the solution of equation $$\mathrm{\Omega }\mathrm{tanh}(\frac{\pi \mathrm{\Omega }}{2\sqrt{2a}})=\frac{4\sqrt{2a}}{\pi }.$$ For this choice of parameters we have $`\mathrm{\Omega }_m=1.89`$. For $`\mathrm{\Omega }2,a1`$ we get the estimate $`\mathrm{\Omega }_m4\sqrt{2a}/\pi `$. It can be seen that, for the high frequencies $`\mathrm{\Omega }>>\sqrt{a}`$, the stochastic layer is exponentially narrow and motion is regular. The analysis of this case will be performed in detail in the next section. In Fig.5 we plot the typical oscillations in the relative atomic population for $`\mathrm{\Lambda }=9.9,K_0=1,K_1=0.2,z(0)=0.6,\mathrm{\Phi }(0)=0`$ for $`\mathrm{\Omega }=1.5`$. It turns out that for $`\mathrm{\Omega }`$ between $`1.5`$ and $`3.22`$ we have chaotic oscillations, and for small or large $`\mathrm{\Omega }`$, regular oscillations. In Fig.6 the phase portrait is given. In Fig.7 the influence of the damping on chaos is illustrated. Parameters are the same as in Fig.5 and $`\eta =0.15`$. In this case the oscillations become regular. ### B The case of the damping term in the form $`\eta z(t)`$. The calculation of the integrals which give the Melnikov function is again cumbersome, and one find the following result: $$M(t_0)=F_1\mathrm{cos}(\mathrm{\Omega }t_0)+\eta F_3,$$ (29) where $$F_3=\frac{\mathrm{\Lambda }2H_s}{\sqrt{2(ba)}}\text{arctan}\left(\sqrt{\frac{a}{ba}}\right)+\frac{3\mathrm{\Lambda }\sqrt{a}}{2\sqrt{2}b}.$$ (30) The zeros of $`M(t_0)`$ are absent when the condition $$\eta >\frac{F_1}{F_3},$$ (31) is satisfied. If we consider for example the case when $`K_1=0.2,\mathrm{\Omega }=0.7,\mathrm{\Lambda }=2.5,z(0)=0.6,\mathrm{\Phi }(0)=\pi `$, the damping constant should be $`\eta =0.15`$. In Figs.(7,8) the chaotic and periodic oscillations of the relative population in the dependence of the frequency for fixed $`z(0),\mathrm{\Phi }(0)`$ are plotted. ## IV Rapidly varying trap potential In the rapidly varying case $`K=K(\frac{t}{ϵ})`$ we can apply he averaging techniques based on the multiscale expansions . Let us search the solution in the form of the sum of the slowly varying components $`U,V`$ and rapidly varying corrections $`u_i,v_i`$ $`\psi _1(t)=U(t_k)+ϵu_1(\zeta )+ϵ^2u_2(\zeta )+\mathrm{}`$ (32) $`\psi _2(t)=V(t_k)+ϵv_2(\zeta )+ϵ^2v_2(\zeta )+\mathrm{}`$ (33) where $`\zeta =t/ϵ,ϵ<<1,`$, $`t_k=ϵ^kt`$ are slow times. The averages over a period of rapid oscillations are $`<u_i>=<v_i>=0`$. The functions $`u_i,v_i`$ are assumed to be functions of the rapid time $`\zeta `$ and the slow varying functions $`U,V`$. The derivatives can be written in the form $$\frac{dU}{dt}=\frac{dU}{dt_0}+ϵ\frac{dU}{dt_1}+\mathrm{}$$ (34) and analogously for the derivative of $`V`$. Below, we will assume $`E_1=E_2,`$. Then the time dependence in $`E(t)`$ can be removed by a simple field transformation $`u(v)=u(v)\mathrm{exp}((i/h)E(t^{})𝑑t^{})`$. In the zeroth order in $`ϵ`$ we get the equations $`ih({\displaystyle \frac{u_0}{t_0}}+{\displaystyle \frac{u_1}{\zeta }})=Eu_0K(\zeta )v_0+\alpha |u_0|^2u_0`$ (35) $`ih({\displaystyle \frac{v_0}{t_0}}+{\displaystyle \frac{v_1}{\zeta }})=Ev_0K(\zeta )u_0+\alpha |v_0|^2v_0`$ (36) Requiring the exclusion of the secular in $`\zeta `$ terms, we find the corrections $`u_1,v_1`$ $`u_1={\displaystyle \frac{i}{h}}v_0(\mu _1<\mu _1>)`$ (37) $`v_1={\displaystyle \frac{i}{h}}u_0(\mu _1<\mu _1>)`$ (38) where $`\mu _1=_0^\zeta (K(s)<K>)𝑑s`$ and the system for $`u_0,v_0`$ becomes: $`ih{\displaystyle \frac{u_0}{v_0}}=Eu_0<K>v_0+\alpha |u_0|^2u_0`$ (39) $`ih{\displaystyle \frac{v_0}{t_0}}=Ev_0<K>u_0+\alpha |v_0|^2v_0`$ (40) In the order $`ϵ`$ we have $$\frac{u_0}{t_1}=0,\frac{v_0}{t_1}=0,$$ (41) and we obtain, for the second order corrections, $`u_2,v_2`$, the elimination of the secular terms, results in: $`u_2={\displaystyle \frac{i}{h}}(\mu _2<\mu _2>)[{\displaystyle \frac{v_0}{t_0}}+{\displaystyle \frac{i}{h}}v_0+{\displaystyle \frac{i\alpha }{h}}[2|u_0|^2v_0+u_0^2v_0^{}]]`$ (42) $`[(\mu _1<\mu _1>)^22M]u_0,`$ (43) $`v_2={\displaystyle \frac{i}{h}}(\mu _2<\mu _2>)[{\displaystyle \frac{u_0}{t_0}}+{\displaystyle \frac{i}{h}}u_0+{\displaystyle \frac{i\alpha }{h}}[2|v_0|^2u_0+v_0^2u_0^{}]]`$ (44) $`[(\mu _1<\mu _1>)^22M]v_0,`$ (45) where $$\mu _2=_0^\zeta (\mu _1\zeta <\mu _1>)𝑑s,M=(<\mu _1^2><\mu _1>^2).$$ ¿From this order we get the equations: $`{\displaystyle \frac{u_0}{t_2}}={\displaystyle \frac{2iM\alpha }{h^3}}[2|v_0|^2u_0v_0^2u_0^{}],`$ (46) $`{\displaystyle \frac{v_0}{t_2}}={\displaystyle \frac{2i\alpha M}{h^3}}[2|u_0|^2v_0u_0^2v_0^{}].`$ (47) Finally, we have in the next order, $`O(ϵ^2)`$, the averaged system $`i\mathrm{}{\displaystyle \frac{u_0}{t}}Eu_0\alpha |u_0|^2u_0=<K>v_0+2{\displaystyle \frac{ϵ^2\alpha M}{\mathrm{}^2}}[2|v_0|^2u_0v_0^2u_0^{}]`$ (48) $`i\mathrm{}{\displaystyle \frac{u_0}{t}}Ev_0\alpha |v_0|^2v_0=<K>v_0+2{\displaystyle \frac{ϵ^2\alpha M}{\mathrm{}^2}}[2|u_0|^2v_0u_0^2v_0^{}].`$ (49) For periodic modulations of tunnel coupling we obtain $`M=K_1^2/\mathrm{\Omega }^2`$. This averaged system is the main result of this section. We see that, in comparison with the constant coupling case, new effects appear. First, the cross term appears, corresponding to a change of type in the effective nonlinearity in coupled BEC’ dynamics. Second, there appears an additional nonlinear phase-sensitive coupling. This term leads to new minima in the effective potential picture. In the constant coupling case, the oscillations of $`z(t)`$ is described by the Duffing oscillator. Now we have a more complicated nonlinear oscillator, describing more complicated interference patterns. ## V Analysis of the averaged dynamics Using the variables introduced in the section II, we get the system $`z_t=2<K>\sqrt{1z^2}\mathrm{sin}(\mathrm{\Phi })\nu \delta (1z^2)\mathrm{sin}(2\mathrm{\Phi })`$ (50) $`\mathrm{\Phi }_t=\mathrm{\Lambda }z+2{\displaystyle \frac{<K>z}{\sqrt{1z^2}}}\mathrm{cos}(\mathrm{\Phi })\nu \delta z\left[2\mathrm{cos}(2\mathrm{\Phi })\right].`$ (51) where we have defined $`\delta =ϵ^2\alpha M/h^2`$. This system has the following Hamiltonian $$H=\frac{\mathrm{\Lambda }_1z^2}{2}2<K>\sqrt{1z^2}\mathrm{cos}\mathrm{\Phi }\frac{\delta }{2}(1z^2)\mathrm{cos}2\mathrm{\Phi },$$ (52) where $`\mathrm{\Lambda }_1=\mathrm{\Lambda }2\delta `$. When the parameter $`\delta =0`$, we get back the system (3). The effective particle motion is that of the double nonrigid pendulum system. Let us consider the different regimes of oscillations. ### A Linear regime This is the case where $`|z|<<1,|\mathrm{\Phi }|<<1`$. Then, from the system (50), it follows that that $$z_{tt}=2(<K>+\delta )(\mathrm{\Lambda }+2<K>\delta )z.$$ Then the frequency of linear oscillations of the number of atoms is $$\omega _L^2=2(<K>+\delta )(\mathrm{\Lambda }+2<K>\delta ).$$ Thus, the frequency of the linear oscillations grows with $`\delta \omega ^2=\omega _L^2\omega _0^2=2\delta (<K>+\mathrm{\Lambda }\delta )`$, in comparison with a nonmodulated case. ### B $`z^2<<1`$, and $`\mathrm{\Phi }`$ is not small. The phase dynamics is described by the equation: $`\mathrm{\Phi }_{tt}+2<K>(\mathrm{\Lambda }2\delta )\mathrm{sin}\mathrm{\Phi }+(\delta \mathrm{\Lambda }+2<K>^22\delta ^2)\mathrm{sin}(2\mathrm{\Phi })+`$ (53) $`2\delta <K>\mathrm{sin}(3\mathrm{\Phi })+{\displaystyle \frac{\delta ^2}{2}}\mathrm{sin}(4\mathrm{\Phi })=0.`$ (54) At derivation of this equation we take into account that $`\mathrm{\Phi }_tz`$, so we neglect by terms $`z\mathrm{\Phi }_t`$. This equation is equivalent to a generalized pendulum motion whose effective potential is given $$U(\mathrm{\Phi })=2<K>(\mathrm{\Lambda }2\delta )\frac{1}{2}(\mathrm{\Lambda }\delta +2<K>^22\delta ^2)\mathrm{cos}(2\mathrm{\Phi })\frac{2\delta <K>}{3}\mathrm{cos}(3\mathrm{\Phi })\frac{\delta ^2}{8}\mathrm{cos}(4\mathrm{\Phi }).$$ (55) The additional minima at $`\mathrm{\Phi }=\pm n\pi `$ correspond to the trapping of oscillations in states witt locked phase. ## VI Macroscopic quantum self-trapping. This phenomenon consists in the self trapping of the atomic number in one of condensates, depending on the initial difference in atomic number and on the relative phase of condensates . This is connected with the phase sensitive linear coupling. The nonlinear guided wave optics analog of this phenomenon, first predicted in , is the switching of the power in the dual core couplers from one core to other, depending on the initial power or the phase difference. The additional coupling in the averaged system (50) is nonlinear and also phase sensitive. Thus the condition of localization of atomic population should modified in comparison with the one considered earlier. Let us find the fixed points of the averaged system (50). The first group is for $`\mathrm{\Phi }_n=\pi n,n=0,\pm 1,\pm 2,\mathrm{}`$ For each $`\mathrm{\Phi }_n`$ there can exist one or three roots for $`z`$. One of the roots is $`z_1=0`$, two others are $$z_{s1,2}=\pm \left(1\frac{4<K>^2}{(\mathrm{\Lambda }\delta )^2}\right)^{1/2}.$$ (56) Note that if $`(\mathrm{\Lambda }\delta )<K><0`$, then 3 roots exists for even $`n`$ and one root ($`z_1=0`$) exist for odd $`n`$. For $`(\mathrm{\Lambda }\delta )<K>>0`$,one has the opposite situation, the critical value is $`\mathrm{\Lambda }_c>2<K>+\delta `$. One can show, however, linear stability analysis predicts that this set is unstable when $`\delta >\mathrm{\Lambda }/2`$. The second group of fixed points is given by: $$\mathrm{cos}(\mathrm{\Phi })=\frac{<K>}{\sqrt{1z^2}}.$$ (57) Again, there are two possibilities: i)$`\mathrm{\Lambda }=3\delta `$, then all line is stationary; ii)$`\mathrm{\Lambda }3\delta `$, then one has the following static point $$\mathrm{cos}(\mathrm{\Phi })=\frac{<K>}{\delta },z=0.$$ (58) This root assumes that $`|<K>/\delta |<1`$. The linear stability analysis shows that the fixed points are unstable when $`\mathrm{\Lambda }>3\delta `$. Following , from the constraint coming from the energy conservation and the boundness of the tunneling energy, given by $$H=\frac{\mathrm{\Lambda }_1z^2(0)}{2}2<<K>>\sqrt{1z^2(0)}\mathrm{cos}(\mathrm{\Phi }(0))\frac{\delta }{2}(1z^2(0))\mathrm{cos}(2\mathrm{\Phi }(0))2<K>\frac{\delta }{2},$$ (59) we obtain the estimate for the critical value of $`\mathrm{\Lambda }_c`$ when the self-localization occurs. We find that: $$\mathrm{\Lambda }_c=\frac{4<K>(\sqrt{1z^2(0)}\mathrm{cos}\mathrm{\Phi }(0)+1)+\delta ((1z^2(0))\mathrm{cos}2\mathrm{\Phi }(0)1)}{z^2(0)}+2\delta .$$ (60) The phase portrait is plotted in Fig.(9,10,11) for initial phase difference $`\mathrm{\Phi }(0)=\pi `$ and $`\mathrm{\Lambda }=2.5`$, for differnt $`\delta =0.2;0.6;0.8`$ The regions corresponding to different dynamics of $`z(t)`$ with $`<z>=0`$ and $`<z>0`$ are shown. The increasing of $`\delta `$ (i.e. $`K_1`$/$`\mathrm{\Omega }`$) leads to the distortion of the MQST regime, and for $`\delta 1`$ to the nonlinear Rabi-like oscillations. The typical oscillations in time of the relative population for different values of $`\delta `$ and for fixed $`\mathrm{\Lambda }`$ are plotted in Fig.9. The influence of rapid modulations leads to the appearance of new minima in the oscillations. This can be described as a result of the overlap of two double well potentials. This conclusion is also confirmed by the results of numerical simulations of the GP equation for the single BEC under rapidly varying trap potential performed in . In this work, it has been shown that the effect of rapid modulations is the appearance of a effective double well trap potential. ## VII Conclusion In this paper we have studied the new effects coming the time variation of the trap potential, and from damping, on the nonlinear oscillations in the relative population behavior between two BEC’s. For the slowly varying trap we predict synchronization of oscillations of the trap with oscillations of the relative population. We find the fixed points for both types of the damping terms occuring in the studies of two coupled condensates. Using the Melnikov approach we study the possibility the appearance of the chaotic oscillations in the tunneling phenomena between two coupled Bose-Einstein condensates. We calculate the width of the stochastic layer in phase space, in which the dynamics is chaotic. We find the lower bound on the region of chaos in $`(K_1,\mathrm{\Omega })`$ plane. We obtain the estimate for the damping coefficient when the chaos is suppressed. For the rapidly varying trap we use the multiscale method and derive the averaged equations for coupled modes describing the tunneling phenomena. We find the fixed points in this case describing the stationary states in two coupled BEC’s. The expression for the critical value of the nonlinearity parameter , when the macroscopic quantum self-trapping occurs is derived. Our results show that there is interest to study the contribution of the quantum tunneling for this driven case and the investigation of resonances and chaos in oscillations of atomic poulation for the strongly overlapped condensates . These problems require separate investigation. ## VIII Acknowledgments Authors are grateful to E.N.Tsoy for useful discussions and comments. This work was partially supported by FAPESP Grant.
warning/0004/astro-ph0004090.html
ar5iv
text
# The “Swiss cheese” cosmological model has no extrinsic curvature discontinuity: A comment on the paper by G.A. Baker, Jr. (astro-ph/0003152) ## 1 Introduction The “Swiss cheese” cosmological model is a general relativistic description of space-time. The name refers to the fact that in this model static spherical voids are created within a larger, time-dependent space-time. A void is constructed by removing the background material inside a spherical boundary and replacing the mass by a concentration of that mass at the centre of the sphere. Mathematically, the model is realized by the matching of a Friedmann-Lemaître-Robertson-Walker (FLRW) metric as the exterior solution, to an exterior Schwarzschild metric as the interior solution, across a spherical boundary. The spherical boundary stays at a fixed coordinate radius in the FLRW frame, but changes with time in the Schwarzschild frame. The smooth matching of two space-times across a three-surface of discontinuity $`\mathrm{\Sigma }`$ is guaranteed if the Darmois junction conditions are satisfied: the first fundamental forms (intrinsic metrics) and the second fundamental forms (extrinsic curvatures) calculated in terms of the coordinates on $`\mathrm{\Sigma }`$, are identical on both sides of the hypersurface . The Darmois junction conditions allow us to use different coordinate systems on both sides of the hypersurface. The continuity of the first and second fundamental forms on a matching hypersurface $`\mathrm{\Sigma }`$ implies the continuity of the fluid pressure on $`\mathrm{\Sigma }`$ (see e.g. ). In the case of the “Swiss cheese” model, it implies a dust filled (i.e. zero pressure) FLRW space-time. Recently, Baker claimed that the extrinsic curvatures for the Schwarzschild and the FLRW metrics used in the “Swiss cheese” model cannot be matched at a spherical boundary. In the following section we show that this claim is erroneous. We prove this by explicitely constructing a smooth matching between the Schwarzschild and FLRW space-times across a spherical hypersurface. In particular, we show that both the intrinsic metrics and the extrinsic curvatures (henceforth often refered to as the first and second fundamental forms respectively) are continuous on the hypersurface. We also verify that the pressure is continuous as required. We conclude by indicating the error in ref. that led to the claim. ## 2 The Matching The general FLRW metric can be written in spherical coordinates as $$ds^2=dt^2K^2(t)\left[r^2\left(d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2\right)+\frac{dr^2}{1kr^2}\right],$$ (1) where $`K(t)`$ is the scale factor and $`k=0,\pm 1`$ the curvature constant of space. We will show that the metric (1) can be joined smoothly on a spherical hypersurface $`\mathrm{\Sigma }`$ to the Schwarzschild metric $$ds^2=\left(1\frac{2M}{\rho }\right)dT^2\rho ^2\left(d\theta ^2+\mathrm{sin}^2\theta d\varphi ^2\right)\left(1\frac{2M}{\rho }\right)^1d\rho ^2.$$ (2) The first fundamental form is the metric which $`\mathrm{\Sigma }`$ inherits from the space-time in which it is imbedded, and may be written as $$\mathrm{{\rm Y}}_{\alpha \beta }=g_{ij}\frac{x^i}{u^\alpha }\frac{x^j}{u^\beta },$$ (3) where $`u^\alpha =(u^1u,u^2v,u^3w)`$ is the coordinate system on the hypersurface. Greek indices run over $`1,\mathrm{},3,`$ while Latin indices over $`1,\mathrm{},4.`$ The second fundamental form is defined by $$\mathrm{\Omega }_{\alpha \beta }=(\mathrm{\Gamma }^p{}_{ij}{}^{}n_{p}^{}n_{i,j})\frac{x^i}{u^\alpha }\frac{x^j}{u^\beta },$$ (4) where $`n_a`$ is a unit normal to $`\mathrm{\Sigma },`$ and $`\mathrm{\Gamma }^p_{ij}`$ are the Christoffel symbols. If $`\mathrm{\Sigma }`$ is given by the function $`f[x^a(u^\alpha )]=0`$, then $`n_i`$ can be calculated from $$n_i=\frac{f_{,i}}{|g^{ab}f_{,a}f_{,b}|^{1/2}},$$ (5) where $`,i`$ denotes $`\frac{}{x^i}`$ . To avoid confusion we will denote indexed quantities associated with the FLRW and Schwarzschild metrics by the letters $`F`$ and $`S`$ respectively. We consider a spherical hypersurface $`\mathrm{\Sigma }`$ given by the function $`f_F(x_F^i)=rr_0=0,`$ where $`r_0`$ is a constant, and parametrized by $`x_F^1=t=u,x_F^2=\theta =v,x_F^3=\varphi =w,`$ and $`x_F^4=r=r_0,`$ in the FLRW frame. In the Schwarzschild frame we choose the parametrization $`x_S^1=T=T(u),x_S^2=\theta =v,x_S^3=\varphi =w,`$ and $`x_S^4=\rho =\rho (u).`$ The condition $`\mathrm{{\rm Y}}_{F\alpha \beta }=\mathrm{{\rm Y}}_{S\alpha \beta }`$ then implies $$1=\left(1\frac{2M}{\rho }\right)\left(\frac{dT}{du}\right)^2\left(1\frac{2M}{\rho }\right)^1\left(\frac{d\rho }{du}\right)^2,$$ (6) $$K^2r_0^2=\rho ^2.$$ (7) We now turn to the two second fundamental forms. The (outward pointing) unit normal in the FLRW frame can be calculated from eq. (5) and $`f_F(x_F^i)=rr_0=0.`$ The result is $`n_{Fi}=\delta _i^4n_{F4},`$ where $`n_{F4}=|g_{F44}|^{1/2}.`$ The unit normal is spacelike, i.e. $`n_F^in_{Fi}=1`$. The unit normal in the Schwarzschild frame cannot be obtained directly from eq. (5) since we do not know the form of $`f_S.`$ However, $`n_{Si}`$ must satisfy the two conditions $$n_S^in_{Si}n_F^in_{Fi}=1\text{ and }n_{Si}\frac{x_S^i}{u^\alpha }=0,$$ (8) where the second condition results from the partial differentiation of $`f_S[x_S^i(u^\alpha )]=0`$ with respect to $`u^\alpha `$. From (8) one obtains $$\begin{array}{c}n_{S2}=n_{S3}=0,\\ n_{S1}\frac{dT}{du}+n_{S4}\frac{d\rho }{du}=0,\\ \left(1\frac{2M}{\rho }\right)^1n_{S1}^{}{}_{}{}^{2}\left(1\frac{2M}{\rho }\right)n_{S4}^{}{}_{}{}^{2}=1.\end{array}\}$$ (9) With the help of eq. (6), equations (9) enable us to derive $`n_{Si}`$ as a function of $`u^\alpha :`$ $$n_{Si}=(ϵ\frac{d\rho }{du},\mathrm{\hspace{0.33em}0},\mathrm{\hspace{0.33em}0},ϵ\frac{dT}{du}),ϵ=\pm 1.$$ (10) Because of the simple form of $`n_{Fi},`$ eq. (4) for the second fundamental form can be much simplified in the FLRW frame. In this case one obtains from (4) $`\mathrm{\Omega }_{F\alpha \beta }`$ $`=`$ $`\mathrm{\Gamma }_{Fij}^4n_{F4}{\displaystyle \frac{x_F^i}{u^\alpha }}{\displaystyle \frac{x_F^j}{u^\beta }}n_{F4,j}{\displaystyle \frac{x_F^4}{u^\alpha }}{\displaystyle \frac{x_F^j}{u^\beta }}`$ (11) $`=`$ $`\mathrm{\Gamma }_{F\mu \nu }^4n_{F4}{\displaystyle \frac{x_F^\mu }{u^\alpha }}{\displaystyle \frac{x_F^\nu }{u^\beta }},\text{since}{\displaystyle \frac{x_F^4}{u^\alpha }}={\displaystyle \frac{r_0}{u^\alpha }}=0,`$ $`=`$ $`\mathrm{\Gamma }_{F\mu \nu }^4n_{F4}\delta _\alpha ^\mu \delta _\beta ^\nu `$ $`=`$ $`n_{F4}\mathrm{\Gamma }_{F\alpha \beta }^4`$ $`=`$ $`{\displaystyle \frac{1}{2}}|g_{F44}|^{1/2}g_F^{4i}(g_{F\alpha i,\beta }+g_{F\beta i,\alpha }g_{F\alpha \beta ,i})`$ $`=`$ $`{\displaystyle \frac{1}{2}}|g_{F44}|^{1/2}g_F^{44}g_{F\alpha \beta ,4},`$ so that $$\mathrm{\Omega }_{F\alpha \beta }=\frac{1}{2}|g_{F44}|^{1/2}g_{F\alpha \beta ,4}.$$ (12) Equation (12) with $`F^{}`$s dropped, i.e. $$\mathrm{\Omega }_{\alpha \beta }=\frac{1}{2}|g_{44}|^{1/2}g_{\alpha \beta ,4},$$ (13) is valid for any coordinate hypersurface $`x^4=\text{constant},`$ in an orthogonal coordinate system and parametrized by $`x^\alpha =u^\alpha .`$ No similar simplification of the second fundamental form of $`\mathrm{\Sigma }`$ is possible in the Schwarzschild frame. Moreover, for us to calculate the second term in eq. (4) we need $`n_{Si}`$ as a function of $`x_S^i.`$ However, if we once more differentiate the second condition in eq. (8) with respect to $`u^\alpha `$ it follows that $$n_{Si,j}\frac{x_S^i}{u^\alpha }\frac{x_S^j}{u^\beta }=n_{Si}\frac{^2x_S^i}{u^\alpha u^\beta },$$ giving $$\mathrm{\Omega }_{S\alpha \beta }=\mathrm{\Gamma }_{Sij}^pn_{Sp}\frac{x_S^i}{u^\alpha }\frac{x_S^j}{u^\beta }+n_{Si}\frac{^2x_S^i}{u^\alpha u^\beta }.$$ (14) Using equations (12) and (14), we find that $`\mathrm{\Omega }_{S\alpha \beta }=0=\mathrm{\Omega }_{F\alpha \beta },\alpha \beta .`$ The remaining, diagonal components of $`\mathrm{\Omega }_{\alpha \beta }`$ and the continuity condition $`\mathrm{\Omega }_{S\alpha \beta }=\mathrm{\Omega }_{F\alpha \beta }`$ on $`\mathrm{\Sigma },`$ result in the following three differential equations $`\mathrm{\Omega }_{S11}`$ $``$ $`\mathrm{\Gamma }_{S11}^4n_{S4}\left({\displaystyle \frac{dT}{du}}\right)^2+\mathrm{\Gamma }_{S44}^4n_{S4}\left({\displaystyle \frac{d\rho }{du}}\right)^2+2\mathrm{\Gamma }_{S14}^1n_{S1}{\displaystyle \frac{dT}{du}}{\displaystyle \frac{d\rho }{du}}`$ (15) $`+n_{S1}{\displaystyle \frac{d^2T}{du^2}}+n_{S4}{\displaystyle \frac{d^2\rho }{du^2}}=0\mathrm{\Omega }_{F11},`$ $`\mathrm{\Omega }_{S22}`$ $``$ $`\mathrm{\Gamma }_{S22}^4n_{S4}=\alpha |K|r_0\mathrm{\Omega }_{F22},`$ (16) $`\mathrm{\Omega }_{S33}`$ $``$ $`\mathrm{\Gamma }_{S33}^4n_{S4}=\alpha |K|r_0\mathrm{sin}^2\theta \mathrm{\Omega }_{F33},`$ (17) where the Christoffel symbols are given by $`\mathrm{\Gamma }_{S14}^1`$ $`=`$ $`\mathrm{\Gamma }_{S44}^4={\displaystyle \frac{M}{\rho (\rho 2M)}},`$ $`\mathrm{\Gamma }_{S11}^4`$ $`=`$ $`{\displaystyle \frac{(\rho 2M)M}{\rho ^3}},`$ $`\mathrm{\Gamma }_{S22}^4`$ $`=`$ $`(\rho 2M),`$ $`\mathrm{\Gamma }_{S33}^4`$ $`=`$ $`\mathrm{sin}^2\theta \mathrm{\Gamma }_{S22}^4,`$ and $`\alpha (1kr_0^2)^{1/2}.`$ Equations (16) and (17) being equivalent, we can use either of them and eq. (7) to obtain $$\frac{dT}{du}=\frac{ϵ\alpha \rho }{\rho 2M}.$$ (18) It then follows from eq. (6) that $$\left(\frac{d\rho }{du}\right)^2=\alpha ^2\left(\frac{\rho 2M}{\rho }\right).$$ (19) Differentiating eqs. (18) and (19) w.r.t. $`u`$, one obtains $$\frac{d^2T}{du^2}=ϵ\alpha \frac{2M}{(\rho 2M)^2}\frac{d\rho }{du}.$$ (20) and $$\frac{d^2\rho }{du^2}=\frac{M}{\rho ^2}.$$ (21) With equations (18)-(21), eq. (15) is now identically satisfied. Thus, both the first and second fundamental forms are continuous on $`\mathrm{\Sigma }.`$ It remains to verify that the pressure is also continuous across the spherical boundary. Since in the Schwarzschild space-time the pressure is zero, it must vanish in the FLRW space-time. The FLRW space-time is a perfect fluid space-time, and as such its Einstein field equations are given by $$G^{ij}R^{ij}\frac{1}{2}Rg^{ij}=8\pi \left[(\mu +p)u^iu^jpg^{ij}\right],$$ (22) where $`G^{ij}`$ is the Einstein tensor, $`R^{ij}`$ the Ricci tensor, $`R`$ the Ricci curvature scalar, $`\mu `$ the matter-energy density, $`p`$ the pressure, and $`u^i`$ the unit four-velocity. Specifically, eq. (22) implies $$G_{}^{1}{}_{1}{}^{}=\frac{3(\dot{K}^2+k)}{K^2}=8\pi \mu $$ (23) and $$G_{}^{2}{}_{2}{}^{}=G_{}^{3}{}_{3}{}^{}=G_{}^{4}{}_{4}{}^{}=\frac{2\ddot{K}}{K}+\frac{(\dot{K}^2+k)}{K^2}=8\pi p,$$ (24) where $`\dot{K}\frac{dK}{du}\frac{dK}{dt},`$ and $`\ddot{K}\frac{d^2K}{du^2}\frac{d^2K}{dt^2}.`$ To obtain $`\dot{K}`$ and $`\ddot{K},`$ we differentiate eq. (7) w.r.t. $`u.`$ The results are $$\dot{K}=\eta \frac{1}{r_0}\frac{d\rho }{du},$$ (25) and $$\ddot{K}=\eta \frac{1}{r_0}\frac{d^2\rho }{du^2},$$ (26) where $`\eta \frac{K}{|K|}.`$ Substituting eqs (19) and (21) into (25) and (26), and using $`K=\eta \frac{\rho }{r_0}`$ (from eq. (7)), it follows from (24) that $`p=0.`$ Finally, we note that, from eq. (23), $`8\pi \mu =\frac{6M}{r_0^3|R|^3},`$ a positive quantity, which shows that the matching is physically admissible. ## 3 Discussion and Conclusions We have shown, in a mathematically rigorous way, that an FLRW space-time can be joined smoothly to a Schwarzschild space-time across a spherical boundary, a construction used in the “Swiss cheese” cosmological model of the Universe. The boundary has a fixed coordinate radius in the FLRW frame, but changes with time in the Schwarzschild frame. In particular, we have shown that the intrinsic metric and extrinsic curvature are both continuous on the boundary as is required for a smooth, permanent matching. These matching requirements also imply the continuity of the pressure on the boundary, which has been verified here. Our results differ from those of ref. . Specifically, we have not found a discontinuity in the extrinsic curvature on the spherical boundary as claimed in . The claim in can be explained out as follows. The author, without justification, uses equation (13) (his eq. (2.10) with $`N_\alpha =0`$) for calculating the components of $`\mathrm{\Omega }_{\alpha \beta }`$ in both the FLRW and Schwarzschild frames. This results in the mismatch of $`\mathrm{\Omega }_{S\alpha \beta }`$ and $`\mathrm{\Omega }_{F\alpha \beta }`$ on the spherical hypersurface, leading to the claim that “the ’Swiss cheese’ model is at best only an approximation, with a singular interface”. The use of eq. (13) in the Schwarzschild frame is incorrect. Assuming, a priori, (as the author of ref. and we have done) that the boundary surface changes with time in the Schwarzschild frame, it follows that the boundary is not a coordinate boundary in that frame, and, therefore, eq. (4) for the extrinsic curvature, cannot be reduced to eq. (13). Equation (13) can be used for calculating $`\mathrm{\Omega }_{S\alpha \beta }`$ only under the assumption that the matching hypersurface in the Schwarzschild frame is the coordinate hypersurface $`\rho =\text{constant.}`$ But then eq. (7) cannot be satisfied unless $`R(t)=\text{constant;}`$ obviously not an interesting case, since then the FLRW space-time reduces to the Minkowski space-time. Thus the use of eq. (13) in the Schwarzschild frame is an error.
warning/0004/hep-ex0004028.html
ar5iv
text
# Measurements of Charm Fragmentation into 𝐷_𝑠^{∗+} and 𝐷_𝑠⁺ in 𝑒⁺⁢𝑒⁻ Annihilations at √𝑠=10.5 GeV ## I Introduction The production cross-sections of $`q\overline{q}`$ pairs in $`e^+e^{}`$ annihilations can be calculated using QCD, but the process of fragmentation whereby hadrons are formed is non-perturbative and phenomenological models are used to describe it. Two properties of hadron production that can be experimentally measured are the hadron momentum distribution and the relative population of available spin states. Measurements of primary hadron fragmentation can be challenging due to cascades from higher order resonances that can be indistinguishable from the primary hadrons. The study of $`D_s^+`$ and $`D_s^+`$ fragmentation in $`e^+e^{}`$ annihilations at $`\sqrt{s}=10.5`$ GeV benefits from the fact that $`L=1`$ charm mesons have not been observed to decay to either $`D_s^+`$ or $`D_s^+`$ and the influence of $`B`$ events is kinematically eliminated for $`x(D_s)>0.4`$, where $`x`$ is the $`D_s`$ momentum divided by the maximum kinematically allowed $`D_s`$ momentum. The $`D_s`$ system is thus particularly well suited for the measurement of the vector to pseudoscalar production ratio. The vector to pseudoscalar production ratio is usually described using the variable $$P_V=\frac{V}{V+P},$$ (1) where P and V represent, respectively, the number of pseudoscalar and vector mesons directly produced through a particular production mechanism, e.g. $`e^+e^{}`$ annihilations. Counting the number of spin states available to an $`L=0`$ meson leads to the expectation that $`P_V=0.75`$. This spin counting model has been shown to be useful for describing the $`D^+`$ spin alignment , but most measured values of $`P_V`$ have been significantly lower than 0.75 for charm mesons. Other models based upon the mass difference between the vector and pseudoscalar states predict values of $`P_V`$ that are less than 0.75 , but more precise measurements are needed to better determine any relationship between $`P_V`$ and the mass difference. ## II Detector and Event Selection The data in this analysis were collected from $`e^+e^{}`$ collisions at the Cornell Electron Storage Ring (CESR) by the CLEO II detector. The CLEO II detector is a general purpose charged and neutral particle spectrometer described in detail elsewhere . The dataset used in this analysis contains $`3.11\pm 0.03`$ fb<sup>-1</sup> of data collected at the $`\mathrm{{\rm Y}}(4S)`$ resonance and $`1.61\pm 0.02`$ fb<sup>-1</sup> of data collected below the $`b\overline{b}`$ threshold (about 60 MeV below the $`\mathrm{{\rm Y}}(4S)`$ resonance), for an approximate total of $`5\times 10^6`$ $`c\overline{c}`$ events. In this analysis, $`D_s^+`$ mesons are reconstructed via the decay $`D_s^+D_s^+\gamma `$ and $`D_s^+`$ mesons are reconstructed via the decay chain $`D_s^+\varphi \pi ^+`$ with $`\varphi K^+K^{}`$ (inclusion of charge conjugate modes is implied throughout this paper). All charged tracks used in this analysis are required to have an origin close to the $`e^+e^{}`$ interaction region and must be well reconstructed. When drift chamber particle identification information is available, the specific ionization, $`dE/dx`$, must be within two standard deviations of the expected value for candidate kaon tracks and within three standard deviations of the expected value for candidate pion tracks. Showers in the crystal calorimeter are considered as photon candidates if they have a minimum energy of 100 MeV, are within either the barrel ($`|\mathrm{cos}\theta _s|<0.71`$, where $`\theta _s`$ is the angle between the shower and the $`e^+`$ beam direction) or endcap ($`0.85<|\mathrm{cos}\theta _s|<0.95`$) regions, have an energy deposition consistent with that expected for a photon, and do not include any crystals near a projected charged track. Candidate $`\varphi `$ mesons are reconstructed using all appropriately signed combinations of candidate kaon tracks in an event. The invariant mass $`M(KK)`$ is required to be within 8.4 MeV/$`c^2`$ (approximately 2 standard deviations) of the known $`\varphi `$ mass . Candidate $`D_s^+`$ mesons are reconstructed using all combinations of $`\varphi `$ candidates and candidate pion tracks in an event. Candidate $`D_s^+`$ mesons are reconstructed using candidate photons, and $`\varphi \pi `$ combinations with invariant mass $`M(KK\pi )`$ within 20 MeV/$`c^2`$ (approximately 2.5 to 3 standard deviations) of the known $`D_s^+`$ mass. Because the $`\varphi `$ must be polarized in the helicity-zero state in a $`D_s\varphi \pi `$ decay, the decay of the $`\varphi `$ has an angular distribution proportional to $`cos^2\alpha `$, where $`\alpha `$ is the angle between the $`K^+`$ and $`D_s^+`$ momentum vectors in the $`\varphi `$ rest frame. Since the background angular distribution is flat, the signal to background ratio is improved by requiring $`|cos\alpha |>0.35`$. The signal to background ratio is further enhanced by requiring that $`cos\theta _\pi 0.8`$, where $`\theta _\pi `$ is the angle of the $`\pi `$ momentum vector in the $`D_s^+`$ rest frame relative to the $`D_s^+`$ momentum vector in the laboratory frame; the signal distribution is flat in this variable while background events peak at $`\mathrm{cos}\theta _\pi =1.0`$. Because of the minimum energy restriction for photon candidates, signal photons traveling in a direction opposite to the $`D_s^+`$ direction in the laboratory frame are excluded from the candidate sample. By requiring $`\mathrm{cos}\theta _\gamma >0.8`$, where $`\theta _\gamma `$ is defined as the angle of the photon momentum vector in the $`D_s^+`$ rest frame relative to the $`D_s^+`$ momentum vector in the laboratory frame, additional background $`D_s^+`$ candidates are suppressed. Low momentum $`D_s^+`$ candidates are difficult to analyze because of the large amount of background from combinatorics as well as $`B`$ decays. The analysis is therefore restricted to $`x(D_s^+)>0.44`$ where $$x(D_s^+)\frac{p(D_s^+)}{p_{max}(D_s^+)},$$ (2) and $$p_{max}(D_s^+)=\sqrt{E_{beam}^2m_{D_s^+}^2}.$$ (3) For $`D_s^+`$ candidates, the $`x(D_s^+)`$ requirement is replaced by $`x(D_s^+)>0.5`$ where $$x(D_s^+)\frac{p(D_s^+)}{p_{max}(D_s^+)},$$ (4) and $$p_{max}(D_s^+)=\sqrt{\left(E_{beam}\frac{m_{D_s^+}^2m_{D_s^+}^2}{4E_{beam}}\right)^2m_{D_s^+}^2}.$$ (5) In principle, $`BD_s^+\pi `$ can result in $`x(D_s^+)0.5`$. However, such decays are $`bu`$ transitions and thus heavily suppressed, so they are expected to be a negligible source of background. Based on the assumption that all observed $`D_s^+`$ are primary, the $`D_s^+`$ momentum spectrum is simply studied by measuring the $`D_s^+`$ yield in eight equal sized bins of $`x(D_s^+)`$ over the range $`0.5<x(D_s^+)<0.98`$. However, the observed $`D_s^+`$ can be primary or $`D_s^+`$ daughters. In order to study the momentum distribution of primary $`D_s^+`$ mesons, it is necessary to subtract out the $`D_s^+`$ contribution to the $`D_s^+`$ yields. Since all $`D_s^+`$ are assumed to decay to $`D_s^+`$, the $`D_s^+`$ yields from $`D_s^+`$ decays can be accounted for by simply measuring the $`D_s^+`$ yields as above, but in bins of the variable $`x(D_s^+)`$ rather than $`x(D_s^+)`$. After the $`D_s^+`$ yields are corrected for efficiency and the branching ratio $`(D_s^+D_s^+\gamma )`$, they are subtracted from the efficiency corrected $`D_s^+`$ yield in each $`x(D_s^+)`$ bin to calculate the primary $`D_s^+`$ yield. ## III Fitting The $`D_s^+`$ yields are projected onto $`\mathrm{\Delta }M=M(KK\pi \gamma )M(KK\pi )`$ for $`D_s^+`$ candidates and the $`D_s^+`$ yields are projected onto $`M(KK\pi )`$ for $`D_s^+`$ candidates. Fitting shapes for the peaks in these distributions are determined using a sample of Monte Carlo events generated using the Lund jetset 7.3 program combined with a geant-based CLEO II detector simulation, where every event contains a $`D_s^+`$ or $`D_s^+`$ decaying through the modes specified above. The $`\mathrm{\Delta }M`$ distributions in data and the signal Monte Carlo sample are simultaneously fit to the sum of an asymmetric Gaussian for the signal and separate second-order Chebyshev polynomials for the background in each distribution. An asymmetric Gaussian is used because of the larger tail on the lower side of the peak attributable to energy leakage in the calorimeter. The fits to data used to determine the $`D_s^+`$ yields in the selected regions of $`x(D_s^+)`$ and $`x(D_s^+)`$ are shown in Figs. 1 and 2. The $`M(KK\pi )`$ distributions in data and the signal Monte Carlo sample are simultaneously fit to the sum of a double Gaussian with common mean for the $`D_s^+`$ signal, a Gaussian for the $`D^+`$ signal, two straight lines joined by a quadratic for the combinatoric background in data, and a first order Chebyshev polynomial for the small amount of background in the Monte Carlo sample. The fits to data used to determine the $`D_s^+`$ yields in the selected regions of $`x(D_s^+)`$ are shown in Fig. 3. ## IV Efficiencies The $`D_s^+`$ and $`D_s^+`$ detection efficiencies are estimated using a sample of Monte Carlo events that contains signal as well as background events and is independent of the signal Monte Carlo sample used in the fitting procedure. The $`D_s^+`$ efficiency values in the $`x(D_s^+)`$ regions are listed in Table I, while the $`D_s^+`$ and $`D_s^+`$ efficiencies in the $`x(D_s^+)`$ regions are listed in Table II. For the $`D_s^+`$ production study, the efficiency for each $`x(D_s^+)`$ bin is measured using the fitting procedure described above. The binned raw efficiency values within the range $`0.50<x(D_s^+)<0.98`$ are fit with a first order Chebyshev polynomial to provide a smoothly varying efficiency as a function of $`x(D_s^+)`$. The smoothed efficiency value at the center of each $`x(D_s^+)`$ region is used to calculate the efficiency corrected $`D_s^+`$ yield and cross-section. For the $`D_s^+`$ fragmentation study, the $`D_s^+`$ and $`D_s^+`$ efficiencies are measured in each $`x(D_s^+)`$ bin using the fitting procedure described above. The binned raw $`D_s^+`$ efficiencies within the range $`0.44<x(D_s^+)<0.98`$ are fit with a first-order Chebyshev and the smoothed efficiency values are used to calculate the efficiency corrected $`D_s^+`$ yields. The binned raw $`D_s^+`$ efficiencies with $`0.44<x(D_s^+)<0.86`$ are fit with a first-order Chebyshev polynomial but the efficiencies in the region $`x(D_s^+)>0.86`$ are excluded from the fit because of expected efficiency loss due to the larger proportion of photons in that region with energies less than 100 MeV. The smoothed efficiency values are used to calculate the efficiency corrected $`D_s^+`$ yields for $`x(D_s^+)<0.86`$, while the raw efficiency values are used for $`x(D_s^+)>0.86`$. ## V Results The $`D_s^+`$ yields, efficiency corrected yields and cross-sections in the eight $`x(D_s^+)`$ regions are all listed in Table III. These same quantities for $`D_s^+`$ and $`D_s^+`$ in the nine $`x(D_s^+)`$ regions are listed in Tables IV and V, respectively. The calculated primary $`D_s^+`$ yields and cross-sections are presented in Table VI. By summing the efficiency corrected $`D_s^+`$ and primary $`D_s^+`$ yields listed in Tables III and VI, respectively, Eq. (1) could be used to calculate $`P_V`$ for $`x(D_s^{()+})>0.5`$. However, the uncertainties in the $`D_s^+`$ yields are essentially counted twice due to the subtraction used to calculate the primary $`D_s^+`$ yields. $`P_V`$ can however be calculated in a way that avoids this subtraction. Since all observed $`D_s^+`$ mesons are assumed to be primary, all observed $`D_s^+`$ mesons are assumed to be either primary or $`D_s^+`$ daughters, and all $`D_s^+`$ are expected to decay to a $`D_s^+`$, Eq. (1) can be rewritten as $$P_V=\frac{T(D_s^+)}{T(D_s^+)},$$ (6) where $`T(M)`$ is the total number of $`M`$ mesons in the CLEO II data sample. In terms of the quantities measured using the decay modes chosen for this analysis, $$P_V=\frac{n(D_s^+)}{n(D_s^+)(D_s^+D_s^+\gamma )},$$ (7) where $`n(M)`$ is the efficiency corrected yield of $`M`$ mesons in a particular $`x(D_s^+)`$ region. Using this method, $`P_V(x(D_s^+)>0.44)(D_s^+D_s^+\gamma )=0.42\pm 0.02`$. Using the value $`(D_s^+D_s^+\gamma )=(94.2\pm 2.5)\%`$ leads to $`P_V(x(D_s)>0.44)=0.44\pm 0.02(stat.)\pm 0.01(br.)`$. ## VI Systematic Uncertainty The systematic error for the total $`D_s^+`$ and $`D_s^+`$ yields is determined by varying the selection and fitting procedures as described below and taking the variance in the total yield as the estimate of the error. The variance is also determined on a bin-by-bin basis and the average percentage variance in the individual bins is taken as the estimated systematic uncertainty for all bins. The uncertainties in the $`D_s^+`$ yields for the range $`0.86<x(D_s^+)<1.0`$ are averaged separately since those values are not smoothed and the errors are quite large due to the limited number of $`D_s^+`$ events in that region. Systematic uncertainties on the various yields are listed in Tables VII and VIII. The acceptance angles for showers implicitly alter the acceptance of tracks since there is a high degree of correlation between the flight directions of the $`D_s^+`$ and the photon in the detector. There is also a correlation between the photon energy and the decay angle of the $`D_s^+`$. Varying the shower acceptance angles to $`|\mathrm{cos}\theta _s|<0.5`$ changes the total $`D_s^+`$ yields and the bin-by-bin yields by approximately 6%, while changing the minimum shower energy to either 90 MeV or 110 MeV changes the total $`D_s^+`$ yield by approximately 2% and the bin-by-bin yields by approximately 3%. A 3% overall systematic uncertainty in photon reconstruction has been estimated by comparing the world average value of $`(\eta \gamma \gamma )/(\eta 3\pi ^0)`$ with the relative yields of $`\eta \gamma \gamma `$ and $`\eta 3\pi ^0`$ in data and Monte Carlo. Additional uncertainty exists because of differences in invariant mass distributions between data and Monte Carlo and possible inadequacies of the fitting functions used to determine the yields. This uncertainty is estimated by altering the fitting shapes used to obtain the $`D_s^+`$ and $`D_s^+`$ yields. Varying the fitting technique for the $`M(KK\pi )`$ projections by e.g. using a Gaussian for the $`D_s^+`$ signal peak, a double Gaussian with common mean for the $`D^+`$ signal peak, or a second-order polynomial for the background alters the total $`D_s^+`$ yield by approximately 3% and the bin-by-bin yields by approximately 4%. Using a single Gaussian or double bifurcated Gaussian with a common mean for the peak in the $`\mathrm{\Delta }M`$ distribution alters the total $`D_s^+`$ yield by approximately 2% and the bin-by-bin yields by approximately 3%. There is also an uncertainty related to the requirement that $`M(KK\pi )`$ be within 20 MeV/$`c^2`$ of its nominal value. Widening this requirement to 25 MeV/$`c^2`$ and narrowing it to 15 MeV/$`c^2`$ has resulted in an approximate 2% error in the total $`D_s^+`$ yield and an approximate 4% error in the bin-by-bin yields. The uncertainties in the efficiency values shown in Tables I and II vary for each region of $`x(D_s^+)`$ and $`x(D_s^+)`$ due to limited Monte Carlo statistics and the smoothing process. For instance, the errors in the smoothed efficiency values near the limits of the $`x`$ region studied are higher than those in the middle of the region due to the uncertainty in the slope of the function used in the smoothing process. The errors in the efficiency contribute to the systematic uncertainty on a bin-by-bin basis and the percentage errors are added in quadrature for the determination of the percentage error for the total yields. All of the individual systematic uncertainties associated with a given yield are added together in quadrature with the percentage error in the efficiency to determine the total systematic uncertainty in the $`D_s^+`$ and $`D_s^+`$ yields. These systematic uncertainties are already included in the errors in the yields in Tables III, IV, V and VI. After including the total systematic uncertainty, $`P_V(x(D_s^+)>0.44)=0.44\pm 0.02(stat.)\pm 0.03(syst.)\pm 0.01(br.)`$. ## VII Discussion of Results The momentum distributions of hadrons created in the fragmentation process are commonly modeled with either the Andersson et al. symmetric fragmentation function or the Peterson et al. fragmentation function . Both of these functions depend upon $`z=\frac{E_h+p_{}}{E+p}`$, where $`E_h`$ is the energy of the hadron, $`p_{}`$ is the hadron momentum parallel to $`p`$, the momentum of the primary quark from the production process, and $`E`$ is the energy of the primary quark. The Andersson function is $$f(z)z^1(1z)^a\mathrm{exp}(bm_{}^2/z),$$ (8) where $`a`$ and $`b`$ are free parameters, $`m_{}=\sqrt{m_q^2+p_{}^2}`$, $`m_q`$ is the mass of the primary quark and $`p_{}`$ is the hadron momentum perpendicular to $`p`$. The Peterson function is $$f(z)\frac{1}{z[1(1/z)ϵ_P/(1z)]^2},$$ (9) where $`ϵ_P`$ is the single free parameter. To properly compare fragmentation models with data it is necessary to use the above functions in a full Monte Carlo simulation that incorporates photon radiation, gluon radiation and other effects. To facilitate comparison with other experimental results, $`x`$ is used as an approximation of $`z`$ and a binned $`\chi ^2`$ fit to the data is performed using these two functions as shown in Figs. 4 and 5. Since the parameters $`b`$ and $`m_{}`$ only appear in Eq. (8) as a product, the constraint $`b=1`$ has been used for the fit, thereby changing the interpretation of the value of $`m_{}`$. The numerical results from the fits are listed in Table IX. The normalizations of these fits are not used to calculate a value of $`P_V`$ due to differences between $`x`$ and $`z`$ that are non-negligible in the low $`D_s`$ momentum regime. The fragmentation spectra for charm mesons has been studied previously by the CLEO collaboration and input parameters for the Andersson et al. model were determined using measured fragmentation distributions for $`D^+`$, $`D^0`$, $`D^+`$, $`D_s`$ and $`\mathrm{\Lambda }_c`$. A comparison of the data presented here with a Monte Carlo distribution using the parameters determined in that study, $`a=0.60`$ and $`b=0.52`$, is shown in Figure 6. The use of these values as input parameters for charm fragmentation into $`D_s`$ mesons clearly result in a momentum distribution that is too soft, which is not surprising since P-wave charm meson decays to $`D^+`$, $`D^0`$ and $`D^+`$ were not excluded in the prior study. High levels of combinatoric background at low values of $`x(D_s^+)`$ prohibit a good measurement of $`P_V`$ for the full range of allowed $`D_s^+`$ momenta. Based on Monte Carlo simulations and the data presented, approximately $`7585\%`$ of all $`D_s^+`$ and $`D_s^+`$ are expected to have $`x(D_s^+)>0.44`$ and the value of $`P_V`$ presented here is not expected to differ much from $`P_V`$(all $`x`$). It is possible to make a model-dependent extrapolation of $`P_V`$(all $`x`$) using $$P_V(x>0)=\frac{1}{1+\left(\frac{1}{P_V(x>0.44)}1\right)\frac{Q_V}{Q_P}},$$ (10) where $`Q_V`$ is the percentage of $`D_s^+`$ that decay to a $`D_s^+`$ with $`x(D_s^+)>0.44`$ and $`Q_P`$ is the fraction of primary $`D_s^+`$ that have $`x(D_s^+)>0.44`$. Since only about one fifth of either fragmentation spectra lies below $`x(D_s^+)=0.44`$, and because both distributions approach zero smoothly as $`x(D_s^+)0`$, the ratio $`Q_V/Q_P`$ is expected to be close to unity and to only depend weakly upon the chosen fragmentation parameters. Using the Andersson et al. model with the parameters $`a=0.60`$ and $`b=0.52`$ results in $`Q_V=0.773`$, $`Q_P=0.795`$, $`Q_V/Q_P=0.972`$, and from Eq. (10), $`P_V`$(all $`x(D_s^+))=0.45\pm 0.05`$. Changing the input parameters to provide a harder spectrum has a very small effect on $`Q_V/Q_P`$. A distribution created with $`a=0.4`$ and $`b=0.9`$, for example, provides a much improved representation of the data and results in $`Q_V=0.860`$, $`Q_P=0.875`$, $`Q_V/Q_P=0.983`$ and $`P_V`$(all $`x(D_s^+))=0.45\pm 0.05`$. This clearly shows that the dependence of the $`P_V`$ extrapolation on the choice of fragmentation parameters is indeed weak. Based on the results of varying the input parameters for the two models, a systematic uncertainty of 3% is estimated for the model-dependent extrapolation resulting in a final extrapolated value of $`P_V`$(all $`x(D_s^+))=0.45\pm 0.05`$, which is significantly different than the expected result based on spin counting. Other measurements of $`P_V`$ for charm and bottom mesons have been presented , but it is difficult to make direct comparisons between those results and the one presented here because of differences in methodology and center-of-mass energies in the other analyses. Nonetheless, measurements of $`P_V(B)`$ are generally close to the spin-counting expectation while measurements of $`P_V(D)`$ are well below that value as shown in Table X. ## VIII Conclusion In summary, studies of $`D_s^+`$ and $`D_s^+`$ fragmentation in $`e^+e^{}`$ annihilations at $`\sqrt{s}=10.5`$ GeV have been presented. $`P_V(x(D_s)>0.44)`$ has been measured to be $`0.44\pm 0.02(stat.)\pm 0.03(syst.)\pm 0.01(br.)`$. When extrapolated to the entire available momentum region this measurement deviates significantly from $`P_V=0.75`$, the expected result based on simple spin counting. ## IX Acknowledgements We gratefully acknowledge the effort of the CESR staff in providing us with excellent luminosity and running conditions. J.R. Patterson and I.P.J. Shipsey thank the NYI program of the NSF, M. Selen thanks the PFF program of the NSF, M. Selen and H. Yamamoto thank the OJI program of DOE, J.R. Patterson, K. Honscheid, M. Selen and V. Sharma thank the A.P. Sloan Foundation, M. Selen and V. Sharma thank the Research Corporation, F. Blanc thanks the Swiss National Science Foundation, and H. Schwarthoff and E. von Toerne thank the Alexander von Humboldt Stiftung for support. This work was supported by the National Science Foundation, the U.S. Department of Energy, and the Natural Sciences and Engineering Research Council of Canada.
warning/0004/astro-ph0004065.html
ar5iv
text
# New challenges for Adaptive Optics: Extremely Large Telescopes ## 1 Introduction The current generation of large ground based optical telescopes has a diameter of the primary mirror in the 8 to 10 metre range. Recently some thoughts have been given to the next generation optical telescopes on the ground. In these projects the diameter of the primary mirror lies in a range between 40 and 100 metres (see Gilmozzi et al. 1998, Andersen et al. 1999, Mountain 1997). The use of Adaptive Optics (AO, Roddier 1999) in the visible is crucial to obtain the full potential in angular resolution, to avoid source confusion for extragalactic studies at high redshifts, and to reduce the background contribution, dramatically increasing limiting magnitude (the signal to noise ratio is then proportional to the square of telescope diameter). Competition with space based observatories, providing diffraction limited imaging on an 8 m class telescope (see Stockman 1997) is also a driver for AO correction in the visible with larger apertures. In this paper we address key issues for a visible light AO system on these Extremely Large Telescopes (ELTs). We have chosen a telescope diameter of 100 m, since it represents the extreme case and we want to investigate the limiting factors of AO on such a large aperture. We shall not address here the astrophysical drivers for such aperture size, which are presented elsewhere (Gilmozzi et al. 1998). We model the performance of an AO system working in the visible on a 100 m telescope, for an on-axis natural guide star (NGS) (section 2). The sky coverage with this approach is close to zero, because only bright objects (R $``$ 10) can be used as AO reference. The use of a single artificial laser guide star (LGS) is ruled out by the huge error introduced by the cone effect or focus anisoplanatism (Foy & Labeyrie 1985). We propose to use turbulence tomography (i.e. 3D mapping of turbulence, Tallon & Foy 1990, hereafter TF90) combined with Multi-Conjugate Adaptive Optics (Foy & Labeyrie 1985, Beckers 1988, hereafter MCAO) as a way to increase the fraction of the sky which can be observed. In section 3 we present the main concepts involved in turbulence tomography. In section 4 we describe a fundamental limitation of the corrected field of view size corrected by a small (1-3) number of deformable mirrors (DMs) and taking into account real turbulence profiles. A solution using 3 NGSs is presented, where the correction is done in the visible (section 5) and in the near-infrared (section 6). In section 7, another solution is presented, based on 4 LGSs for visible correction. In the following section, we present and quantify some technical aspects of AO on ELTs. Finally, in section 9, the conclusions are given. ## 2 AO performance with an on-axis NGS There is a strong scientific interest in visible light studies with the ELTs. Using the software described in Le Louarn et al. (1998) to perform analytical calculations of the AO system performance, we modeled a system with a Strehl ratio (ratio of the peak intensity of the corrected image to the peak intensity of a diffraction limited image, hereafter SR) of 60 % at 0.5 $`\mu `$m, based on a Shack-Hartmann wavefront sensor (e.g. Rousset 1994). The target SR is higher than required by the scientific goals, $``$ 40 %, to take into account potential error sources arising outside the AO system (e.g. aberration of the optics or co-phasing errors of the telescope primary mirror segments). Considering the current performance of AO systems, this is a challenging goal. However, the start of the operation of ELTs is planned in 10-20 years from now, and AO technology is bound to evolve considerably. The atmospheric model we used in these calculations corresponds to good observing conditions at Very Large Telescope observatory of Cerro-Paranal in Chile (Le Louarn et al. 1998). The main atmospheric parameters and the AO hardware characteristics are summarized in Tab. 1. The effects of scintillation on the wavefront sensing were neglected. Preliminary studies (Rousset 1999, private communication) have shown that the wavefront error contribution could be between 20 and 30 nm rms, reducing the SR by $``$ 10 %. The effects of the outer scale of turbulence were also neglected. Measurements (Martin et al. 1998) yield values usually between 20 to 30 m, significantly smaller than the diameter of the ELT. This is a new situation compared to current large telescopes. The effect of the outer scale is mainly to reduce the relative contribution of low order modes of wavefront distortions (Sasiela 1994) and to decrease the stroke needed for the DM to several microns, independently of telescope diameter. This relaxes constraints on the design of DMs, but does not change the overall on-axis system performance. The simulation results are presented in Fig. 1. The target SR of 60 % is obtained in the visible, providing 1.03 milli-arcsecond (mas) diffraction limit at 0.5 $`\mu `$m. The peak SR is over 95 % in K band (2.2 $`\mu `$m), the diffraction limit being 4.5 mas. Due to the fine wavefront sampling needed for correction in the visible, the limiting magnitude (at 0.5 $`\mu `$m) is $`R10`$, which is bright compared to current AO systems working in the near-infrared (around $`R16`$, e.g. Graves et al. 1998). This implies that with a single NGS the sky coverage is extremely small (see Rigaut & Gendron 1992, Le Louarn et al. 1998 for a more extensive discussion on sky coverage with AO systems). To overcome this limitation, we propose two different options, both involving multiple reference sources: NGS and LGS approaches are investigated in the following sections. ## 3 Turbulence tomography Turbulence tomography is a technique to measure the wavefront corrugations produced by discrete atmospheric turbulent layers with the help of several reference sources (TF90). Assuming weak turbulence, the phase corrugations produced by each layer add linearly (Roddier 1981). Knowing the configuration of the guide sources (position in the sky, height above ground in the case of an artificial star) and the altitudes of the layers to be measured, it is possible to reconstruct the phase at the selected turbulent layers. Foy & Labeyrie (1985) proposed to use Multiple DMs to correct them individually, a concept called Multi-Conjugate AO (MCAO). There must be at least as many measurements (number of guide stars times number of measurements points on the pupil) as there are unknowns (number of corrected layers times actuators on the correcting mirrors). Therefore, only a small number (2-4) of turbulent layers can be reconstructed, if a small number ($``$ 4) of reference sources are to be used. Recent papers have tackled the problems of turbulence tomography (TF90, Tallon et al. 1992, Ragazzoni et al. 1999, Fusco et al. 1999, Le Louarn & Tallon 2000), and reconstruction of turbulent wavefronts has been demonstrated in numerical simulations. The maximum size $`\theta `$ of the tomographic corrected Field of View (FOV) is given by geometrical considerations: $$\theta =\frac{D}{h_{max}}(1\frac{h_{max}}{H}),$$ (1) where $`D`$ is the diameter of the telescope, $`h_{max}`$ is the height of the highest turbulent layer and $`H`$ the height of the guide star (infinity for a NGS). As pointed out by TF90, in circular geometry, a small fraction of turbulence is not probed with this maximum FOV (pupil plane vignetting). This problem can be alleviated with a modal approach to turbulence tomography, which allows a slight interpolation of the wavefront within the corrected FOV (Fusco et al. 1999). With a 100 m telescope it may be possible to search reference stars in a much larger patch of the sky than with an 8 m class telescopes. The probability to find a references source can be dramatically increased (Ragazzoni 1999). For a 100 m telescope, the maximum tomographic field is 17′ in diameter with a NGS, or 13′ for LGSs, if the highest turbulent layer is at 20 km above ground. The image is corrected in the whole tomographic FOV only if the whole turbulence is concentrated in few thin layers and if each layer is optically conjugated to its correcting mirror. Taking into account real turbulence profiles, we compute in the next section the FOV size which can be corrected with few DMs and we show that it is much less than the tomographic FOV. ## 4 Limitations of multi-conjugate AO ### 4.1 Turbulence vertical profile measurements We have analyzed the PARSCA (Paranal Seeing Campaign, Fuchs & Vernin 1993) balloon data on the vertical distribution of turbulence to test the assumption that all turbulence is concentrated within a few layers. During the site testing campaign, 12 balloons were launched at nighttime to measure the profile of the refraction index constant, $`C_n^2(h)`$. SCIDAR (Scintillation Detection and Ranging, Azouit & Vernin 1980) measurements were also made simultaneously, confirming the balloon soundings (Sarazin 1996). In Tab. 2 we summarize some parameters of the balloon flights. The average Fried parameter (Fried 1966), $`r_0`$, was 19 cm at 0.5 $`\mu `$m, corresponding to a seeing of 0.55″– slightly better than the average seeing at Paranal, 0.65″. Considering the small time span during which the balloons were launched (19 days), these data are not fully representative of the site. The parameters have been corrected for the height difference between the observatory (2638 m), and the launching site (2500 m), which explains the slight difference with other publications (e.g. Sarazin 1996). In Fig. 2 the $`C_n^2`$ profiles obtained by the balloon flights are plotted. The height resolution of the balloons is $``$ 5 m. For clarity these measurements have been convolved with a Gaussian of standard deviation 500 m. The physics and formation of very thin turbulence laminae is described in Coulman et al. 1995. For most of the flights the thin turbulent layers form larger structures which can be identified with the turbulent layers seen by SCIDAR (see for example the concentration of turbulence near 15 km on flight 45, altitudes are expressed in kilometres above sea level). The strongest of these layers is the boundary layer, in the first kilometres of the atmosphere, present on all plots. Another layer, present on most flights, is located near 10-12 km. These measurements confirm the existence of numerous layers. However, a continuous component of small but significant amplitude is also present on most of the soundings. ### 4.2 Anisoplanatism in MCAO We used the high resolution profiles (not convolved with a Gaussian) and applied the analytical formula derived by Tokovinin, Le Louarn, & Sarazin (2000) to calculate the size of the FOV $`\theta _M`$ which can be corrected with $`M`$ deformable mirrors. This is a generalized isoplanatic angle in the sense of Fried (1982), expressed as $`\theta _M`$ $`=`$ $`[2.905(2\pi /\lambda )^2`$ (2) $`\times {\displaystyle }C_n^2(h)F_M(h,H_1,H_2,\mathrm{},H_M)\mathrm{d}h]^{3/5},`$ where $`F_M`$ is a function depending on the conjugation heights of the DMs, $`H_i`$ the height of conjugation, above ground. This expression assumes that the correction signals applied to each DM are optimized. It assumes an infinite turbulence outer scale and an infinite $`D/r_0`$ ratio. For 1 DM conjugated to altitude $`H_1`$, Eq. 2 contains: $$F_1(h)=|hH_1|^{5/3},$$ (3) which reduces to $`F_1(h)=h^{5/3}`$ if $`H_1=0`$ as in conventional AO and yields the classical $`\theta _0`$. For a two mirror configuration the function has the form: $`F_2(h,H_1,H_2)=0.5[|hH_1|^{5/3}+|hH_2|^{5/3}`$ $`0.5|H_2H_1|^{5/3}`$ $`0.5|H_2H_1|^{5/3}(|hH_1|^{5/3}|hH_2|^{5/3})^2]`$ For 3 or more DMs the expression for $`F_M`$ is much more complex. The heights $`H_i`$ were computed with a multi-parameter optimization algorithm to maximize $`\theta _M`$. We explored the possibilities with 1, 2 and 3 DMs in different altitude combinations, from all $`H_i`$ fixed to all $`H_i`$ optimized. In the optimized setups, the height of the mirrors were adapted for each flight to maximize the isoplanatic angle. For fixed DMs, we chose the conjugation height as the median of the heights found by optimization. The DM configurations are summarized in Tab. 3, and our results are shown in Fig. 3. With 3 DMs the increase in $`\theta _3`$ (compared to $`\theta _0`$) ranges from a factor of 2.6 to 13, depending on the profile. The median increase of $`\theta _3`$ is a factor of 7.7, which means that the isoplanatic angle in the visible increases from 2.2″ to 17″. On particular nights (Flight 43 for example, which has a lot of extended high altitude turbulence) $`\theta _3`$ stays small, $``$ 6″. The largest $`\theta _3`$ found was 28.9″ (Flight 38). A wavelength of 2.2 $`\mu `$m yields a median $`\theta _3`$ of 102″ (for comparison, $`\theta _0`$ = 13″). A two mirror configuration brings improvement factors between 1.6 and 8.7, with a median of 4.6. Therefore, with 2 DMs, one can expect to increase the isoplanatic angle to $``$ 10″ in the visible. Adapting the conjugate height of the DMs to profile variations is not crucial ( $`\theta _3`$ increases only by $``$ 7 % when using 3 optimized heights instead of fixed ones). Fig. 4 shows the optimal conjugate heights for the DMs as a function of flight number. Three main heights are identified: ground, $``$ 10–12 km and 15–20 km. Considering the observed stability of the optimum heights (due to the stability of the main turbulent layers), it is not surprising that optimizing the heights does not improve significantly the FOV. Notice the large deviation for point 2 (Flight 39). As shown in Fig 2, the turbulence was located very low, and the balloon reached only a maximum altitude of $``$ 15 km, leaving part of the turbulence unmeasured. These results show that anisoplanatic effects occur in the visible even with 3 DMs used in an MCAO approach. They represent only one site, on a relatively short timescale. Other sites with similar isoplanatic angles exist (e.g. the measurements at Maidanak, Uzbekistan, provide a median $`\theta _0`$ of 2.48″(Ziad et al. 2000)). Moreover, the $`\theta _M`$ computed here is somewhat pessimistic, since it contains a piston term (which reduces the isoplanatic angle but does not affect image quality) and does not take into account the finite number of corrected turbulent modes. This is similar to the effect seen with $`\theta _0`$, which overestimates isoplanatic effects (Chun 1998). Therefore, it is reasonable to expect a corrected FOV between 30″ and 60″ in diameter, in the visible. This is a considerable improvement over the few arcsecond isoplanatic field in the visible (roughly equal to $`\theta _0`$), but much less than the tomographic FOV given by Eq. 1. We suggest that the site where an ELT is built be optimized in terms of turbulence profiles, and not only total turbulence, as it used to be in previous surveys. It is more effective to correct a few strong layers (even if the total turbulence is higher), than a continuous repartition of lower amplitude turbulence. Indeed, comparing for example Flights 46 and 55 shows that a similar $`\theta _0`$ 1.7″can be well corrected with 3 DMs (Flight 46, $`\theta _3`$ 14″) if turbulence is concentrated in a few peaks (see Fig. 2), whereas a quasi-continuous turbulence benefits much less from MCAO correction (Flight 55, $`\theta _3`$9 ″). The location of the turbulent layers should also be as stable as possible, to minimize the changes in DM conjugate height. Of course, some other parameters of the site will have impact on the telescope performance, like the wind (which is likely to be an important factor on such a large structure). ### 4.3 Required field of view In tomographic wavefront sensing using LGSs the reference sources are placed at the edges of the corrected field (TF90). Therefore with 3 DMs the LGSs are positioned $`\theta _{LGS}=\theta _3`$ apart. The telescope FOV, $`\theta _{tel}`$, must however be larger (see Fig. 5) for the laser spots to be imaged by the telescope: $$\theta _{tel}=\theta _{LGS}+\frac{D}{H}.$$ (5) For a 100 m telescope and a sodium LGS placed at a 90 km height, and for $`\theta _{LGS}=`$60″, we get $`\theta _{tel}=`$ 290″, or almost 5′ in diameter. This can be a severe requirement for the telescope optical design. ## 5 Natural Guide Stars for visible correction The use of several NGSs on an ELT to increase the corrected FOV and to find reference stars outside the isoplanatic patch was proposed by Ragazzoni (1999). He pointed out that with turbulence tomography the maximum FOV which can be corrected increases linearly with telescope diameter, as shown by Eq. 1. Therefore, it would be possible to use the huge tomographic FOV to search for natural references. This work assumed that anisoplanatism was not present in turbulence tomography (turbulence concentrated in a few thin layers). In the previous paragraph we have shown that this is unfortunately not the case with real turbulence profiles. As a consequence, if the reference stars are much further away than $`\theta _M`$ they will not benefit from AO correction. The wavefront measurement would therefore be done in open-loop. This is a very unusual situation in AO (Roddier 1999), and experiments must be carried out to verify the feasibility of that approach. Moreover, our further studies show that for widely separated NGSs, the errors of tomographic wavefront reconstruction with real turbulence profiles can be very high. So the use of 3 NGSs in a wide tomographic field seems problematic. Still, we estimate the sky coverage for this option. Another constraint comes from the telescope design. The telescope FOV of an ELT is a strong cost driver and, at the moment, a full tomographic FOV (17′) does not seem to be feasible. Current optical designs for a 100 m telescope (Dierickx et al. 1999) provide a maximum FOV of 12′. We have computed the sky coverage (SC) for the case when reference stars are sought within a 12′ FOV (Fig. 6). Full SC is obtained only near the Galactic plane. A 60 % SC can be achieved with a SR of 0.2 at average Galactic coordinates ($`l=180^{}`$, $`b=20^{}`$), or 30 % near the pole. If a telescope design can be improved to have the maximum FOV allowed by tomography (Eq. 1) the SC will be significantly increased. A full SC can be achieved with a SR of 0.1 everywhere. SC of 50 % is achieved on the whole sky with a SR of at least 0.4. Given the performance of the AO system shown in Fig. 1, the telescope FOV size is identified here as a limiting factor for the sky coverage. Initially we presumed in these simulations that the limiting magnitude for 3 NGSs is the same as for one NGS, e.g. R $``$ 10 (Fig. 1). This is conservative with regards to the results obtained by Johnston & Welsh (1994): when using four reference stars, the flux from the individual reference sources could be divided by four, i.e. a gain of 1.5 magnitudes. We have therefore also studied the cases where the limiting NGS magnitudes were one and two magnitudes fainter. Such gains could be achieved by efficient tomographic reconstruction algorithms. If the limiting magnitude can be increased by one magnitude, a SC of 40 % at the Galactic pole and 90 % at average Galactic latitudes can be obtained with a SR of 0.2. With the maximum tomographic FOV, a SC of 50 % is obtained with a SR 0.5 at the Galactic pole. ## 6 Natural guide stars and correction in the infrared The main problem of the NGS approach in the visible is caused by residual anisoplanatism. This problem is alleviated when only correction in the infrared is needed. At 1.25 $`\mu `$m $`\theta _M`$ is increased by a factor of 3 compared to the visible (see Eq. 2). For $``$ 60″ FOV in the visible (diameter), a 3′ corrected FOV is obtained. The limiting magnitude, as shown by Fig. 1, increases from R $``$ 10 to R $``$ 13. The sky coverage is plotted in Fig. 7. It shows that with a Strehl ratio of 0.2, SCs of 0.4 %, 30 %, 100 % are obtained respectively at Galactic poles, at average latitudes and in the Galactic disk. If a 1<sup>m</sup> gain in limiting magnitude is obtained compared to a single NGS, the coverages increase only slighlty. At 2.2 $`\mu `$m, the FOV is $``$ 6′(diameter), the limiting magnitude is about R $``$ 15. The sky coverage is 10 % at the Galactic pole and complete elsewhere. ## 7 Laser Guide Stars For astronomical AO systems LGSs based on resonant scattering in the sodium layer (Foy & Labeyrie 1985) are usually considered because they provide the highest reference source available, reducing the cone effect (also called focus isoplanatism, Foy & Labeyrie 1985, Fried & Belsher 1994, Tyler 1994). This effect is due to the finite altitude of the laser guide star. It prevents obtaining high AO correction in the visible already with 8 m telescopes. ### 7.1 Power requirements The laser power requirements for current AO systems working in the near-IR is of about 5 W (Continuous-Wave, CW), providing LGS brightness equivalent to a $`9^m`$ guide star (Jacobsen et al. 1994; Max et al. 1997; Davies et al. 1998). The typical sub-aperture size for those systems is 60 cm. Scaling to the subaperture size in the visible (16 cm) to obtain similar performance, the power of the laser should be 14 times higher (assuming a linear scaling of the guide star brightness with laser power), or about 70 W (CW). This scaling does not take saturation of the sodium layer into account. Milonni et al. 1998 provide an analytical tool to compute the power requirement in the case of a pulsed laser for a given guide star brightness with saturation. Using pulsed laser characteristics of the Keck LGS implementation (Sandler 1999) – 11 kHz repetition rate, 100 ns pulse duration, – we infer that to receive the same number of photons as for a 70 W CW laser, a $``$ 175 W pulsed laser is needed. However, considering Fig. 1, we can see that a 9<sup>th</sup> magnitude guide star would provide a Strehl ratio of 40 %. Therefore, if a slight loss of the AO system performance is acceptable, a significantly smaller amount of laser power would be sufficient. One could instead use a Rayleigh-scattering based LGS system (Fugate et al. 1994). This has the advantage of being able to use any laser (producing a bright LGS at an arbitrary wavelength is currently not a problem, see Fugate et al. 1994). However, the low altitude of Rayleigh LGSs ($``$ 15 km) reduces its suitability for tomography. The position of the LGSs to obtain a zero corrected FOV (only the cone effect is removed) is: $$\theta _{null}=\frac{D}{H}.$$ (6) $`\theta _{null}`$ 23′ ($`D=`$ 100 m, $`H=`$ 15 km), whereas the maximum tomographic FOV (Eq. 1) allowed by the highest turbulent layer (10 km, optimistic considering Fig. 2) is $``$ 11′(for a guide star placed at 15 km). Therefore, the cone effect can not be fully corrected with only 4 Rayleigh LGSs on ELTs and we will not consider this option in the remainder of this paper. ### 7.2 Multiple sodium laser guide stars On a 100 m telescope the use of a single LGS is totally impossible because of the huge cone effect involved. The option of using multiple (4) sodium laser guide stars in a tomographic fashion has therefore been investigated. We should stress that LGSs are placed on the edges of the corrected FOV (TF90), and therefore the problem of open-loop wavefront measurements does not affect this approach (the required FOV is given by $`\theta _M`$). The problem with LGSs in turbulence tomography is that the wavefront tilt cannot be obtained from the LGS (Pilkington 1987) and propagates into the global reconstructed wavefront. In addition to global tilt, other low order modes (like forms of defocus and astigmatism) have to be measured from an NGS located in the reconstructed FOV (Le Louarn & Tallon 2000). Elaborate techniques have been proposed to measure the tilt from the LGS (see e.g. Foy et al. 1995, Ragazzoni 1996). Unfortunately, real time correction has not been demonstrated. If tilt can be retrieved, this problem disappears and full SC is achieved. To solve the problem of LGS tilt indetermination, we propose to use in conjunction with LGS a very low order wavefront sensor (for example a curvature sensor, Roddier et al. 1988) working on a faint NGS. The limiting magnitude with 19 sub-apertures (4 sub-apertures across the pupil) is currently of R $``$ 17 (Rigaut et al. 1998) on a 3.6 m telescope, with correction at 2.2 $`\mu `$m. In Tab. 4, we summarize the scaling factors to be taken into account to convert this limiting magnitude to that of a 100 m telescope with a correction in the visible. The limiting magnitude is R $``$ 22. This scaling is only valid if compensation is done in the visible, so that wavefront sensing benefits from the AO correction (Rousset 1994). Otherwise, as shown by Rigaut & Gendron (1992), there is no gain in limiting magnitude for low order wavefront sensing on a large aperture compared to 4 m class telescopes. We used a model of the Galaxy developed by Robin & Crézé (1986) to get the probability to find a star of a given magnitude within a given FOV. Considering the faint magnitudes this system will be able to use, we also took into account the density of galaxies in the sky. We used galaxy counts given by Fynbo et al. (1999), based on a combination of measurements from the Hubble Deep Fields (North and South (Williams et al. 1996)) and the ESO NTT deep field (Arnouts et al. 1999)). Near the Galactic pole galaxies become more numerous than stars for magnitudes fainter than $`R`$ 22. A bias may exist since not all of these galaxies can be used as a reference due to their size (a source size smaller than 4 mas was assumed in Tab. 4). However, usually, the fainter the galaxies the smaller they are. We have assumed that galaxies are distributed evenly in the sky. Poisson statistics give the probability to find a reference object for a given AO limiting magnitude. In Fig. 8 the probability to find an NGS within a field of 30″ in diameter is shown. The SC is $``$ 13 % for the Galactic pole at R$``$22. A twice larger corrected isoplanatic angle (60″ in diameter, Fig. 9), yields a SC of 40 % at the poles, 70 % at average latitudes and 100 % near the Galactic plane. Scaling the SR vs limiting magnitude of current curvature systems, we expect a SR between 0.2 and 0.4 for this reference magnitude. At a magnitude of $`R`$ 22, most of the wavefront references sources will be galaxies when observing near the Galactic pole. ## 8 Technical challenges In the previous sections, we have shown that there are no fundamental limitations imposed by the laws of atmospheric turbulence to building a visible light AO system on a 100 m optical telescope. In this section, we shall discuss the technical difficulties which have to be addressed to build such a system. ### 8.1 Wavefront sensor The number of sub-apertures of the wavefront sensor impose the use of a large detector. Centroiding computations require at least 2$`\times `$2 pixels per sub-aperture. For 16 cm sub-apertures, this means that the wavefront sensor detector must have at least 1250<sup>2</sup> pixels. Moreover, if guard pixels are used, this number could increase to 2500<sup>2</sup> (4$`\times `$4 pixels per sub-aperture). The pyramid wavefront sensor concept (Ragazzoni & Farinato 1999) requires only 2$`\times `$2 per sampling area and therefore could be an interesting alternative to a SH sensor. The detector noise requirement could be loosened slightly from the 1$`e^{}`$ level we have used, if bright LGSs can be created in the atmosphere. This is however unlikely, since saturation problems in the sodium layer will arise (see section 7.1). Currently, the state of the art detectors for wavefront sensors are 128<sup>2</sup> (Laurent et al. 2000). The required number of pixels could however be reduced by two means. One could use a curvature wavefront sensing method, coupled to a CCD detector. This approach has been proposed by Beletic, Dorn & Burke (1999) and has the advantage to reduce the number of pixels needed on the detector to one per sub-aperture. This would bring the total required number of pixels to $``$625<sup>2</sup>, which is realistic. It does not seem possible, with current technology, to produce a bimorph mirror (usually associated to curvature sensors) with 500000 actuators. This problem could be solved by coupling a curvature sensor to a piezo-stack deformable mirror but the approach clearly deserves more studies. The read-out rate of the wavefront sensor detector (SH) can be obtained by scaling the typical current frame-rate in the IR ($``$ 200 Hz) to the visible. We obtain a frame rate of $``$ 1.5 kHz. Therefore, there is a choice to be made between a smaller number of pixels but high frame-rate and a larger but slower system. Both large number of pixels and high read-out speed can be achieved by butting small chips together, with multiple read-out ports (like in the Nasmyth Adaptive Optics System wavefront sensor, Laurent et al. 2000), or even more efficiently by adapting the CCD designing technique described in Beletic et al. (1999) to Shack-Hartmann systems, which allows a very efficient parallelization of the read-out process. Therefore the wavefront sensor detector should not be technically the most challenging part of the AO system. ### 8.2 Deformable mirror With a typical DM diameter of 0.5 m which could be feasible on a 100 m telescope (Dierickx et al. 1999), the spacing requirement between the DM actuators would be 0.8 mm. This a value ten times smaller than on existing DMs. Therefore, the production of a DM with 500000 actuators clearly requires new methods. Current development based on MOEMS (Micro-Opto-Electro-Mechanical systems) could lead to spacings down to 0.3 mm, (e.g. Bifano et al. 1997, Vdovin et al. 1997, Roggeman et al. 1997) making possible a DM size of $``$20 cm. One of the key issues in the design of these DMs is the required stroke. Assuming an outer scale of turbulence of 25 m and a von Kármán model, a stroke of $`\pm 5\mu `$m (3 $`\sigma `$) would be sufficient. However, the actual turbulence spectrum at low spatial frequencies must be measured on 8 m class telescopes for realistic estimates of the required stroke. ### 8.3 Computing power By using Moore’s law, which states that the computing power doubles every 1.5 years, the computing power in 20 years will be increased by a factor of 10<sup>4</sup>. Current wavefront computers have a delay smaller than 200 $`\mu `$s, which is compatible with use in the visible (Rabaud et al. 2000). The required computing power increase can therefore be estimated as the squared ratio of the number of controlled actuators: $$\gamma =(\frac{N_{ELT}}{N_{IRAO}})^2,$$ (7) where $`N_{IRAO}`$ is the number of actuators of current IR AO systems (200), and $`N_{ELT}`$ the number of actuators for the ELT (500000). We get $`\gamma 6\times 10^6`$. However, this does not take into account that the cross-talk between actuators will be negligible for actuators far away from each other and therefore the interaction matrix will be very sparse. This will reduce significantly the computing load. If, for example, the interaction matrix (Boyer, Michaud & Rousset 1990) can be broken up into 6 times $`100\times 100`$ matrices, the likely evolution in technology would bring the adequate power in 20 years. Another possibility would be to use a curvature sensing, in which the interaction matrix is almost diagonal (if no modal control is employed), minimizing the computing power requirements. However, this approach, as noted earlier, seems to be prohibited by the availability of large bimorph mirrors. ### 8.4 Optics The use of a small pitch between the actuators of the DM allows to maintain a small pupil diameter: with a pitch of 300 $`\mu `$m, the pupil size is 187 mm. This facilitates the imaging of the pupil on the wavefront sensor detector. Indeed, with 625 sub-apertures across the pupil, the WFS detector-size is roughly 25 mm (assuming 2 pixels per sub-aperture and 20 $`\mu `$m pixels). The reduction factor from the pupil to the detector is then 7.5, which is not a problem if each subaperture has a FOV of a few arcseconds. Atmospheric dispersion (AD) correction is currently an unsolved problem and has to be tackled at the level of telescope design. For example, AD produces an elongation of the object of 184 mas (if AD is not corrected, assuming imaging between 0.5 and 0.6$`\mu `$m, at a zenith angle of 30) which is unacceptably high. The design of the AD corrector will be challenging, since an optimal combination of glasses, allowing a correction with an accuracy better than 1 mas must be found. The physical sizes of these AD correctors is also a problem, because of the large size of the optics. The required precision puts severe constraints on the measurement of atmospheric parameters (air temperature, humidity, pressure). For the multi-NGS scheme this problem is even more crucial, since the NGSs must be far apart to increase the sky coverage, and will therefore suffer immensely from AD. The multi-LGS has the advantage to be in sensitive to AD, because the sources are highly monochromatic. If proper correctors cannot be built for technological reasons, narrow band operation of the telescope should be used if the highest spatial resolution is required: at 30 from zenith a bandpass of 0.4 nm produces a dispersion of $``$1 mas, if no correction is made. The use of 3D detectors (like integral field spectrographs), would solve the problem, since images in different colors can then be disentangled. In the multi-NGS case, if anisoplanatism limits the correction and the sources do not benefit from AO correction, non-common path aberrations between the sources will be difficult to maintain. ### 8.5 Laser spot elongation The atmospheric sodium layer is roughly 10 km thick (e.g. Papen, Gardner & Yu 1996). This causes the LGS to be extended, for sub-apertures which are not on the optical axis of the telescope (assuming a projection of the LGS from behind the secondary mirror of the telescope). The apparent size of a laser spot is given by simple geometry: $$\theta _{spot}\frac{\mathrm{\Delta }Hd}{H_{Na}^2}$$ (8) where $`\mathrm{\Delta }H`$ is the thickness of the sodium layer (10 km), $`H_{Na}`$ the altitude of the Sodium layer ($``$ 90 km), $`d`$ is the separation of the beam-projector and the considered sub-aperture. With $`d`$=50 m we get $`\theta _{spot}13`$″. Since the multiple LGSs will be off-axis, the spots will be even more elongated. This is clearly too large for standard wavefront sensors, which typically have a field of view of 2-3″. Several methods have been proposed to eliminate spot elongation. The conceptually simplest is to use a pulsed laser and to select only a small portion of the laser stripe by time gating the photons coming from the LGS. This has the advantage of being technically simple, at the cost of the effective brightness of the LGS. Other solution have been proposed in the literature (e.g. Beckers (1992), improving the previous scheme by shifting the wavefront sensor measurements synchronously with the propagation of the beam in the sodium layer, thus removing the loss of photons at the price of complexity. Other less technically challenging solutions should certainly be investigated. ## 9 Conclusions Although a realization of an adaptive optical system working with a 100 m telescope in the visible represents a technical challenge, it is shown here that very large aperture opens a number of new possibilities and such a correction becomes feasible for a significant fraction of the sky. The new approaches involve either use of several widely spaced bright NGS (in the near IR) or a very faint NGS combined with few LGS. In both cases a 3-D tomographic measurement of instantaneous phase screens is needed. Wavefront correction will be made with few (2-3) DMs conjugated to the optimum heights; in this way the FOV size is increased $``$ 8 times compared to the single-DM AO systems, and FOV diameter may reach 1′ in the visible. Additional criteria for site selection related to operation in this mode are formulated. ## acknowledgements The authors would like to thank Roberto Ragazzoni for many useful discussions, Johan Fynbo for his data on the magnitude distribution of galaxies. This paper benefitted from many discussions with B. Delabre, Ph. Dierickx and R. Gilmozzi regarding the design of 100 m telescopes. We are also grateful to an anonymous referee for improving the quality of this paper. This work was done with the help of the European TMR network “Laser guide star for 8-metre class telescopes” of the European Union, contract #ERBFMRXCT960094.
warning/0004/hep-ph0004176.html
ar5iv
text
# NUCLEAR GEOMETRY OF JET QUENCHING ## 1 Introduction The experimental investigation of ultrarelativistic nuclear collisions offers a unique possibility of studying the properties of strongly interacting matter at the high energy density when the hadronic matter is expected to become deconfined and a gas of asymptotically free quarks and gluons is formed. This is called quark-gluon plasma (QGP), in which the colour interactions between partons are screened owing to collective effects (see, for example, reviews ). In recent years, a great deal of attention has been paid to the study of ”hard” probes of QGP – heavy quarkonia and hard partonic jets, which do not appear as constituents of the thermalized system, but can carry information about the earliest stages of its evolution. In particular, the strong suppression of yield of heavy quark vector mesons as $`J/\mathrm{\Psi }`$, $`\mathrm{\Psi }^{}`$ ($`c\overline{c}`$ states) and $`\mathrm{{\rm Y}}`$, $`\mathrm{{\rm Y}}^{}`$, $`\mathrm{{\rm Y}}^{\prime \prime }`$ ($`b\overline{b}`$ states) is one of the promising signatures of the quark-gluon plasma formation in heavy ion collisions . An intriguing phenomenon is the ”anomalously” small yield of $`\mathrm{\Psi }`$-resonances, observed in Pb-Pb collisions in the NA50 experiment (CERN-SPS) and inconsistent with the conventional model of pre-resonance absorption in cold nuclear matter. Although the interpretation of this phenomenon as a result of the formation of a QGP is quite plausible , alternative explanations have also been put forward, such, for example, as $`\mathrm{\Psi }h`$ rescattering on comoving hadrons . Thus the nature of this ”anomalous” suppression of $`\mathrm{\Psi }`$-resonance production is not yet fully understood, and it should be completely explained in future . For heavier ($`b\overline{b}`$) systems, a similar suppression effect in super-dense strongly interacting matter is expected at higher temperatures than for $`c\overline{c}`$, which are expected to be reached in central collisions of heavy ions at the RHIC at BNL and LHC at CERN colliders. Along with the suppression of heavy quarkonia, one of the processes which may give information about the earliest stages of evolution of the dense matter formed in ultrarelativistic nuclear collisions is the passage through the matter of hard jets of colour-charged partons, pairs of which are created at the very beginning of the collision process (typically, at $`\begin{array}{c}<\hfill \\ \hfill \end{array}0.01`$ fm/c) as a result of individual initial hard nucleon-nucleon (parton-parton) scatterings. Such jets pass through the dense parton matter formed due to mini-jet production at larger time scales ($`0.1`$ fm/c), and interact strongly with the comoving constituents in the medium, changing its original properties as a result of additional rescatterings. The inclusive cross section for hard jet production processes is still very small for performing a systematic analysis at the SPS energies ($`\sqrt{s}20`$ GeV per nucleon pair), but it increases fast with the energy of collided nuclei. Thus these will play important role in the formation of the initial state at the energies of RHIC ($`\sqrt{s}=200`$ GeV per nucleon pair) and LHC ($`\sqrt{s}=5.5`$ TeV per nucleon pair) colliders. The actual problem is to study the energy losses of a hard jet evolving through the dense matter. We know two possible mechanisms of energy losses: $`(1)`$ radiative losses due to gluon ”bremsstrahlung” induced by multiple scattering and $`(2)`$ collisional losses due to the final state interactions (elastic rescatterings) of high $`p_T`$ partons off the medium constituents . Since the jet rescattering intensity strongly increases with temperature, formation of a super-dense and hot partonic matter in heavy ion collisions (with initial temperature up to $`T_01`$ GeV at LHC ) should result in significantly larger jet energy losses as compared with the case of ”cold” nuclear matter or hadronic gas at $`T\begin{array}{c}<\hfill \\ \hfill \end{array}0.2`$ GeV. Although the radiative energy losses of a high energy parton have been shown to dominate over the collisional losses by up to an order of magnitude , a direct experimental verification of this phenomenon remains an open problem. Indeed, with increasing of hard parton energy the maximum of the angular distribution of bremsstrahlung gluons has shift towards the parent parton direction. This means that measuring the jet energy as a sum of the energies of final hadrons moving inside an angular cone with a given finite size $`\theta _0`$ will allow the bulk of the gluon radiation to belong to the jet and thus the major fraction of the initial parton energy to be reconstructed. Therefore, the medium-induced radiation will, in the first place, soften particle energy distributions inside the jet, increase the multiplicity of secondary particles, but will not affect the total jet energy. It was recently shown that the radiation of energetic gluons in a QCD medium is essentially different from the Bethe-Heitler independent radiation pattern. Such gluons have formation times exceeding the mean free path for QCD parton scattering in the medium. In these circumstances the coherent effects play a crucial role leading to a strong suppression of the medium-induced gluon radiation. This coherent suppression is a QCD analogue of the Landau-Pomeranchuk-Migdal effect in QED. It is important to notice that the coherent LPM radiation induces a strong dependence of the jet energy on the jet cone size $`\theta _0`$ . On the other hand, the collisional energy losses represent an incoherent sum over all rescatterings. It is almost independent of the initial parton energy. Meanwhile, the angular distribution of the collisional energy loss is essentially different from that of the radiative one. The bulk of ”thermal” particles knocked out of the dense matter by elastic scatterings fly away in almost transverse direction relative to the hard jet axis. As a result, the collisional energy loss turns out to be practically independent on $`\theta _0`$ and emerges outside the narrow jet cone. Thus the relative contribution of collisional losses would likely become significant for jets with finite cone size propagating through the QGP . In a search for experimental evidences in favour of the medium-induced energy losses a significant dijet quenching (a suppression of high-$`p_T`$ jet pair yield) and a monojet-to-dijet ratio enhancement were proposed as possible signals of dense matter formation in ultrarelativistic collisions of nuclei. Other possible signatures that could directly measure the energy losses involve tagging the hard jet opposite a particle that does not interact strongly as a $`Z`$-boson (mostly $`q+gq+Z(\mu ^+\mu ^{})`$, but also $`q+\overline{q}q+Z`$ ) or a photon (mostly $`q+gq+\gamma `$, also $`q+\overline{q}q+\gamma `$). The jet energy losses in dense matter should result in the non-symmetric shape of the distribution of differences in $`P_T`$ between the Z-boson ($`\gamma `$) and jet. The above phenomena can be studied in heavy ion collisions with Compact Muon Solenoid (CMS), which is the general purpose detector designed to run at the LHC . Note, that using $`\gamma +jet`$ channel in this case is complicated due to large background from $`jet+jet`$ production when one of the jet in an event is misidentified as a photon (the leading $`\pi ^0`$). However the shape of the distribution of differences in $`E_T`$ between the $`\gamma `$ and jet is very different for signal and background, and still sensitive to the jet energy losses . The advantage of $`\gamma +jet`$ and $`Z(\mu ^+\mu ^{})+jet`$ channels is that one can determine the average initial transverse momentum of the hard jet, $`P_T^{jet}P_T^{\gamma ,Z}`$. It gives the attractive possibility to search for coherent effects in QCD-medium: the dependence of energy losses of the distance traversed can be studied experimentally in different bins of impact parameter distribution of nucleus-nucleus collision, or by varying collided ions and selecting the most central collisions. The intriguing prediction associated with the coherence pattern of the medium induced radiation is that radiative energy losses per unit distance $`dE/dx`$ depend on the total distance traversed $`L`$ . The value $`dE/dx`$ is approaching to being proportional to $`L`$ for static medium , and it has weaker $`L`$-dependence for the case of expanding medium . The main goal of the present paper is to analyze the possibility of observing the $`L`$-dependence of jet energy losses $`dE/dx`$ for realistic nuclear geometry. In particular, we are studying the impact parameter dependence of collisional and radiative jet energy losses in dense QCD-matter, created in ultrarelativistic heavy ion collisions. ## 2 The geometrical model for jet production in nuclear collisions Let us to consider the simple geometrical model of jet production and jet passing through a dense matter in high energy symmetric nucleus-nucleus collision. The figure 1 shows the essence of the problem in the plane of impact parameter b of two colliding nuclei $`A`$-$`A`$. The impact parameter $`b`$ here is the transverse distance between nucleus centers $`O_1`$ and $`O_2`$, $`OO_2=O_1O=b/2`$. Let $`B(r\mathrm{cos}\psi ,r\mathrm{sin}\psi )`$ be denoted as a jet (dijet) production vertex, with $`r`$ being the distance from the nuclear collision axis to the $`B`$. Then the distance between nucleus centers ($`O_1,O_2`$) and vertex $`B`$ can be found as $$r_{1,2}=\sqrt{r^2+\frac{b^2}{4}\pm rb\mathrm{cos}\psi }.$$ (1) The distribution over jet production vertex $`B(r,\psi )`$ at given impact parameter $`b`$ is written as $$P_{AA}(𝐫,b)=\frac{T_A(r_1)T_A(r_2)}{T_{AA}(b)},$$ (2) where $$T_{AA}(b)=d^2𝐬T_A(𝐬)T_A(𝐛𝐬)=\underset{0}{\overset{2\pi }{}}𝑑\psi \underset{0}{\overset{r_{max}}{}}r𝑑rT_A(r_1)T_A(r_2)$$ (3) is the nuclear overlap function, $`T_A(𝐫)=A\underset{\mathrm{}}{\overset{+\mathrm{}}{}}\rho _A(𝐫,z)𝑑z`$ is the nuclear thickness function with nucleon density distribution $`\rho _A(𝐫,z)`$. The maximum possible value of $`r`$ in nuclear overlapping zone can be estimated from the equation $$max\{r_1(r=r_{max}),r_2(r=r_{max})\}=R_A$$ (4) ($`R_A`$ is the radius of the nucleus $`A`$). This gives $$r_{max}=min\{\sqrt{R_A^2\frac{b^2}{4}\mathrm{sin}^2\psi }+\frac{b}{2}\mathrm{cos}\psi ,\sqrt{R_A^2\frac{b^2}{4}\mathrm{sin}^2\psi }\frac{b}{2}\mathrm{cos}\psi \}.$$ (5) In particular, for the uniform nucleon density distribution, $`\rho _A^{un}(𝐑)=\rho _0\mathrm{\Theta }(R_A|𝐑|)`$, the nuclear overlap function is equal to $`T_A^{un}(r)=3A\sqrt{R_A^2r^2}/(2\pi R_A^3)`$. Then the distribution $`P_{AA}^{un}(𝐫,b)`$ is proportional to $$P_{AA}^{un}(𝐫,b)\sqrt{R_A^2r_1^2(r,\psi ,b)}\sqrt{R_A^2r_2^2(r,\psi ,b)}.$$ (6) For central $`AA`$ collisions ($`b=0`$, $`r_{max}=R_A`$) we get simply $`P_{AA}^{un}(𝐫,b=0)(R_A^2r^2)`$. It is straightforward to evaluate the time $`\tau _L=L`$ it takes for jet to traverse the dense zone: $$\tau _L=min\{\sqrt{R_A^2r_1^2\mathrm{sin}^2\phi }r_1\mathrm{cos}\phi ,\sqrt{R_A^2r_2^2\mathrm{sin}^2(\phi \phi _0)}r_2\mathrm{cos}(\phi \phi _0)\},$$ (7) where $`\phi `$ is the azimuthal angle which determines the direction of a jet motion in the transverse plane, and $`\phi _0`$ is the angle between vectors $`𝐫_\mathrm{𝟏}`$ and $`𝐫_\mathrm{𝟐}`$. The expression for $$\phi _0=\mathrm{arccos}\frac{r^2b^2/4}{r_1r_2}$$ (8) can be obtained from the condition $$r_1r_2\mathrm{cos}\phi _0=𝐫_\mathrm{𝟏}𝐫_\mathrm{𝟐}=(b/2r\mathrm{cos}\psi )(b/2r\mathrm{cos}\psi )+r^2\mathrm{sin}^2\psi =r^2b^2/4.$$ (9) Finally, we are going to estimate the dependence of initial energy density in nuclear overlapping zone on impact parameter of the collision. At collider energies the minijet system (the semi-hard gluons, quarks and antiquarks with $`p_T\begin{array}{c}>\hfill \\ \hfill \end{array}p_01÷2`$ GeV/c) in the central rapidity region is typically formed in parton-parton scatterings at very early times, $`\tau _01/p_T\begin{array}{c}<\hfill \\ \hfill \end{array}1/p_00.1`$ fm/c, and it will then serve as initial condition for the further evolution of the system . Strictly speaking, the soft particle production mechanisms (like the decay of the colour field) can also contribute to initial conditions in nuclear interactions. However, the relative strength of soft part decreases strongly with increasing c.m.s. energy of the ion beams. In particular, at LHC energies $`\sqrt{s}=7`$ TeV$`\times (2Z/A)`$ per nucleon pair the hard and semi-hard processes contribute over $`80\%`$ to the transverse energy in heavy ion collisions . Moreover, soft processes with small momentum transfer $`Q^2\mathrm{\Lambda }_{QCD}^2(200`$ MeV$`)^2`$ $`p_0^2`$ can be partially or fully suppressed, owing to screening of the colour interaction in the dense parton matter produced from the system of minijets in the early stages of the reaction . Therefore, at LHC energies, we will consider only dominant semi-hard contribution to the formation of initial state. The initial energy density inside the comoving volume of longitudinal size $`\mathrm{\Delta }z=\tau _02\mathrm{\Delta }y`$ can be estimated using the Bjorken formula as $$\epsilon (\tau =\tau _0)=\frac{E_T^A(|y|<\mathrm{\Delta }y)}{S(b)\mathrm{\Delta }z}=\frac{E_T^A(|y|<\mathrm{\Delta }y)p_0}{S(b)2\mathrm{\Delta }y},$$ (10) where $$S_{AA}(b)=\underset{0}{\overset{2\pi }{}}𝑑\psi \underset{0}{\overset{r_{max}}{}}r𝑑r=\left(\pi 2\mathrm{arcsin}\frac{b}{2R_A}\right)R_A^2b\sqrt{R_A^2\frac{b^2}{4}}$$ (11) is the effective transverse area of nuclear overlapping zone at impact parameter $`b`$. The total initial transverse energy deposition in mid-rapidity region can be calculated as $$E_T^A(b,\sqrt{s},p_0,|y|<\mathrm{\Delta }y)=T_{AA}(b)\sigma _{NN}^{jet}(\sqrt{s},p_0)p_T,$$ (12) where the first $`p_T`$-moment of inclusive differential minijet cross section $`\sigma _{NN}^{jet}p_T`$ is determined by the dynamics of nucleon-nucleon interactions at the corresponding c.m.s. energy. Then the dependence of initial energy density $`\epsilon _0`$ in nuclear overlapping zone on impact parameter $`b`$ has the form: $$\epsilon _0(b)T_{AA}(b)/S_{AA}(b),$$ (13) or $$\epsilon _0(b)=\epsilon _0(b=0)\frac{T_{AA}(b)}{T_{AA}(b=0)}\frac{S_{AA}(b=0)}{S_{AA}(b)}.$$ (14) For central $`AA`$ collisions we have $`S_{AA}(b=0)=\pi R_A^2`$ and $`T_{AA}(b=0)=9A^2/(8\pi R_A^2)`$. It is worth noting that although this simple geometrical model for jet production in nucleus-nucleus collisions is formally can be applicable up to impact parameter $`b=2R_A`$, the major informative domain of our interest is central and semi-central collisions with $`b\begin{array}{c}<\hfill \\ \hfill \end{array}R_A`$ only. We have the following reasons in favour of this. 1) The contribution of such events to total jet rate is dominant, although these events represent only a few percents of total inelastic $`AA`$ cross section . For example, the $`PbPb`$ collisions with impact parameter $`b<0.9R_{Pb}=6`$ fm contribute $`50\%`$ to the total dijet rate at LHC energy, their relative fraction of total cross section being only $`10\%`$ in this case . 2) In the most central heavy ion collisions the maximum initial energy density is expected to be achieved in a fairly large (compared with typical hadronic scales) volume, when the effect of super-dense and hot matter formation, like quark-gluon plasma, can be really observable. The result for impact parameter dependence of initial energy density $`\epsilon _0`$ (14) in nuclear overlapping zone for uniform nucleon density is shown in figure 2: it is very weakly dependent of $`b`$ ($`\delta \epsilon _0\begin{array}{c}<\hfill \\ \hfill \end{array}10\%`$) up to $`bR_A`$, and decreases rapidly at $`b\begin{array}{c}>\hfill \\ \hfill \end{array}R_A`$. On the other hand, the averaged over all possible jet production vertices proper time $`\tau _L`$ (7) of jet escaping from the dense zone is found to go down almost linearly with increasing impact parameter $`b`$ (see the second curve in fig.2). Therefore the variation of impact parameter $`b`$ of nucleus-nucleus collision (which can be measured, for example, using the total transverse energy deposition detected in different parts of calorimeters ) up to $`bR_A`$ gives the possibility to study jet quenching as a function of distance traversed without significant changing initial energy density $`\epsilon _0`$. Meanwhile, the weakness of $`b`$-dependence of $`\epsilon _0`$ gives us the advantage as compared with using of beams of different ions at a fixed bin of impact parameter distribution, when the scaling $`\tau _L(b=0)R_AA^{1/3}`$ exists. Eq.(13) gives $`\epsilon _0(b=0)A^2/R_A^4`$, i.e. $`\epsilon _0(b=0)A^{2/3}`$. Figure 3 illustrates changing of average time $`\tau _L`$ of jet travel and initial energy density $`\epsilon _0`$ in dense zone with variation of impact parameter $`b`$ (at fixed $`A=207`$, Pb) and atomic weight $`A`$ (at fixed $`b=0`$). For example, decreasing $`\tau _L(Pb,b=0)6`$ fm by the factor $`1.7`$ can be obtained by: a) increasing $`b`$ up to $`b=0.9R_{Pb}6`$ fm (at the expense of only $`10\%`$ of $`\epsilon _0`$ reduction); b) decreasing $`A`$ down to $`A=40`$ (Ca) (at the expense of $`\epsilon _0`$ reduction by the factor $`3`$). 3) It is well known, that the uniform nucleon density distribution in the nucleus, $`\rho _A^{un}(𝐑)=\rho _0\mathrm{\Theta }(R_A|𝐑|)`$, can serve as a good approximation for central and semi-central collisions (see figure 4, which shows the nuclear overlap function profile for the uniform <sup>1</sup><sup>1</sup>1Moreover, in this case the explicit form of $`T_{AA}(b)`$ can be obtained: $$T_{AA}^{un}(b)=T_{AA}^{un}(b=0)\left[1\stackrel{~}{b}\left[1+\left(1\frac{\stackrel{~}{b}}{4}\right)\mathrm{ln}\frac{1}{\stackrel{~}{b}}+2\left(1\frac{\stackrel{~}{b}}{4}\right)\left(\mathrm{ln}(1+\sqrt{1\stackrel{~}{b}})\frac{\sqrt{1\stackrel{~}{b}}}{1+\sqrt{1\stackrel{~}{b}}}\right)\frac{\stackrel{~}{b}(1\stackrel{~}{b})}{2(1+\sqrt{1\stackrel{~}{b}})^2}\right]\right],$$ $`\stackrel{~}{b}=b^2/(4R_A^2)`$, the weak $`b`$-dependence of $`\epsilon _0`$ and approximately linear drop of $`\tau _L(b)`$ being derived for $`b\begin{array}{c}<\hfill \\ \hfill \end{array}R_A`$ analytically. and the standard Woods-Saxon nucleon densities). The edge effects near the surface of the nucleus, impact parameter dependence of nuclear parton structure functions (”nuclear shadowing”) , early transverse expansion of the system and other potentially important phenomena for peripheral ($`b2R_A`$) collisions are beyond our consideration here. ## 3 Impact parameter dependence of jet energy losses The intensity of final state rescattering and collisional and radiative energy losses of hard jet partons in dense QCD-matter, created in nuclear overlapping zone, are sensitive to their initial parameters (energy density, formation time) and space-time evolution . In order to analyze the impact parameter dependence of jet energy losses and jet quenching, we treat the medium as a boost-invariant longitudinally expanding quark-gluon fluid, and partons as being produced on a hyper-surface of equal proper times $`\tau =\sqrt{t^2z^2}`$ . We are expecting this is a adequate approximation for central and semi-central collisions for our semi-qualitative discourse. The approach relies on an accumulative energy losses, when both initial and final state gluon radiation is associated with each scattering in expanding medium together including the interference effect by the modified radiation spectrum as a function of decreasing temperature $`dE/dx(T)`$. Note that recently the radiative energy losses of a fast parton propagating through expanding (according to Bjorken’s model) QCD plasma have been explicitly evaluated in as $`dE/dx|_{expanding}=cdE/dx|_{T_L}`$ with numerical factor $`c2`$ $`(6)`$ for a parton created inside (outside) the medium, $`T_L`$ being the temperature at which the dense matter was left . The total energy losses in transverse direction experienced by a hard parton due to multiple scattering in matter are the result of averaging over the jet production vertex $`P_{AA}(𝐫,b)`$ (2), the transfer momentum squared $`t`$ in a single rescattering and space-time evolution of the medium: $$\mathrm{\Delta }E_T(b)=\underset{0}{\overset{2\pi }{}}𝑑\psi \underset{0}{\overset{r_{max}}{}}r𝑑r\frac{T_A(r_1)T_A(r_2)}{T_{AA}(b)}\underset{0}{\overset{2\pi }{}}\frac{d\phi }{2\pi }\underset{\tau _0}{\overset{\tau _L}{}}𝑑\tau \left(\frac{dE}{dx}^{rad}(\tau )+\underset{b}{}\sigma _{ab}(\tau )\rho _b(\tau )\nu (\tau )\right).$$ (15) Here $`\tau _0`$ and $`\tau _L`$ (7) are the proper time of the plasma formation and the time of jet escaping from the dense zone respectively; $`\rho _bT^3`$ is the density of plasma constituents of type $`b`$ at temperature $`T`$; $`\sigma _{ab}`$ is the integral cross section of scattering of the jet parton $`a`$ off the comoving constituent $`b`$ (with the same longitudinal rapidity $`y`$); $`\nu `$ and $`dE/dx^{rad}`$ are the thermal-averaged collisional energy loss of a jet parton due to single elastic scattering and radiative energy losses per unit distance respectively. If the mean free path of a hard parton is larger than the screening radius in the QCD medium, $`\lambda \mu _D^1`$, the successive scatterings can be treated as independent . The transverse distance between successive scatterings, $`\mathrm{\Delta }r_i=(\tau _{i+1}\tau _i)v_T=(\tau _{i+1}\tau _i)p_T/E`$, is determined in linear kinetic theory according to the probability density: $$\frac{dP}{d(\mathrm{\Delta }r_i)}=\lambda ^1(\tau _{i+1})\mathrm{exp}(\underset{0}{\overset{\mathrm{\Delta }r_i}{}}\lambda ^1(\tau _i+s)𝑑s),$$ (16) where the mean inverse free path is given by $`\lambda _a^1(\tau )=_b\sigma _{ab}(\tau )\rho _b(\tau )`$. The dominant contribution to the differential cross section $`d\sigma /dt`$ for scattering of a parton with energy $`E`$ off the ”thermal” partons with energy (or effective mass) $`m_03TE`$ at temperature $`T`$ can be written as $$\frac{d\sigma _{ab}}{dt}C_{ab}\frac{2\pi \alpha _s^2(t)}{t^2},$$ (17) where $`C_{ab}=9/4,1,4/9`$ for $`gg`$, $`gq`$ and $`qq`$ scatterings respectively, $$\alpha _s=\frac{12\pi }{(332N_f)\mathrm{ln}(t/\mathrm{\Lambda }_{QCD}^2)}$$ (18) is the QCD running coupling constant for $`N_f`$ active quark flavours, and $`\mathrm{\Lambda }_{QCD}`$ is the QCD scale parameter which is of the order of the critical temperature, $`\mathrm{\Lambda }_{QCD}T_c`$. The integrated parton scattering cross section, $$\sigma _{ab}=\underset{\mu _D^2\left(\tau \right)}{\overset{m_0\left(\tau \right)E/2}{}}𝑑t\frac{d\sigma _{ab}}{dt},$$ (19) is regularized by the Debye screening mass squared $`\mu _D^2`$. The collisional energy losses due to elastic scattering with high-momentum transfer have been originally estimated by Bjorken in , and recalculated later in taking also into account the loss with low-momentum transfer dominated by the interactions with plasma collective modes. Since latter process contributes to the total collisional energy losses without the large factor $`\mathrm{ln}(E/\mu _D)`$ in comparison with high-momentum scattering and it can be effectively ”absorbed” by the redefinition of minimal $`t\mu _D^2`$ under the numerical estimates, we shall concentrate on collisional energy losses with high-momentum transfer only <sup>2</sup><sup>2</sup>2Anyway, high- and low-momentum parts of collisional energy losses have the same dependence on distance traversed.. The thermal average of such loss can be written as $$\nu =\frac{t}{2m_0}=\frac{1}{2}\frac{1}{m_0}t\frac{1}{4T\sigma _{ab}}\underset{\mu _D^2}{\overset{3TE/2}{}}𝑑t\frac{d\sigma _{ab}}{dt}t.$$ (20) The value $`\nu `$ is independent of total distance traversed and determined by temperature, roughly $`\nu T`$. Then total collisional energy losses integrated over whole jet path are estimated as $`\mathrm{\Delta }E_{col}T_0^2\sqrt{\epsilon _0}`$, as it has been pointed out in . The $`\tau _L`$-dependence of $`\mathrm{\Delta }E_{col}`$ can be weaker than linear for expanding medium ($`\mathrm{\Delta }E_{col}\tau _L`$ for static matter). The energy spectrum of coherent medium-induced gluon radiation and the corresponding dominated part of radiative energy losses, $`dE/dx`$, were analyzed in by means of the Schrödinger-like equation whose ”potential” is determined by the single-scattering cross section of the hard parton in the medium. For the quark produced in the medium it gives <sup>3</sup><sup>3</sup>3The gluons with formation times $`\tau _f`$ exceeding the time $`\tau _L=L`$ that are formed outside the medium (the factorization medium-independent component) carry away a fraction of the initial parton energy proportional to $`\alpha _s(E)`$. This part of gluon radiation produces the standard jet energy profile which is identical to that of a jet produced in a hard process in the vacuum. Hereafter we shall concentrate on the medium-dependent effects and will not include the ”vacuum” part of the jet profile. $`{\displaystyle \frac{dE}{dx}}^{rad}={\displaystyle \frac{2\alpha _sC_R}{\pi \tau _L}}{\displaystyle \underset{\omega _{\mathrm{min}}}{\overset{E}{}}}𝑑\omega \left[1y+{\displaystyle \frac{y^2}{2}}\right]\mathrm{ln}\left|\mathrm{cos}(\omega _1\tau _1)\right|,`$ (21) $`\omega _1=\sqrt{i\left(1y+{\displaystyle \frac{C_R}{3}}y^2\right)\overline{\kappa }\mathrm{ln}{\displaystyle \frac{16}{\overline{\kappa }}}}\text{with}\overline{\kappa }={\displaystyle \frac{\mu _D^2\lambda _g}{\omega (1y)}}.`$ (22) Here $`\tau _1=\tau _L/(2\lambda _g)`$, and $`y=\omega /E`$ is the fraction of the hard parton energy carried by the radiated gluon, and $`C_R=4/3`$ is the quark colour factor. A similar expression for the gluon jet can be obtained by substituting $`C_R=3`$ and a proper change of the factor in the square bracket in (21), see . The integral (21) is carried out over all energies from $`\omega _{\mathrm{min}}=E_{LPM}=\mu _D^2\lambda _g`$ ($`\lambda _g`$ is the gluon mean free path), the minimal radiated gluon energy in the coherent LPM regime, up to initial jet energy $`E`$. The complex form of the expression (21) does not allow us in general case to extract the explicit form of $`\tau _L`$\- and $`T`$\- dependences of $`dE/dx^{rad}`$. In the limit of ”strong” LPM effect, $`\omega \mu _D^2\lambda _g`$, we have $`dE/dx^{rad}T^3`$ and $`dE/dx^{rad}\tau _L`$ with logarithmic accuracy. Then total radiative energy losses $`\mathrm{\Delta }E_{rad}=𝑑\tau 𝑑E/𝑑x^{rad}`$ are estimated as $`\mathrm{\Delta }E_{rad}T_0^3\epsilon _0^{3/4}`$ and $`\mathrm{\Delta }E_{rad}\tau _L^\beta `$, where $`\beta \begin{array}{c}<\hfill \\ \hfill \end{array}2`$ for expanding medium ($`\beta 2`$ in the case of static matter). In order to simplify numerical calculations (and not to introduce new parameters) we omit the transverse expansion and viscosity of the fluid using the well-known scaling Bjorken’s solution for temperature and density of QGP at $`T>T_c200`$ MeV: $$\epsilon (\tau )\tau ^{4/3}=\epsilon _0\tau _0^{4/3},T(\tau )\tau ^{1/3}=T_0\tau _0^{1/3},\rho (\tau )\tau =\rho _0\tau _0.$$ (23) Let us remark that the influence of the transverse flow, as well as of the mixed phase at $`T=T_c`$, on the intensity of jet rescattering (which is a strongly increasing function of $`T`$) seems to be inessential for high initial temperatures $`T_0T_c`$ . On the contrary, the presence of viscosity slows down the cooling rate, which leads to a jet parton spending more time in the hottest regions of the medium. As a result the rescattering intensity goes up, i.e., in fact an effective temperature of the medium gets lifted as compared with the perfect QGP case . We also do not take into account here the probability of jet rescattering in nuclear matter, because the intensity of this process and corresponding contribution to total energy losses are not significant due to much smaller energy density in a ”cold” nuclei. For certainty we used the initial conditions for the gluon-dominated plasma formation ($`N_f0`$, $`\rho _q1.95T^3`$) expected for central $`PbPb`$ collisions at LHC : $`\tau _00.1`$ fm/c, $`T_01`$ GeV <sup>4</sup><sup>4</sup>4These estimates are of course rather approximate and model-depending: the discount of higher order $`\alpha _s`$ terms, uncertainties of structure functions in the low-$`x`$ region, and nuclear shadowing can result in variations of the initial energy density .. Figure 5 represents the calculated $`\tau _L`$-dependence of coherent medium-induced radiative and collisional energy losses of a quark-initiated jet with initial energy $`E_T^q=100`$ GeV. We see that the $`\tau _L`$-dependence of radiative and collisional losses is very different: $`\mathrm{\Delta }E_{rad}(\tau _L)`$ grows somewhat stronger than linearly, meanwhile $`\mathrm{\Delta }E_{col}(\tau _L)`$ looks rather logarithmic. This results in the corresponding difference in the impact parameter dependence of radiative and collisional losses, the normalized profile of which are presented in figure 6. To make the plot more visual the energy losses $`\mathrm{\Delta }E_T(b)`$ are normalized to the corresponding average values at zero impact parameter, $`\mathrm{\Delta }E_{Trad}^q(b=0)45`$ GeV and $`\mathrm{\Delta }E_{Tcol}^q(b=0)5`$ GeV for the parameters used. For example, decreasing of impact parameter from $`b=0`$ to $`b=R_{Pb}`$ gives $`30\%`$ collisional and $`50\%`$ radiative losses reduction. We have also found that the the form of $`b`$-dependence of collisional losses is almost independent of scenarios of space-time evolution of QGP (perfect or viscous fluid), $`b`$-dependence of radiative losses being somewhat more sensitive to these effects. Note that the choice of the scale for a minimal jet energy $`E_T^q100`$ GeV corresponds to the threshold for ”true” QCD-jet recognition against the ”thermal” background jets (statistical fluctuations of the transverse energy flux) with reconstruction efficiency closed to $`1`$ in heavy ion collisions at LHC . We hope that the separation of collisional and radiative contribution to the total energy losses, when doing the experimental data analysis for jets with finite cone size, could be performed basing on essential difference of their angular distribution : the radiative losses are expected to dominate at small jet cone size $`\theta _0`$, while the relative contribution to collisional losses grows with increasing $`\theta _0`$. ## 4 Impact parameter dependence of dijet production rate In previous section we have analyzed the impact parameter dependence of jet energy losses, which can be directly observed in $`\gamma +jet`$ and $`Z(\mu ^+\mu ^{})+jet`$ production processes. Another observable effect is a suppression of high-$`p_T`$ jet pair yield (dijet quenching) due to final state rescattering and energy losses. In connection with this, we would like to estimate the impact parameter dependence of $`jet+jet`$ production rates in heavy ion collisions. The observed number of $`\{ij\}`$ type dijets with transverse momenta $`p_{T1},p_{T2}`$ produced in initial hard scattering processes in minimum bias $`AA`$ collisions is written as: $`{\displaystyle \frac{dN_{ij}^{dijet}}{dp_{T1}dp_{T2}}}`$ $`=`$ $`{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}d^2b{\displaystyle \frac{d^2\sigma _{jet}^0}{d^2b}}{\displaystyle \frac{dN_{ij}^{dijet}}{dp_{T1}dp_{T2}}}(b)/{\displaystyle \underset{0}{\overset{\mathrm{}}{}}}d^2b{\displaystyle \frac{d^2\sigma _{jet}^0}{d^2b}},`$ (24) $`{\displaystyle \frac{dN_{ij}^{dijet}}{dp_{T1}dp_{T2}}}(b)`$ $`=`$ $`{\displaystyle \underset{0}{\overset{2\pi }{}}}d\psi {\displaystyle \underset{0}{\overset{r_{max}}{}}}rdrT_A(r_1)T_A(r_2){\displaystyle \underset{0}{\overset{2\pi }{}}}{\displaystyle \frac{d\phi }{2\pi }}{\displaystyle }dp_T^2{\displaystyle \frac{d\sigma _{ij}}{dp_T^2}}\delta (p_{T1}p_T+`$ (25) $`\mathrm{\Delta }E_T^i(r,\psi ,\phi ,b))\delta (p_{T2}p_T+\mathrm{\Delta }E_T^j(r,\psi ,\pi \phi ,b)),`$ where parton differential cross section $`d\sigma _{ij}/dp_T^2`$ is calculated in the perturbative QCD: $$\frac{d\sigma _{ij}}{dp_T^2}=K𝑑x_1𝑑x_2𝑑\widehat{t}f_i(x_1,p_T^2)f_j(x_2,p_T^2)\frac{d\widehat{\sigma }_{ij}}{d\widehat{t}}\delta (p_T^2\frac{\widehat{t}\widehat{u}}{\widehat{s}}),$$ (26) $`d\widehat{\sigma _{ij}}/d\widehat{t}`$ expresses the differential cross-section for a parton-parton scattering as a function of the kinematical Mandelstam variables $`\widehat{s}`$, $`\widehat{t}`$ and $`\widehat{u}`$, $`f_{i,j}`$ are the structure functions, $`x`$ is the nucleon-momentum fraction carried by a parton, the correction factor $`K`$ takes into account higher order contributions. We have tested with the program of S.D.Ellis et al. that next-to-leading order (NLO) corrections are insignificant ($`K1`$) for jets with $`p_T50÷100`$ GeV/c and reasonable cone radius in the ($`y,\varphi `$)-plane $`R=0.3÷0.5`$ (see also ). Note also that the region of sufficiently hard jets, $`x_{1,2}\sqrt{\widehat{s}/s}\begin{array}{c}>\hfill \\ \hfill \end{array}0.2`$, almost does not affected by the initial state nuclear interactions like gluon depletion (”nuclear shadowing” of nucleon structure functions) . Anyway, the integrated above the threshold value $`p_T^{cut}`$ dijet rate, $$R_{AA}^{dijet}(p_{T1},p_{T2}>p_T^{cut})=\underset{p_T^{cut}}{}𝑑p_{T1}\underset{p_T^{cut}}{}𝑑p_{T2}\underset{i,j}{}(\frac{dN_{ij}^{dijet}}{dp_{T1}dp_{T2}})_{AA},$$ (27) in $`AA`$ relative to $`pp`$ collisions can be studied by introducing a reference process, unaffected by energy losses and with a production cross section proportional to the number of nucleon-nucleon collisions, such as Drell-Yan dimuons or (suitable for LHC ) Z$`(\mu ^+\mu ^{})`$ production, $$R_{AA}^{dijet}/R_{pp}^{dijet}=\left(\sigma _{AA}^{dijet}/\sigma _{pp}^{dijet}\right)/\left(\sigma _{AA}^{DY(Z)}/\sigma _{pp}^{DY(Z)}\right).$$ (28) The cross section $`d^2\sigma _{jet}^0/d^2b`$ for initially produced jets in $`AA`$ collisions at given $`b`$ can be written as : $$\frac{d^2\sigma _{jet}^0}{d^2b}(𝐛,\sqrt{s})=T_{AA}(𝐛)\sigma _{NN}^{jet}(\sqrt{s})\frac{d^2\sigma _{in}^{AA}}{d^2b}(𝐛,\sqrt{s}),$$ (29) where nucleon-nucleon collision cross section of the hard process $`\sigma _{NN}^{jet}`$ has been computed with PYTHIA model . The differential inelastic $`AA`$ cross section is calculated as: $$\frac{d^2\sigma _{in}^{AA}}{d^2b}(𝐛,\sqrt{s})=\left[1\left(1\frac{1}{A^2}T_{AA}(𝐛)\sigma _{NN}^{in}(\sqrt{s})\right)^{A^2}\right]$$ (30) with inelastic non-diffractive nucleon-nucleon cross section $`\sigma _{NN}^{in}`$ ($`60`$ mb for $`\sqrt{s}=5.5`$ TeV). Figure 7 shows dijet rates $`\sigma _{AA}^{in}R_{AA}^{dijet}L`$ ($`E_T^{jet}>p_T^{cut}=100`$ GeV, rapidity window $`|y^{jet}|<2.5`$) in different impact parameter bins for three cases: $`(i)`$ without energy losses, $`(ii)`$ with collisional losses only, $`(iii)`$ with collisional and radiative losses. The rates are normalized to the expected number of events produced in Pb-Pb collisions during two weeks ($`1.2\times 10^6`$ s) of LHC run time, assuming luminosity $`L=10^{27}cm^2s^1`$ . The total initial dijet rate with $`E_T^{jet}>100`$ GeV is estimated as $`1.1\times 10^7`$ events ($`gggg`$ $`60\%`$, $`qgqg`$ $`30\%`$, $`qq,ggqq`$ $`10\%`$). Since the dijet quenching is much stronger in central collisions than in peripheral one’s, the maximum and mean values of $`dN^{dijet}/db`$ distribution get shifted towards the larger $`b`$. The corresponding result for jets with non-zero cone size $`\theta _0`$ is expected to be somewhere between $`(iii)`$ ($`\theta _00`$) and $`(ii)`$ cases. The observation of a dramatic change in the $`b`$-dependence of dijet rates in heavy ion collisions as compared to what is expected from independent nucleon-nucleon interactions pattern, would indicate the existence of medium-induced parton rescattering. As we have mentioned above, the measurement of the centrality of events can be performed from total transverse energy deposition $`E_T^{tot}`$ in calorimeters, which strongly decreases from central to peripheral collisions , roughly as $`\overline{E_T^{tot}}(b)T_{AA}(b)`$. If jet energy losses $`\mathrm{\Delta }E_T^{jet}`$ (15) or dijet production rates $`R^{dijet}`$ (27,28) are measured in different bins of $`E_T^{tot}`$, then one can relate $`b`$\- and $`E_T^{tot}`$\- dependences of $`F=(\mathrm{\Delta }E_T^{jet},R^{dijet})`$ using $`E_T^{tot}b`$ correlation functions $`C_{AA}`$: $`F(E_T^{tot})={\displaystyle d^2bF(b)C_{AA}(E_T^{tot},b)},C_{AA}(E_T^{tot},b)={\displaystyle \frac{1}{\sqrt{2\pi }\sigma _{E_T}(b)}}\mathrm{exp}\left({\displaystyle \frac{\left(E_T^{tot}\overline{E_T^{tot}}(b)\right)^2}{2\sigma _{E_T}^2(b)}}\right),`$ (31) $`F(b)={\displaystyle 𝑑E_T^{tot}F(E_T^{tot})C_{AA}(b,E_T^{tot})},C_{AA}(b,E_T^{tot})={\displaystyle \frac{1}{\sqrt{2\pi }\sigma _b(E_T^{tot})}}\mathrm{exp}\left({\displaystyle \frac{\left(b\overline{b}(E_T^{tot})\right)^2}{2\sigma _b^2(E_T^{tot})}}\right).`$ (32) The estimated with the HIJING model accuracy of impact parameter determination $`\sigma _b(E_T^{tot})12`$ fm in $`AA`$ collisions at LHC seems to be enough to observe the above effects. ## 5 Conclusions To summarize, we have considered the impact parameter dependence of medium-induced radiative and collisional jet energy losses in dense QCD-matter, created in ultrarelativistic heavy ion collisions. We have found that this $`b`$-dependence is very different for each mechanism due to coherent effects (the dependence of radiative energy losses per unit distance $`dE/dx`$ of total distance traversed). As a consequence, the radiative losses are more sensitive to the impact parameter of nucleus-nucleus collision, which determines the effective volume of nuclear overlapping dense zone, and the space-time evolution of the medium. A possible way to directly observe the energy losses at different impact parameter (or total detected $`E_T`$ deposition) bins, involves tagging the hard jet opposite a particle that does not interact strongly, like in $`\gamma +jet`$ and $`Z(\mu ^+\mu ^{})+jet`$ production processes. Since initial energy density $`\epsilon _0`$ in dense zone depends on $`b`$ very slightly ($`\delta \epsilon _0\begin{array}{c}<\hfill \\ \hfill \end{array}10\%`$) up to $`bR_A`$, studying $`b`$-dependence appears to be advantageous than using of different ions at fixed impact parameter $`b0`$ (when $`\epsilon _0(b0)A^{2/3}`$). We hope that the separation of collisional and radiative contribution to total energy losses when doing the experimental data analysis for jets with finite cone size could be performed basing on essential difference in their angular distributions. Another process of interest is high-$`p_T`$ jet pair production. The expected statistics for dijet rates in heavy ion collisions at LHC will be large enough to study the impact parameter dependence. Since suppression of dijet yield (jet quenching) due to medium-induced energy losses should be much stronger in central collisions than in the peripheral one’s, the maximum and mean values of $`dN^{dijet}/db`$ distribution predicted to be shifted towards the larger $`b`$. Finally, the study of the impact parameter dependences in the hard jet production processes ($`jet+jet`$, $`\gamma +jet`$ and $`Z+jet`$ channels) is important for extracting information about the properties of super-dense QCD-matter to be created in heavy ion collisions at LHC. $`Acknowledgements`$. Discussions with Yu.L.Dokshitzer, L.I.Sarycheva and R.Vogt are gratefully acknowledged.
warning/0004/gr-qc0004067.html
ar5iv
text
# Untitled Document Compatibility of the Expansive Nondecelerative Universe Model with the Newton Gravitational Theory and the General Theory of Relativity Miroslav Sukenik<sup>a</sup>, Jozef Sima<sup>a</sup> and Julius Vanko<sup>b</sup> <sup>b</sup>Slovak Technical University, Dep. Inorg. Chem., Radlinskeho 9, 812 37 Bratislava, Slovakia <sup>a</sup>Comenius University, Dep. Nucl. Physics, Mlynska dolina F1, 842 48 Bratislava, Slovakia e-mail: sima@chelin.chtf.stuba.sk; vanko@fmph.uniba.sk Abstract. Applying the Vaidya metrics in the model of Expansive Nondecelerative Universe (ENU) leads to compatibility of the ENU model both with the classic Newton gravitational theory and the general theory of relativity in weak fields. Applying the Vaidya metrics to gravitational field has led us to relation $`\epsilon _g=\frac{R.c^4}{8\pi .G}=\frac{3m.c^2}{4\pi .a.r^2}`$(1) where $`\epsilon _g`$ is the density of the gravitational energy created by a body with the mass m in the distance r, R is the scalar curvature, a is the gauge factor reaching at the present $`a10^{26}m`$(2) The mean energy density of ENU, $`\epsilon _{ENU}`$ is expressed as $`\epsilon _{ENU}=\frac{3c^4}{8\pi .G.a^2}`$(3) When both densities are of the identical value $`\left|\epsilon _g\right|=\epsilon _{ENU}`$(4) then the following relations hold $`R=r_{ef}=(R_g.a)^{1/2}`$(5) in which $`r_{ef}`$ is the effective radius of a body with the mass m, $`R_g`$ is its gravitational diameter. Compton wave $`\lambda _C`$ can be expressed as $`\lambda _C=\frac{\text{h}\text{ }\text{ }}{m.c}`$(6) and then, based on identity of (5) and (6), relation expressing the lightest particle capable to have gravitational influence on its environment appears $`m_x^3=\frac{\text{h}\text{ }\text{ }^2}{2G.a}`$(7) The above mentioned relations are taken as starting points to manifest mutual compatibility of the Vaidya metrics incorporating ENU model and Newton gravitational theory. In the Newtonian approach, relation describing the density of gravitational energy is usually written as follows $`\epsilon _g=\frac{3G.m^2}{4\pi .r^4}`$(8) Providing that $`\lambda _C=r`$(9) the identity of (3) and (8) leads to (7) which means that the ENU model, Vaidya metrics and Newton gravitational theory are mutually compatible. Compatibility of the ENU model and the general theory of relativity will be exemplified on Hawking model of black hole evaporation . It was Hawking who, stemming from principles of quantum mechanics, thermodynamics and cosmology, evidenced that the output P of quantum evaporation of a black hole with the diamater $`R_{BH}`$ is $`P=\frac{\text{h}\text{ }\text{ }.c^2}{R_{BH}^2}`$ (10) and the energy E of a single quantum of the evaporation is ``` ``` $`E=\frac{\text{h}\text{ }\text{ }.c}{R_{BH}}`$ (11) The mass $`m_{lim}`$ of a limiting black hole is given by relation $`m_{lim}=\left(\frac{\text{h}\text{ }\text{ }.c^4.t}{4G^2}\right)^{1/3}`$ (12) where t is the cosmologic time. Comparing (11) and (12), for the energy of a limiting black hole evaporation quantum it follows $`E=m_x.c^2`$(13) where $`m_x=\left(\frac{\text{h}\text{ }\text{ }^2}{2G.a}\right)^{1/3}`$(14) i.e. identity of (7) and (13) is manifested. The mentioned identity can be taken as an example documenting compatibility of the ENU model, Vaidya metrics and the general theory of relativity. Conclusions The present results document the justifiability of Vaidya metrics incorporation into mathematical tools used in cosmological problems solving; The inclusion of Vaidya metrics in the ENU model allows to localize and quantify the energy of gravitational field outside a body; The present contribution suggests the existence of unity of the ENU model, Newton gravitational theory and general theory of relativity. References 1. P.C. Vaidya, Proc. Indian Acad. Sci., A33 (1951) 264 2. J. Sima, M. Sukenik, Preprint: gr-qc 9903090 (1999) 3. J. Sima, M. Sukenik, M. Sukenikova, Preprint: qr-qc 9910094 (1999) 4. S. Hawking, Sci. Amer., 236 (1980) 34
warning/0004/cond-mat0004064.html
ar5iv
text
# Experimental and numerical signatures of dynamical crossover in orientationally disordered crystals ## Abstract By means of NMR experiment and MD computer simulation we investigate the dynamical properties of a chloroadamantane orientationally disordered crystal. We find a plastic-plastic dynamical transition at $`T_x330`$ K in the pico-nanosecond regime. It is interpreted as the rotational analogue of the Goldstein crossing temperature between quasi-free diffusion and activated regime predicted in liquids. Below $`T_x`$, NMR experimental data are well described by a Frenkel model corresponding to a strongly anisotropic motion. At higher temperatures, a drastic deviation is observed toward quasi-isotropic rotational diffusion. Close to $`T_x`$, we observe that two-step relaxations emerge. An interpretation which is based on the present study of a specific heat anomaly detected by a recent calorimetric experiment is proposed. The elusive nature of the glass formation relates to the sharp rise of the transport coefficients (viscosity or relaxation times) in a narrow temperature range above the calorimetric transition temperature $`T_g`$ . Some of the recent active investigations focus on the existence of a dynamical crossover which could occur above $`T_g`$ in glass-forming liquids. The idealized version of the Mode Coupling Theory(MCT) predicts that a dynamical decoupling occurs at a critical temperature $`T_c>T_g`$ . It is however suspected that $`T_c`$ corresponds to or is close to the temperature $`T_x`$ at which the system also starts being sensitive to the energy landscape in its exploration of the configurational hyperspace . Motions in liquids would thus be dominate at low temperature by activated processes between potential energy minima while free diffusion occurs at high temperatures. Such dynamical crossover was already argued by Goldstein thirty years ago . In the last few years a significant effort has been undertaken to investigate experimentally and numerically the properties of glass-forming liquids in order to detect such dynamical changeover. However, most of the good glass-formers which allow easy investigations of the supercooled metastable state such as o-terphenyl are molecular liquids where contributions due to translational, orientational and internal degrees of freedom (TDOF, ODOF) are inextricably mixed. It makes the properties of the energy landscape topography, MCT predictions or detection of dynamical transitions extremely hard to investigate. A clear visualization of the different processes at the molecular level of simple model compounds is needed. The discovery that some orientationally disordered crystals exhibit the phenomenology of conventional glasses is therefore an exciting opportunity. Indeed, some molecular crystals present a *plastic phase* in which the molecular centers of mass show a perfect average crystalline translational order while the orientations are dynamically disordered. Some plastic crystals, i.e glassy crystals , can be undercooled and present many properties characteristic of conventional molecular-liquid glasses, both in the way the glass occurs (undercooling) and the thermodynamic signatures. Glassy crystals which are the very rotational analogue of liquid glass-formers offer an excellent opportunity to focus on the role of the ODOF. In , Barrat *et Al.* investigated the rotational dynamics of a colliding hard needle model on a rigid lattice and have shown the emergence of a dynamical decoupling. Substituted adamantanes plastic crystals such as the cyanoadamantane noted CNa in the following are good experimental candidates and some of them exhibit a glass transition signature. In , MD numerical simulations of a simple model inspired from the CNa molecular geometry allowed us to investigate both TDOF and ODOF properties. In this work a dynamical decoupling and two-step relaxations were found to be possible in a rotator phase. In the present letter, we address the possibility of finding such a behaviour in a real system. We have found chloroadamantane $`\mathrm{C}_{10}\mathrm{H}_{15}\mathrm{Cl}`$ (Cla) to be a very favorable and relevant system for this investigation. It shows a plastic phase over a wide interval of temperature $`[244442]`$ K (see Fig. 1c) and a structure isomorphous to CNa . The Cla molecule possesses a smaller substitute than does CNa. This gives rise to faster dynamics in the plastic phase where rotational motions are also suspected to change in nature as it is reported from earlier incoherent quasielastic neutron scattering experiments . Our results show for the first time, for a rotator phase, using experimental NMR and numerical MD simulations, the evidence of a dynamical crossover between quasi-free rotational diffusion and activated motion occurring in the pico-nanosecond regime. It is also shown that this accident can explain the existence of a calorimetric anomaly (see Fig. 1c) recently detected in Cla by Oguni *et Al.* . NMR experiments were performed on Chloroadamantane on Bruker spectrometers ASX100 and AMX 400. $`{}_{}{}^{1}\mathrm{H}`$ and $`{}_{}{}^{13}\mathrm{C}`$ linewidths and relaxation times were measured. The experiments were carried out on powder samples sealed in glass tubes for $`{}_{}{}^{1}\mathrm{H}`$ experiments or packed into zirconium rotors for $`{}_{}{}^{13}\mathrm{C}`$ MAS experiments. The spinning speed of the sample was 4 kHz. The sample temperature was regulated to within $`\pm 0.5`$ K by a Brucker BVT 2000 temperature controller. The spin lattice relaxation times were measured using the inversion recovery pulse sequence, the recovery of the magnetization was always found to be exponential within experimental accuracy. The $`{}_{}{}^{1}\mathrm{H}`$ spin lattice $`T_1`$ vs temperature is reported in Fig. 1a. The main feature of this curve is the occurrence of an abrupt change of the activation energy, from 1400 K to 3400 K, at a temperature close to $`T_x330`$ K which is far from both the melting temperature and the brittle to plastic transition. Slow diffusive motions of the molecules are generally observed in plastic phases. In the present study, their contribution to the spin-lattice relaxation time $`T_1`$ was found to be negligible in the temperature range investigated. This fact was particularly evidenced by measurements of the dipolar order $`T_{1D}`$ (see Fig. 1a). The diffusional correlation time ($`\tau _{diff}30000`$ s at room temperature) was found to be strongly temperature dependent with a very high activation energy $`14000`$ K. Self diffusion has thus to be discarded as the origin of the break observed in $`T_1`$. Previous neutron studies have suggested that two types of motion occur in the rotator phase of Cla: tumbling of the molecular $`C_3`$ axis among six fourfold crystallographic axes along the directions, and fast uniaxial rotation about the threefold axis. It was shown that the motion around the molecular axes is very fast ($`10^{12}`$ s at room temperature) with an activation energy of about 1240 K. Therefore, it takes place outside the time window where the dynamical crossover is detected at $`T_x`$. Dipolar tumbling motions are thus mainly involved in the evolution of the spin-lattice relaxation. Two separate temperature domains are clearly identified: below $`T_x`$, both NMR experimental proton and carbon data are well described by a Frenkel model assuming an Arrhenian behavior of tumbling and uniaxial rotations. Above $`T_x`$, it was not possible to describe the evolution of $`T_1`$ using the previous model. Hence, we expect that some hypotheses of this model in particular the finite number of equilibrium positions become irrelevant at high temperatures. Anisotropy in the rotational motion can be precisely seen from the spin lattice relaxation times of the three different protonated carbons. Effectively the main relaxation process of the methylene and methine carbons is the dipole-dipole interaction between the $`{}_{}{}^{13}\mathrm{C}`$ nuclei and their attached protons (see Fig. 1b). Relaxation times for carbon were measured as a function of temperature in the rotator phase. The ratio $`T_1(\mathrm{C4})/T_1(\mathrm{C2})`$ of the relaxational times for the methylene (C2, C4) is displayed in Fig. 1b. A strongly anisotropic motion is found below $`T_x`$ where the ratio of the carbon relaxational times is about 0.25. This latter value can be well reproduced by the previous Frenkel model. A drastic deviation from the anisotropic behaviour occurs at high temperatures where a ratio value of 0.7 is reached. Since free isotropic rotational diffusion processus would correspond to identical relaxational times for the methylene, this result reveals the influence of the residual crystalline field on the rotational dynamics. The experimental heat capacity obtained by Oguni et Al. , in the Cla plastic phase is reported in Fig. 1c. The evolution of $`C_p`$ shows an anomalous hump which was reported to resemble somewhat to a Schottky anomaly. Owing to the nature of the Cla molecule which is mainly rigid, conformational changes have to be discarded for the origin of the Schottky anomaly. Clearly, the hump covers the entire temperature range where the dynamical crossover is seen in our NMR investigation. Its origin will be discussed in the light of the MD simulation results. MD simulations have been performed at 29 different temperatures from $`T=`$ 220 to 500 K for a sample corresponding to the crystalline fcc rotator phase on a simple model of Cla (see Fig. 2). A complete description of the model is given in . Owing to the very long MD runs, we succeeded in investigating truly equilibrated states of the system down to 220 K. The reorientational motions can be described by a set of correlation functions which are defined as: $`C_l(t)=\frac{1}{N}_{i=1,N}P_l(\stackrel{}{\mu }_i(t).\stackrel{}{\mu }_i(0))`$ where $`P_l`$ is the $`l`$-order Legendre polynomial and $`\stackrel{}{\mu }_i`$ is the individual dipolar moment of the molecule $`i`$. The nature of dynamical changes and anisotropy can be checked by computing $`C_{l=1,2}`$ functions which can also be related to the information obtained from NMR relaxation measurements. Fig. 3 shows the $`C_2(t)`$ time correlation function for all investigated temperatures above 220 K. Clearly, a two step relaxational behaviour classically seen in supercooled molecular liquids is shown in our simulations. When lowering the temperature, an intermediate plateau region emerges which proves the presence of an orientational caging between neighboring molecules. This is the rotational analogue of the translational *cage effect* observed in liquids. This transient regime is followed by a slow $`\alpha `$ process which can be associated in our system, as checked for CNa , with large tumbling motion between fourfold crystallographic directions where dipoles are preferably localized (see Fig. 2). A consequence is that the tumbling of one molecule is allowed only if an orientational rearrangement of its local neighbors occurs. This *cooperative motion* is clearly displayed using projections of the individual dipolar moments on one crystallographic plane in Fig. 2. The $`\tau _1`$ and $`\tau _2`$ relaxational times are defined as the time it takes for their respective time correlation functions to decay $`e^1`$ of their initial values. They are displayed in Fig. 4a. Clearly, close to $`T_x330`$ K, both $`\tau _1`$ and $`\tau _2`$ relaxational times start diverging simultenously from the Arrhenian high temperature behaviour. Both high and low temperature apparent activation energies are found to be in good agreement with NMR experimental results. The nature of the rotational motions involved in this dynamical change is given by the ratio $`\tau _1/\tau _2`$ in Fig. 4b. This latter evolves from a value equal to 3 corresponding to free small-step rotational diffusion to a value of 1 classically associated to activated jump-like motion (see also Fig. 2. These results prove for the first time the existence of a plastic-plastic transition which can be interpreted as the rotational analogue of the Goldstein crossing temperature in liquids. It could also be associated with a different process of exploring the energy landscape topography above and below $`T_x`$. As it has been recently proposed, dynamics of the system can be viewed as hopping processes between separated inherent structures at sufficiently low temperatures. Specific heats have been directly extracted from the calculated enthalpy ($`C_p=dH/dT`$) and are shown in Fig. 4c. Clearly, a pronounced hump is seen which extends over the low temperature part of the rotator phase below 350 K. It proves that our MD simulation captures the main features which are responsible to the anomaly found in calorimetric experiments by Oguni et Al. . A good agreement with the experimental results is found for the area under the $`C_p`$ hump and its temperature extension while for the position of the maximum it is only fair. Obviously, the underlying absolute value of the specific heat is not reproduced in this computation owing to the simplicity of our model, where the internal vibrational contributions of the Cla molecule and the fast uniaxial rotations have not been taken into account. Thermodynamically, the anomalous hump requires an abrupt decrease and a sharp change in curvature of the entropy vs temperature evolution. Fig. 4c shows that as the temperature is lowered the $`C_p`$ hump begins to develop when dynamics change in nature. The strong localization of the dipole along the directions (see Fig. 2) which develops below $`T_x`$, is thus the prime suspect for the drastic change of the configurational entropy. In conclusion, convincing experimental and numerical evidence for a dynamical crossover ($`T_x`$) in the rotator phase of chloroadamantane have been obtained, very much in keeping with recent views on structural glass-formers. It is seen as a transition from quasi-free rotational diffusion of the molecular dipoles to activated geared tumblings at low temperatures. This rotational localization is shown numerically to give rise to a specific heat hump as was recently measured . The possible existence of such a calorimetric signature of a dynamical accident in liquids is highly debated and was unsuccessfully searched until now . Certainly, the underlying lattice and specificity of the orientational disorder contribute to make dynamical changes in the system much sharper than in liquids. We now feel confident that orientationally disordered crystals provide a valuable analog and alternative to the conventional liquid glass-formers. Results of the present study offer new interesting possibilities of testing the different theoretical approaches of the glass formation. The authors are indebted to Professor K. Ngai for stimulating discussions. This work was supported by the INTERREG II program (Nord Pas de Calais/Kent). The authors wish to acknowledge the use of the facilities of the IDRIS where some of the simulations were carried out.
warning/0004/astro-ph0004275.html
ar5iv
text
# Turmoil on the accretion disk of GRO J1655-40 ## 1 Introduction The black-hole candidate and soft X-ray transient GRO J1655$``$40 (X-ray Nova Sco 1994) was discovered on 1994 July 27 by the Burst and Transient Source Experiment (BATSE) onboard the Compton Gamma Ray Observatory (Zhang et al. 1994, Harmon et al. 1995). Soon thereafter the optical counterpart was found (Bailyn et al. 1995a). It has since shown irregular outburst activity (e.g. Zhang et al. 1997). After the 1995 July/August hard X-ray outburst the source settled back to quiescence. However, on 1996 April 25 it became active again (Remillard et al. 1996, Levine et al. 1996, Orosz et al. 1997) for a period which lasted $``$16 months (see e.g. Remillard et al. 1999). GRO J1655$``$40 has an orbital period of 2.62 days, has dynamically been shown to contain a black hole with a mass of $``$7 M, and is viewed at an inclination of $``$70 (Bailyn et al. 1995b, Orosz & Bailyn 1997, van der Hooft et al. 1998, Shahbaz et al. 1999). The object has become notorious for being one of the systems showing relativistic radio jets, during its first outburst in 1994 (Tingay et al. 1995; Hjellming & Rupen 1995). Deep short dips in the X-ray light curves of GRO J1655$``$40 were discovered during its 1996/1997 outburst These dips appeared periodically between orbital phases $``$0.7 and $``$0.85 and were shown to be due to a medium absorbing the radiation coming from the inner parts of the accretion disk (Kuulkers et al. 1998a). Such “orbital” dips are known to occur also in various other low-mass and high-mass X-ray binaries (see Parmar & White 1988; Marshall et al. 1993; Saraswat et al. 1996; and references therein). They probe the interaction between the stream from the companion and the accretion disk, as well as the X-ray emitting region itself (see e.g. Kuulkers et al. 1998a; Tomsick et al. 1998). We report here on observations made with the Wide Field Cameras (WFC) onboard the Beppo Satellite per Astronomia X (BeppoSAX) observatory, obtained during various stages of the last outburst in 1996/1997 of GRO J1655$``$40. Due to the relatively long orbital period of GRO J1655$``$40, the WFC is ideally suited for monitoring such a system at all orbital phases almost continuously, which is not possible or undertaken with other X-ray instruments carrying out short programs of typically less than a day or with all-sky monitors viewing sources typically up to ten times a day for a short period of order minutes. We summarize all dip events seen during the 1996/1997 outburst with the WFC and other instruments and discuss their origin. ## 2 Observations and Analysis The Wide Field Camera (WFC) instrument (Jager et al. 1997) consists of two coded aperture cameras that point in opposite directions and perpendicular to the Narrow-Field Instruments on the same BeppoSAX satellite (Boella et al. 1997). The field of view of each camera is 40 by 40 square degrees full-width to zero response with an angular resolution about 5 arcminute in each direction. The detectors are Xenon-filled multi-wire proportional counters with band passes of 2 to 28 keV. The sensitivity in the 2–10 keV band is about 10 mCrab in 10<sup>4</sup> s. The imaging capability and sensitivity allow an accurate monitoring of complex sky regions, like the Galactic bulge. The WFCs are carrying out a program of monitoring observations of the field around the Galactic Center. The purpose is to detect X-ray transient activity, particularly from low-mass X-ray binaries (LMXBs) whose Galactic population exhibits a strong concentration in this field, and to monitor the behaviour of persistently bright X-ray sources (see e.g. Heise 1998). This program consists of campaigns during the spring and autumn of each year. Each campaign lasts about two months and typically comprises weekly observations. The favorable position of GRO J1655$``$40 ($`l=344\stackrel{}{.}98`$, $`b=+2\stackrel{}{.}46`$ or $`\mathrm{R}.\mathrm{A}.=16^h54^m00^s,\mathrm{Dec}.=39\mathrm{°}50\mathrm{}45\mathrm{}`$, J2000.0) within the vicinity of the Galactic Center implies therefore a relatively large coverage of the source during its 1996/1997 outburst. In this paper we report on part of the X-ray monitoring program during 1996 August-October and 1997 March-April and August, as well as serendipitous observations close to the Galactic Center (see Table 1), yielding a total of 1.37 Msec of good data. These observations were either performed with WFC unit 1 or 2. For the light curves we used as intensity measure the count rate per cm<sup>2</sup> in the 2–8 keV band. All measurements are corrected for dead time. The All-Sky Monitor (ASM; Levine et al. 1996) onboard the Rossi X-ray Timing Explorer (RXTE) scans the X-ray sky in series of 90 sec dwells in three energy bands between 2 and 12 keV. A given source is observed typically 5 to 10 times a day. For our purpose we used the ASM quick-look results in the 2–12 keV band as provided by the RXTE/ASM team at the Massachusetts Institute of Technology (MIT). Numerous pointed target-of-opportunity observations were obtained during the 1996/1997 outburst of GRO J1655$``$40 with the Proportional Counter Array (PCA: 2–60 keV; Bradt et al. 1993) onboard RXTE (Sobczak et al. 1999; Remillard et al. 1999). Dips in the light curves were seen in observations taken on 1996 June 20 (Hynes et al. 2000, in preparation), 1997 January 5 (Remillard et al. 1999) and 1997 February 26 (Kuulkers et al. 1998a,b). We used the data collected with a time resolution of 0.125 sec in the total 2–60 keV energy band from the so-called ‘Standard 1’ mode. The High Resolution Imager (HRI: 0.1–2.4 keV; e.g. Zombeck et al. 1995) onboard the Röntgen Satellite (ROSAT) observed GRO J1655$``$40 various times in the periods 1996 August 28 to September 24 and 1997 March 10 to 13. During the second epoch deep short drops in intensity were found on March 12, similar to those seen with the RXTE/PCA. For the light curve we used all counts registered by the HRI; we did not correct for background, vignetting and dead-time, since the source is bright and since we are mainly interested in the dip light curves and times of dip occurrence. The Advanced Satellite for Cosmology and Astrophysics (ASCA) satellite (Tanaka et al. 1994) is equipped with two Solid State Imaging Spectrometers (SIS; 0.4–10 keV) and two Gas Imaging Spectrometers (GIS; 0.7–10 keV). For our purposes we only used the data taken by the GIS instruments since they provide a higher time resolution compared to the SIS instruments. ASCA observed GRO J1655$``$40 during its 1996/1997 outburst on 1997 February 26–28. The GIS instruments were set to PH mode with timing resolution of 62.5 ms in high bit rate and 0.5 s in medium bit rate. Standard data screening was employed<sup>1</sup><sup>1</sup>1See the ASCA Data Reduction Guide, Version 2.0 (http://heasarc.gsfc.nasa.gov/docs/asca/abc/abc.html).. Data taken at a geomagnetic cut-off rigidity lower than 4 GeV, at an elevation angle less than $`5^{}`$ from the Earth limb and during passage through the South Atlantic Anomaly were rejected. After filtering, the total net exposure time of each GIS was 33.7 ks. We extracted the GIS light curves from a circular region of radius $`6^{}`$ centered at the position of the source. ## 3 Results ### 3.1 The quest for X-ray dips In the upper panel of Fig. 1 we show the outburst light curve of GRO J1655$``$40 as obtained with the ASM. The outburst seems to consist of two stages separated by JD 2450450 (1997 January 1). The first stage shows strong flaring behaviour, whereas the second stage shows a smoother light curve. In the lower panel of Fig. 1 we show the observations as obtained with the WFC. Note that the last set of WFC observations were taken just before the source returned to quiescence (see also Sobczak et al. 1999, Remillard et al. 1999). The deep drops clearly seen in the ASM light curve are due to absorption dips, previously identified by Kuulkers et al. (1998a). In the WFC light curve such dipping behaviour is clearly present at the end of the second period of WFC observations. We constructed WFC light curves with a time resolution of 5 sec, and searched for clear drops in intensity. On several occasions we found clear dipping behaviour in the X-ray light curves; they all appeared between orbital phases 0.68 and 0.89. During the observations on 1996 August 22 and 1997 April 14 various dips appeared; their detailed light curves are given in Figs. 2 (left panel) and 3 (right panel). On other occasions only single dips were found (see Table 2). It is clear that the dips are very deep, i.e. down to $``$1% of the out-of-dip intensity, as was already noticed for other dips from RXTE/PCA and ASM observations discussed by Kuulkers et al. (1998a). Their durations range from $``$15 sec to $``$3.5 min. The profiles of the dips are reminiscent of those of the dips seen previously with the RXTE PCA (Kuulkers et al. 1998a). Particularly interesting is the light curve of 1997 April 14 (Fig. 3, right panel), during which continuing dip activity occurred for a total time span of $``$9.2 hr, i.e. between orbital phases 0.68–0.79 of binary cycle<sup>2</sup><sup>2</sup>2Binary cycle is defined as the number of binary orbits since JD 2449838.4198 (van der Hooft et al. 1998). 272. In Fig. 4 we show our collection of X-ray light curves from other instruments (a.–e.: RXTE/PCA, f.: ROSAT/HRI) during which clear dips were seen. Again, the dips are short, i.e. they have durations of up to a minute, and deep. Some of the light curves (e.g. Fig. 4c.) show already a slight depression in the intensity just before and after the dips. The main differences in the dip profiles can be discerned by comparing Figs. 4d. and e. Apart from a difference in the out-of-dip intensity, the dips on 1996 June 20 (Fig. 4d) are more gradual and somewhat less deep (down to $``$10% of the out-of-dip intensity), than those that occurred on 1997 January 5 (Fig. 4a–c) and February 26 (Fig. 4e). We note that ASCA observed GRO J1655$``$40 from 1997 February 26 00:46:49 to February 28 16:50:49, i.e. around the time of the RXTE/PCA observations on 1997 February 26. However, the X-ray coverage was rather sparse between orbital phases 0.76–0.94, and no X-ray dips could be found at these or other phases. We determined all times of the dips found in the observations described in this paper; the time and orbital phase ranges of their occurrences are displayed in Table 2 and indicated in the upper panel of Fig. 1. The total orbital phase range between which dips occur is 0.68–0.92. In Fig. 5 we plot the histogram of the orbital phase occurrences of all the dips. The occurrences cluster near $`\varphi _{\mathrm{orb}}0.8`$ and the distribution is consistent with being Gaussian. It may be interesting to note that all dips were only seen during the first halfs of the two stages of the outburst, i.e. dips occurred during JD 2450205–331, i.e. 1996 May 1 to August 17, and JD 2450453–567, i.e. 1997 January 4 to April 28 (Table 2, Fig. 1). ### 3.2 From binary orbit to binary orbit Due to the long continuous coverage of the WFC it is now possible for the first time to follow the behaviour from binary orbit to the next binary orbit. If one looks at the light curves exactly one or several binary orbit(s) earlier and/or later than the binary orbit which did display dips, no dips are found. This is illustrated in Figs. 2 and 3. In Fig. 2 we show the WFC light curves from orbital phases 0.65–0.95 taken one binary orbit apart. Clearly, the dips observed in binary cycle 182 (left panel) just after orbital phase 0.79 are not there in binary cycle 183 (right panel) at the same orbital phase. In Fig. 3 we show light curves 6 binary orbits apart from each other. The strong dipping behaviour in binary cycle 272 (right panel) is clearly not repeated at exactly the same phases in binary cycle 266 (left panel). The dipping behaviour seen in cycle 272 is stronger, i.e. much more dips occur, than in cycle 182, as well as compared to other binary orbits covered by the WFC. We attribute this to the transient nature of the dips, since we do not think that most or all of the dips are missed by earth occultations or SAA passages. To illustrate the changes per binary orbit further we constructed a light curve in the 2–20 keV band (Fig. 6, upper panel) and hardness curve given by the ratio of 5–20 keV and 2–5 keV (Fig. 6, lower panel) of the first nine consecutive days of observations in 1996. This is a unique data set from a continuous monitoring campaign of the Galactic Center (Cornelisse et al. 2000, in preparation). For clarity we have indicated the phase intervals of observed dipping behaviour (dashed lines) and the time when the companion star is closest to us, i.e. phase zero (dotted lines). Overall, both the intensity and hardness increased during the nine-day period. No clear repeatable features per orbit are seen, apart from a depression in the intensity near orbital phases 0.6–0.9 and 2.6–2.9. Moreover, the dip features which occur in the beginning of the observation (see also Fig. 2), are not repeated in subsequent orbits, although some might have been missed due to earth occultations or SAA passages (see also above). Note that the depressions in the light curves near orbital phases 0.6–0.9 and 2.6–2.9 are accompanied by a softening in the hardness values. On the other hand, no clear hardness changes are discerned during the depressions in the light curves near orbital phases 3.3–3.4 and 3.9–4.0. ## 4 Discussion ### 4.1 Orbital phase dependence Combining all times of X-ray dip occurrences during the 1996/1997 outburst of GRO J1655$``$40 we find that the dips occur between orbital phases 0.68 and 0.92, i.e. extending the dip range with respect to that quoted by Kuulkers et al. (1998a). The occurrence of the observed dips concentrate near orbital phases 0.75–0.80. Similar deep short dips have now been seen in other X-ray transient systems: 4U 1630$``$47 (Tomsick et al. 1998; Kuulkers et al. 1998a) and Cir X-1 (Shirey et al. 1999). As noted by Kuulkers et al. (1998a) the duration of the dips is also of the same order as those seen in Cyg X-1 and Her X-1 (e.g., Kitamoto et al. 1984, Leahy 1997), but is shorter than those typically seen in LMXB dip sources (see e.g. Parmar & White 1988). For the first time, we were able to obtain a nearly-continuous view of the system for several binary orbits during different parts of the outburst. On one occasion we found long irregular dipping behaviour for a period of $``$9.2 hr, which spans about $``$0.15 in orbital phase. This duty cycle is much larger than quoted by Tomsick et al. (1998) for GRO J1655$``$40; it is similar to that seen in low-mass X-ray binary dip sources (‘dippers’; e.g. Parmar & White 1988; White et al. 1995). Do the dips recur every binary orbit? At certain times they do, as was shown by Kuulkers et al. (1998b) to be the case during the early part of the 1996/1997 outburst using the RXTE/ASM light curves. Our collection of dip times suggests the same during other parts of the outburst (see Table 2). If one compares in detail the light curves from binary orbit to binary orbit, it is clear that the dips do not persist at exactly the same phase. Moreover, if the strong dipping behaviour seen in binary cycle 272 were present every orbit, we would certainly have seen this in the WFC light curves. In contrast most of the WFC light curves do not show evidence for any strong dipping behaviour. Such a transient nature of the dip activity has been noted before as well in the other LMXB dip sources, see e.g. Parmar et al. (1986) for the case of EXO 0748$``$676. We note that it looks as if dipping only occurred during the first halfs of the two stages of the outburst, i.e. they appear during the rise and plateau phases of the outburst. ### 4.2 Stream-disk interaction The orbital phase dependence of the dip occurrences suggests the dip mechanism to be fixed in the binary orbit. For illustrative purposes (Fig. 7) we show views of the system at the extremes of the dip phase range. The most straightforward site is the interaction region between the mass transfer stream from the companion star and the accretion disk (e.g. Frank et al. 1987). However, it is not the bulge on the accretion disk itself which obscures the X-ray emitting region. This would result in a rather long X-ray dip as seen in e.g. accretion disk corona sources (see e.g. White et al. 1995); moreover, the inclination at which we view GRO J1655$``$40 is too low for such an effect to be seen (see Frank et al. 1987). Rather the short duration of the dips points to an absorbing medium which is filamentary in nature and probably situated above and/or below the impact region (Kuulkers et al. 1998b). A similar filamentary nature has been proposed for the medium in Cyg X-1 that causes the second peak in the distribution of (short) dips with orbital phase at $`\varphi _{\mathrm{orb}}`$$``$0.6 (where phase zero corresponds to superior conjunction of the black hole). This peak has been attributed also to a mass transfer stream from the O-supergiant interacting with the accretion disk (Bałucińska-Church et al. 2000). Note that the main peak occurs at $`\varphi _{\mathrm{orb}}`$$``$0.95, which is due to absorption in the wind of the supergiant. The dynamics of mass transfer streams from the companion up to the place of impact onto the accretion disk was first investigated by Lubow & Shu (1975, 1976). More recently it has become possible to study in more detail the impact region itself and what happens with the matter after it reaches this region (e.g. Armitage & Livio 1998, and references therein). The hydrodynamic simulations by Armitage & Livio (1998) show that the fate of the stream after it impacts the disk depends on the efficiency of cooling in the shock-heated gas created by the impact. If the cooling is efficient, the upper and/or lower parts of the stream are able to freely overflow the disk, unless the disk rim is too thick. This ballistic stream will reimpact the disk near to its closest approach to the compact object (see also Frank et al. 1987, Lubow 1989). Observationally this may lead to rather broad dips occurring near orbital phase 0.6 (Lubow 1989, see also Armitage & Livio 1996). If cooling is inefficient than the stream ‘splashes’ onto the disk rim, i.e. the material overflowing the disk is a coherent stream but rather erratic. Moreover, more material is able to reach greater heights from the disk, which leads to substantially higher absorption columns with respect to the former case, for lines-of-sight well away from the disk plane. At such lines-of-sight one expects a plethora of dips to be observed from orbital phases $``$0.7–1.0, with a peak near orbital phase 0.8 (Armitage & Livio 1996, see also Frank et al. 1987). Moreover, in the latter case one expects a bulge extending along the disk rim as well (Armitage & Livio 1998). The dips observed during the 1996/1997 outburst of GRO J1655$``$40 all share the characteristics of the case of non-efficient cooling: they are short, show absorption columns up to $``$10<sup>24</sup> atoms cm<sup>-2</sup> (Kuulkers et al. 1998a) and the distribution of dip occurrence times extends from binary orbital phases $``$0.7–0.9, with a peak near binary orbital 0.8. Apart from the ‘splashing’ material causing the deep short dips, the presence of an extended bulge along the disk rim is supported by the observed decrease in optical polarisation near orbital phase $``$0.7–0.8 (Gliozzi et al. 1998). The observations were obtained just after the peak of the second part of the outburst during 1997 July 2–9. The polarisation region must be rather extended and is located primarily around the inner disk regions. A thickening of the disk rim or the remnant stream passing under and over the disk then obscures smoothly this polarisation region (Gliozzi et al. 1998). ASCA observed a $``$4 hr drop in intensity by $``$75% in the light curve of GRO J1655$``$40 on 1994 August 23 (Ueda et al. 1998). These observations were obtained just after the peak of the first hard X-ray and radio outburst (Harmon et al. 1995; Tavani et al. 1996). Using the ephemeris by van der Hooft et al. (1998), the ASCA dip occurred between orbital phases $``$0.53–0.61, i.e. outside the phase range we see for the short and deep dips during the 1996/1997 outburst<sup>3</sup><sup>3</sup>3We note that an eclipse-like feature in the optical light curves was found on 1994, August 17 by Bailyn et al. (1995a), i.e. close in time to the ASCA dip; a possibly similar optical feature occurred one binary orbit earlier (Bianchini et al. 1997). The eclipse-like light curve of Bailyn et al. (1995a) lasted from orbital phase $``$0.0–0.1. Bianchini et al. (1997) suggested this to be due to a bright spot region, or an extended optically thick disk rim shielding part of the disk.. No deep short dips have been reported for the outbursts which occurred in 1994/1995, however, the soft X-ray ($`<`$10 keV) coverage during these outbursts was very sparse. The long ASCA dip is in agreement with what is expected when the stream encounters a cold accretion disk rim, so that part of the stream can continue to flow above and under the accretion disk (see above). According to Ueda et al. (1998), the material causing the dip by absorption must consist of slightly ionized (warm) and non-ionized (cold) material, which is also in line with the fact that the stream will re-impact the disk sufficiently close to the region where ionisation becomes important (Frank et al. 1987). We suggest that the difference between the observed X-ray dips during the 1994 and 1996/1997 outbursts may be due to the efficiency in irradiating the outer disk regions. In 1994 the disk rim may have had a shape such that it does not see the X-ray emitting regions (i.e. self-screening by the disc of its outer regions; see Dubus et al. 1999), possibly due to a large enhancement of mass transfer from the secondary, and the gas at the stream impact region could cool efficiently. During the 1996/1997 outburst the outer disk regions were less affected by the mass transfer stream from the secondary, which were therefore more easily prone to X-ray irradiation by the X-ray emitting region, keeping the outer disk hot. ###### Acknowledgements. EK thanks Jerry Orosz and Frank Verbunt for comments on an earlier draft of this paper and Keith Horne for supplying the source code of ‘The CV Eclipse Movie (not showing at a theatre near you)’. AKH is supported by a Hong Kong Oxford Scholarship. The BeppoSAX satellite is a joint Italian and Dutch programme.
warning/0004/math0004114.html
ar5iv
text
# Classification of semisimple Hopf algebras of dimension 16. ## 1. Introduction. Recently various classification results were obtained for finite-dimensional semisimple Hopf algebras over an algebraically closed field of characteristic $`0`$. The smallest dimension, for which the question was still open, was $`16`$. In this paper we completely classify all nontrivial (i.e. non-commutative and non-cocommutative) Hopf algebras of dimension $`16`$. Moreover, we consider all possible structures of Grothendieck rings $`K_0\left(H\right)`$ for semisimple non-commutative Hopf algebras of dimension $`16`$. Let $`H`$ be a non-commutative semisimple Hopf algebra of dimension $`16`$ over an algebraically closed field $`k`$ of characteristic $`0`$. Then irreducible representations of $`H`$ of degree $`1`$ are exactly the grouplike elements of $`H^{}`$. Let $`𝐆\left(H^{}\right)`$ denote the group of grouplikes of $`H^{}`$, then $`k𝐆\left(H^{}\right)`$ is a subHopfalgebra of $`H^{}`$ and thus, by Nichols-Zoeller Theorem , $`\left|𝐆\left(H^{}\right)\right|=dimk𝐆\left(H^{}\right)`$ divides $`dimH^{}=dimH=16`$. Therefore by the Artin-Wedderburn Theorem, as an algebra $`H`$ is isomorphic to either (1.1) $`k^{\left(8\right)}M_2\left(k\right)M_2\left(k\right)\text{ or}`$ (1.2) $`k^{\left(4\right)}M_2\left(k\right)M_2\left(k\right)M_2\left(k\right)`$ $`dimZ\left(H\right)`$ equals the number of summands in the Artin-Wedderburn decomposition of $`H`$, thus in the case $`\left(\text{1.1}\right)`$ $`dimZ\left(H\right)=10`$ and $`\left|𝐆\left(H^{}\right)\right|=8`$ and in the case $`\left(\text{1.2}\right)`$ $`dimZ\left(H\right)=7`$ and $`\left|𝐆\left(H^{}\right)\right|=4`$. Our first result, which will be proved in the beginning of Section 3 , is the following: ###### Theorem 1.1. Let $`H`$ be a semisimple Hopf algebra of dimension $`p^n`$ over an algebraically closed field $`k`$ of characteristic $`0`$. If $`HkC_{p^n}`$ then $`𝐆\left(H\right)`$ is not cyclic. Our main result will be proved in Section 9: ###### Theorem 1.2. Let $`k`$ be an algebraically closed field of characteristic $`0`$. Then there are exactly $`16`$ nonisomorphic nontrivial semisimple Hopf algebras of dimension $`16`$. 1. $`𝐆\left(H\right)`$ is abelian of order $`8`$ if and only if $`𝐆\left(H^{}\right)`$ is abelian of order $`8`$. There are $`11`$ Hopf algebras with such a group of grouplikes. 2. If $`H`$ has a nonabelian group of grouplikes then $`𝐆\left(H\right)=D_8`$ and $`𝐆\left(H^{}\right)=C_2\times C_2`$. There are $`2`$ Hopf algebras with a nonabelian group of grouplikes. 3. There are $`3`$ Hopf algebras with $`𝐆\left(H\right)=C_2\times C_2`$. Two of them are dual to the Hopf algebras with a nonabelian group of grouplikes and one of them is selfdual. ###### Remark 1.1. A part of Theorem 1.2, saying that if $`H`$ has a nonabelian group of grouplikes then $`𝐆\left(H\right)=D_8`$ and $`𝐆\left(H^{}\right)=C_2\times C_2`$ can also be obtained as a corollary to a theorem of Natale , and Proposition 3.1. This theorem states that if $`𝐆\left(H\right)`$ is nonabelian then $`H^{}`$ has $`4`$ central grouplikes. One method of constructing a new Hopf algebra from a known one $`H`$ is to twist the comultiplication of $`H`$ by a $`2`$-pseudo-cocycle $`\mathrm{\Omega }HH`$ (or a $`2`$-cocycle $`JHH`$). The new Hopf algebra is denoted $`H_\mathrm{\Omega }`$ (or $`H_J`$). The next theorem summarizes the results of Sections 5 and 6: ###### Theorem 1.3. Let $`H`$ be a semisimple Hopf algebra of dimension $`16`$ over an algebraically closed field $`k`$ of characteristic $`0`$. Then there are exactly $`7`$ possible structures of the Grothendieck ring $`K_0\left(H\right)`$. Moreover 1. $`𝐆\left(H^{}\right)`$ is abelian if and only if the Grothendieck ring of $`H`$ is commutative. Then 1. If $`\left|𝐆\left(H^{}\right)\right|=8`$, as algebras $`K_0\left(H\right)_{}kk^{\left(10\right)}`$. 2. If $`\left|𝐆\left(H^{}\right)\right|=4`$, as algebras $`K_0\left(H\right)_{}kk^{\left(7\right)}`$. 3. $`K_0\left(H\right)=K_0\left(kG\right)`$, where $`G`$ is one of the $`9`$ nonabelian groups of order $`16`$ (although only $`6`$ of those $`K_0`$-rings are distinct). 4. $`H`$ is a twisting with a $`2`$-pseudo-cocycle of some group algebra. 2. If $`K_0\left(H\right)`$ is not commutative then 1. As algebras $`K_0\left(H\right)_{}kk^{\left(6\right)}M_2\left(k\right)`$. 2. $`H`$ is not a twisting of a group algebra. 3. There is only one possible structure of the $`K_0`$-ring. 4. All Hopf algebras with non-commutative $`K_0`$-ring are twistings of each other. ###### Remark 1.2. By Theorem 1.2 there are only $`2`$ Hopf algebras with nonabelian $`𝐆\left(H^{}\right)`$. We summarize the distinct non-commutative, non-cocommutative semisimple Hopf algebras of dimension $`16`$ in the following table. We try to distinguish nonisomorphic examples of Hopf algebras using the groups $`𝐆\left(H\right)`$ and $`𝐆\left(H^{}\right)`$ and the Grothendieck rings $`K_0\left(H\right)`$ (defined in Section 2). Here we consider twistings of group algebras $`kG`$, where $`G`$ is a nonabelian group of order $`16`$. There are exactly nine such groups, described in (see Section 4). The twistings appearing here are explained in Section 7. The coproduct $`\mathrm{\#}^\alpha `$ is explained in Section 8. $`H_8`$ denotes the unique nontrivial semisimple Hopf algebra of dimension $`8`$ (see and ). ###### Remark 1.3. $`H_{C:\sigma _1}`$ is not triangular for the following reasons. If it were triangular then by \[4, Theorem 2.1\] it would be equal to a twisting with a $`2`$-cocycle of a group algebra $`kG`$. Then by \[24, Theorem 4.1\] $`K_0\left(H_{C:\sigma _1}\right)=K_0\left(kG\right)`$ and therefore $`H_{C:\sigma _1}`$ would be a twisting of $`kD_{16}`$ or of $`kQ_{16}`$. But by \[16, Theorem 4.1\], $`kQ_{16}`$ doesn’t have nontrivial cocycle twistings and $`H_{C:1}\left(kD_{16}\right)_J`$ is the only cocycle twisting of $`kD_{16}`$. ###### Remark 1.4. The following Hopf algebras are selfdual: $`H_{d:1,1}H_8kC_2`$ (since $`H_8`$ is selfdual), $`H_{c:\sigma _0}`$ (since comparing $`K_0`$-rings we see that $`H_{c:\sigma _0}A_3^+\left(A_3^+\right)^{}`$, described in and ), $`H_{d:1,1}k\left(D_8\times C_2\right)_J`$ and $`H_{C:\sigma _1}`$ (since there is no other choice for the dual). ## 2. Preliminaries. First we will need the following definition, which was introduced in . ###### Definition 2.1. Let $`K_0\left(H\right)^+`$ denote the abelian semigroup of all equivalence classes of representations of $`H`$ with the addition given by a direct sum. Then its enveloping group $`K_0\left(H\right)`$ has the structure of an ordered ring with involution and is called the Grothendieck ring. In the structure of $`K_0\left(H\right)`$ was described for comodules; it was then translated into the language of modules in . The multiplication in this ring is defined as follows: let $`\left[\pi _1\right]`$ and $`\left[\pi _2\right]`$ denote the classes of representations equivalent to $`\pi _1`$ and $`\pi _2`$, then $`\left[\pi _1\right]\left[\pi _2\right]`$ is the class of the representation $`\left(\pi _1\pi _2\right)\mathrm{\Delta }`$; the unit of this ring is the class $`\left[\epsilon \right]`$ and $`\left[\pi \right]^{}`$ is the equivalence class of the dual representation $`{}_{}{}^{t}(\pi S)`$ defined by $`{}_{}{}^{t}(\pi S\left(h\right))\left(f\right),v=f,(\pi S\left(h\right))\left(v\right)`$. The equivalence classes of irreducible representations of $`H`$ form a basis of $`K_0\left(H\right)`$ and are called basic elements. If $`\left[\pi _1\right],\mathrm{},\left[\pi _d\right]`$ are the basic elements then $`\left[\rho \right]=_{i=1}^d\mathrm{deg}\pi _i\left[\pi _i\right]`$ is called the marked element. For basic elements $`x`$ and $`y`$ we write $$xy=\underset{z\text{ - basic}}{}m(z,xy)z$$ where $`m(z,xy)`$ are non-negative integers. Then the following properties are true (see and ): (2.1) $`m(z,xy)`$ $`=`$ $`m(x^{},yz^{})`$ (2.2) $`m(1,xy^{})`$ $`=`$ $`\delta _{x,y}`$ (2.3) $`{\displaystyle m(z,xy)\mathrm{deg}\left(z\right)}`$ $`=`$ $`\mathrm{deg}\left(xy\right)`$ For simplicity of notation we will write $`\pi `$ instead of $`\left[\pi \right]`$ for elements of $`K_0\left(H\right)`$. We will denote the degree $`2`$ irreducible representations of $`H`$ by $`\pi _i`$ and the degree $`1`$ irreducible representations of $`H`$ (i.e. elements of $`𝐆\left(H^{}\right)`$ or multiplicative characters of $`H`$) by $`\chi _i`$. We denote the generators of $`𝐆\left(H^{}\right)`$ by $`\chi `$, $`\phi `$ and $`\psi `$. If $`H=kG`$ then $`𝐆\left(\left(kG\right)^{}\right)`$ is the group of multiplicative characters of $`G`$. The following proposition can be also obtained as a corollary to \[16, Proposition 2.4\]: ###### Proposition 2.2. Let $`H`$ be a nontrivial semisimple Hopf algebra of dimension $`16`$. Assume that there exists an element $`\chi 𝐆\left(H^{}\right)Z\left(H^{}\right)`$ of order $`2`$ such that $`\chi \pi =\pi `$ for every $`2`$-dimensional representation $`\pi `$ of $`H`$. Then $`H^{}`$ has a group algebra of dimension $`8`$ as a quotient. ###### Proof. Write $`G=𝐆\left(H^{}\right)`$. Dualizing formulas (1.1) and (1.2) we get that as coalgebras $`H^{}`$ $`=`$ $`kGE_1E_2\text{if }\left|𝐆\left(H^{}\right)\right|=8\text{ or}`$ $`H^{}`$ $`=`$ $`kGE_1E_2E_3\text{if }\left|𝐆\left(H^{}\right)\right|=4`$ where $`E_i`$ are simple subcoalgebras of dimension $`4`$ and $`\chi E_i=E_i`$. $`\left(\chi 1\right)H^{}`$ is a normal Hopf ideal of $`H^{}`$. Then $`L=H^{}/\left(\chi 1\right)H^{}`$ is a Hopf algebra of dimension $`8`$. Consider the projection $`p:H^{}L`$. Since $`\chi E_i=E_i`$, $`p\left(E_i\right)=E_i/\left(\chi 1\right)E_i`$. Therefore $`L`$ $`=`$ $`k\left(G/<\chi >\right)p\left(E_1\right)p\left(E_2\right)\text{if }\left|𝐆\left(H^{}\right)\right|=8\text{ or}`$ $`L`$ $`=`$ $`k\left(G/<\chi >\right)p\left(E_1\right)p\left(E_2\right)p\left(E_3\right)\text{if }\left|𝐆\left(H^{}\right)\right|=4`$ $`p\left(E_i\right)`$ are cosemisimple coalgebras of dimension $`2`$, therefore each of them is spanned by two grouplikes. Thus $`L`$ is spanned by $`8`$ grouplikes and $`L`$ is a group algebra. ∎ We will also need the notion of a twisting of a Hopf algebra (see , , ): ###### Definition 2.3. The twisting $`H_\mathrm{\Omega }`$ of a Hopf algebra $`H`$ is a Hopf algebra with the same algebra structure and counit and with comultiplication and antipode given by $`\mathrm{\Delta }_\mathrm{\Omega }\left(h\right)=\mathrm{\Omega }\mathrm{\Delta }\left(h\right)\mathrm{\Omega }^1`$ $`S_\mathrm{\Omega }\left(h\right)=uS\left(h\right)u^1`$ for all $`hH`$, where $`\mathrm{\Omega }HH`$ and $`uH`$ are invertible elements. The new comultiplication $`\mathrm{\Delta }_\mathrm{\Omega }`$ is coassociative if and only if $`\mathrm{\Omega }`$ is a $`2`$-pseudo-cocycle, that is $`_2\left(\mathrm{\Omega }\right)`$ lies in the centralizer of $`\left(\mathrm{\Delta }id\right)\mathrm{\Delta }\left(H\right)`$ in $`HHH`$, where $$_2\left(\mathrm{\Omega }\right)=\left(id\mathrm{\Delta }\right)\left(\mathrm{\Omega }^1\right)\left(1\mathrm{\Omega }^1\right)\left(\mathrm{\Omega }1\right)\left(\mathrm{\Delta }id\right)\left(\mathrm{\Omega }\right).$$ $`\mathrm{\Omega }`$ is called a $`2`$-cocycle if $`_2\left(\mathrm{\Omega }\right)=111`$ and in this case we will denote it by $`J`$. ###### Remark 2.1. By \[24, Theorem 4.1\] $`K_0\left(H\right)K_0\left(H_\mathrm{\Omega }\right)`$ as ordered rings with marked elements, and thus $`𝐆\left(H^{}\right)`$ $`𝐆\left(\left(H_\mathrm{\Omega }\right)^{}\right)`$. ## 3. Hopf algebras of dimension $`16`$ with a commutative subHopfalgebra of dimension $`8`$. We apply the methods used by Masuoka in , and . Let $`H`$ be a nontrivial semisimple Hopf algebra of dimension $`16`$ with a subHopfalgebra $`K=\left(kG\right)^{}`$ of dimension $`8`$. Since $`K`$ is a subHopfalgebra of index $`2`$, by \[10, Proposition 2\] or \[20, Theorem 2.1.1\] $`K`$ is normal in $`H`$ and thus we have an exact sequence of Hopf algebras (3.1) $$K\stackrel{i}{}H\stackrel{\pi }{}F$$ where $`F=ktkC_2`$ and $`K=\left(kG\right)^{}`$, which is cleft by or . Such a sequence is called an extension of $`F`$ by $`K`$ and was first studied by Kac in . The construction of extensions from cohomological data was done in and . $`K`$ is commutative and $`F`$ is cocommutative and thus $`(F,K)`$ form an abelian matched pair of Hopf algebras and $`(G,t)`$ form an abelian matched pair of groups (see , , , \[12, Section 1\]). Therefore $`H`$ becomes a bicrossed product $`K\mathrm{\#}_\sigma ^\theta F`$ with an action $`:FKK`$, a coaction $`\rho :FFK`$, a cocycle $`\sigma :FFK`$ and a dual cocycle $`\theta :FKK`$. $`G`$ is a normal subgroup of the group $`G\times t`$, arising from a matched pair $`(G,t)`$, since $`G`$ has index $`2`$ in $`G\times t`$. Thus $`\rho `$ is trivial and the action by $`t`$ is a Hopf algebra automorphism of $`K`$ (see \[12, Section 1\]). $``$ is a nontrivial action on $`K`$, since otherwise $`HK^t\left[C_2\right]`$ as an algebra, and thus $`H`$ is commutative. Let $`v=\sigma (t,t)K.`$ Then by the properties of the cocycle $`v`$ is a unit and (3.2) $$tv=v$$ Multiplication in $`H`$ gives us (3.3) $`\overline{t}^2`$ $`=`$ $`v`$ (3.4) $`\overline{t}c`$ $`=`$ $`\left(tc\right)\overline{t}`$ where $`\overline{t}=1\mathrm{\#}t`$ and $`cK`$. Moreover, if a unit $`vK`$ satisfies $`\left(\text{3.2}\right)`$, $`\left(\text{3.3}\right)`$ and $`\left(\text{3.4}\right)`$, we can define a cocycle $`\sigma `$ by $`\sigma (1,1)=\sigma (1,t)=\sigma (t,1)=1`$ and $`\sigma (t,t)=v.`$ We proceed by considering the possible $`G`$, namely $`C_8`$, $`C_4\times C_2`$, $`C_2\times C_2\times C_2`$, $`D_8`$ and $`Q_8`$. Theorem 1.1 says that the first case cannot appear. ###### Theorem 1.1. Let’s prove the statement by induction on $`n`$. When $`n=2`$, by \[13, Theorem 2\] $`H`$ is a group algebra and if $`HkC_{p^2}`$ then $`Hk\left(C_p\times C_p\right)`$ and $`𝐆\left(H\right)C_p\times C_p`$. Now assume the statement is true for $`n=m`$. Consider $`H`$ of dimension $`p^{m+1}`$. $`dim\left(H^{}\right)=p^{m+1}`$ and thus, by \[13, Theorem 1\], there exists a central grouplike of order $`p`$ in $`H^{}`$ and therefore $`H^{}`$ contains a normal subHopfalgebra $`KkC_p`$. Thus we get a short exact sequence of Hopf algebras (3.5) $$K\stackrel{𝑖}{}H^{}\stackrel{𝜋}{}F$$ where $`F=H^{}/K^+H^{}`$. Dualizing $`\left(\text{3.5}\right)`$ we get another short exact sequence of Hopf algebras: (3.6) $$F^{}\stackrel{\pi ^{}}{}H\stackrel{i^{}}{}K^{}$$ where $`K^{}KkC_p`$ and $`dimF^{}=dimF=p^m`$. Thus we get $`𝐆\left(F^{}\right)𝐆\left(H\right)`$ and $`𝐆\left(F^{}\right)`$ is not cyclic unless $`F^{}kC_{p^m}`$. In the first case we are done since it implies that $`𝐆\left(H\right)`$ is not cyclic. In the second case, since $`K`$ is normal in $`H^{}`$, $`H^{}`$ is isomorphic as an algebra to a twisted group ring $`K^t\left[F\right]`$ where $`FF^{}kC_{p^m}`$. It is easy to show that, since$`F`$ is a group algebra of a cyclic group, $`K^t\left[F\right]`$ is commutative. Thus $`H`$ is cocommutative and the only possible $`H`$ with a cyclic group of grouplikes is $`kC_{p^{m+1}}`$. ∎ ### 3.1. Case of $`𝐆\left(H\right)=C_4\times C_2`$. We will show that there are at most $`7`$ possible Hopf algebras of this kind. Let $`H`$ be a nontrivial semisimple Hopf algebra of dimension $`16`$ with a subHopfalgebra $`K=k\left(C_4\times C_2\right)^{}k\left(C_4\times C_2\right)`$. Then $`𝐆\left(H\right)=GC_4\times C_2`$. Let $`G=x\times y`$ with $`\left|x\right|=4`$ and $`\left|y\right|=2`$. Then the dual basis of $`KK^{}`$ is given by $$e_{pq}=1/8\left(1+i^px+i^{2p}x^2+i^{3p}x^3\right)\left(1+\left(1\right)^qy\right),\text{ }p=0,1,2,3;q=0,1$$ Then $`\mathrm{\Delta }_H\left(e_{pq}\right)`$ $`=`$ $`\mathrm{\Delta }_K\left(e_{pq}\right)={\displaystyle \underset{\begin{array}{c}p_1+p_2pmod4\\ q_1+q_2qmod2\end{array}}{}}e_{p_1q_1}e_{p_2q_2}`$ $`\mathrm{\Delta }_H\left(\overline{t}\right)`$ $`=`$ $`\theta \left(t\right)\overline{t}\overline{t}`$ where $`\overline{t}=1\mathrm{\#}t.`$ Dualizing $`\left(\text{3.1}\right)`$ we get another extension $$F^{}\stackrel{\pi ^{}}{}H^{}\stackrel{i^{}}{}K^{}$$ and as in \[11, 2.4\], \[12, 2.11\] or \[15, 2.1\], since $`k`$ is algebraically closed, there exist units $`\overline{x}`$ and $`\overline{y}H^{}`$, such that $`\overline{x}^4=\overline{y}^2=1_H^{}`$, $`e_{pq},\overline{x}^i\overline{y}^j=\delta _{ip}\delta _{jq}`$ and $`\alpha =\overline{y}^1\overline{x}^1\overline{y}\overline{x}F^{}=k\{e_0,e_1\}`$, where $`\left\{e_r\right\}`$ is a dual basis of $`\left\{t^r\right\}`$, $`r=0,1`$. $`\epsilon \left(\alpha \right)=\epsilon \left(\overline{y}^1\overline{x}^1\overline{y}\overline{x}\right)=1`$ and therefore $`\alpha =e_0+\xi e_1`$. The right action $`\rho ^{}:F^{}K^{}F^{}`$ is trivial, thus $`F^{}`$ lies in the center of $`H^{}`$. $$\overline{x}=\overline{y}^2\overline{x}=\overline{y}\overline{x}\overline{y}\alpha =\overline{x}\overline{y}\alpha \overline{y}\alpha =\overline{x}\overline{y}^2\alpha ^2=\overline{x}\alpha ^2$$ Thus $`\alpha ^2=1`$ and therefore $`\xi =\pm 1`$. $$\mathrm{\Delta }_H\left(\overline{t}\right),\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r=\overline{t},\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r=\delta _{kr}\overline{t},\overline{x}^{i+p}\overline{y}^{j+q}\alpha ^{jp}e_k=\xi ^{jp}\delta _{k1}\delta _{r1}$$ On the other hand $$\mathrm{\Delta }_H\left(\overline{t}\right),\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r=\theta \left(t\right)\overline{t}\overline{t},\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r=\theta \left(t\right),\overline{x}^i\overline{y}^j\overline{x}^p\overline{y}^q\delta _{k1}\delta _{r1}$$ Therefore $$\theta \left(t\right)=\underset{ijpq}{}\xi ^{jp}e_{ij}e_{pq}$$ and since $`H`$ should be non-cocommutative, $`\theta (t,t)`$ is nontrivial, and thus $`\xi =1`$ and $$\theta \left(t\right)=\underset{ijpq}{}\left(1\right)^{jp}e_{ij}e_{pq}=\frac{1}{2}\left(\left(1+y\right)1+\left(1y\right)x^2\right)$$ Write $`v=\sigma (t,t)=c_{i,j}e_{i,j}`$ then $`c_{0,0}=\epsilon \left(v\right)=1`$ and $`c_{i,j}0,`$ since $`v`$ is a unit, and $$\mathrm{\Delta }_H\left(\overline{t}^2\right)=\mathrm{\Delta }_H\left(v\right)=\mathrm{\Delta }_K\left(c_{i,j}e_{i,j}\right)=c_{i+p,j+q}e_{i,j}e_{p,q}$$ On the other hand, if we write $$te_{p,q}=e_{\alpha _1(p,q),\alpha _2(p,q)}$$ $`\mathrm{\Delta }\left(\overline{t}\right)\mathrm{\Delta }\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \underset{ijpq}{}}\left(1\right)^{jp}e_{ij}\overline{t}e_{pq}\overline{t}{\displaystyle \underset{ijpq}{}}\left(1\right)^{jp}e_{ij}\overline{t}e_{pq}\overline{t}`$ $`=`$ $`{\displaystyle \underset{ijpq}{}}\left(1\right)^{jp}e_{ij}e_{pq}{\displaystyle \underset{ijpq}{}}\left(1\right)^{jp}e_{\alpha _1(i,j),\alpha _2(i,j)}\overline{t}^2e_{\alpha _1(p,q),\alpha _2(p,q)}\overline{t}^2`$ $`=`$ $`{\displaystyle \underset{ijpq}{}}\left(1\right)^{jp}e_{ij}e_{pq}{\displaystyle \underset{ijpq}{}}\left(1\right)^{\alpha _2(i,j)\alpha _1(p,q)}e_{ij}\overline{t}^2e_{pq}\overline{t}^2`$ $`=`$ $`{\displaystyle \underset{ijpq}{}}\left(1\right)^{jp+\alpha _2(i,j)\alpha _1(p,q)}c_{ij}c_{pq}e_{ij}e_{pq}`$ Thus for $`H`$ to be a bialgebra we should have (3.7) $$c_{i+p,j+q}=\left(1\right)^{jp+\alpha _2(i,j)\alpha _1(p,q)}c_{i,j}c_{p,q}$$ Action by $`t`$ is a Hopf algebra map and therefore $`tG=G`$ and $`f_t:GG`$ defined by $`f_t\left(g\right)=tg`$ is a group automorphism of order $`2`$. There are three possibilities for such an automorphism; we consider them below: #### Case a) The action is given by $`t`$ $``$ $`x=xy`$ $`t`$ $``$ $`y=y`$ Then $`te_{i,j}=e_{i+2j,j}`$. Write $`v=\sigma (t,t)=c_{i,j}e_{i,j}`$ . By $`\left(\text{3.2}\right)`$ and $`\left(\text{3.7}\right)`$ (3.8) $`c_{i,j}`$ $`=`$ $`c_{i+2j,j}`$ (3.9) $`c_{i+p,j+q}`$ $`=`$ $`c_{i,j}c_{p,q}`$ Conditions $`\left(\text{3.8}\right)`$ and $`\left(\text{3.9}\right)`$ imply that $`c_{1,0}=\left(1\right)^k`$ and $`c_{0,1}=\left(1\right)^l`$ for $`k,l=0,1`$ and $`\sigma (t,t)={\displaystyle \underset{p,q}{}}\left(1\right)^{kp+lq}e_{p,q}={\displaystyle \left(1\right)^{kp}e_{p,q}\left(1\right)^{lq}e_{p,q}}=x^{2k}y^lk,l=0,1`$ For $`k,l=0,1`$ let $`H_{k,l}`$ be the Hopf algebras with the structures described above with cocycles $`\sigma _{k,l}(t,t)=x^{2k}y^l`$. Define $`f:H_{k,l}`$ $``$ $`H_{k+1,l+1}\text{by}`$ $`f\left(e_{r,s}\right)`$ $`=`$ $`e_{r,s}`$ $`f\left(\overline{t}\right)`$ $`=`$ $`x\overline{t}`$ and extend it multiplicatively to $`f\left(e_{r,s}\overline{t}\right)`$. Then $`f`$ is a trivial group homomorphism on $`𝐆\left(H_{k,l}\right)`$ and $`f\left(\overline{t}\right)f\left(\overline{t}\right)`$ $`=`$ $`x\overline{t}x\overline{t}=x^2y\overline{t}^2=x^2yx^{2\left(k+1\right)}y^{\left(l+1\right)}=x^{2k}y^l=f\left(\overline{t}^2\right)`$ $`f\left(\overline{t}x\right)`$ $`=`$ $`f\left(xy\overline{t}\right)=xyx\overline{t}=x\overline{t}x=f\left(\overline{t}\right)f\left(x\right)`$ $`\left(ff\right)\mathrm{\Delta }\left(\overline{t}\right)`$ $`=`$ $`\left(ff\right)\left(\theta \left(t\right)\overline{t}\overline{t}\right)=\theta \left(t\right)\left(f\left(\overline{t}\right)f\left(\overline{t}\right)\right)=\theta \left(t\right)\left(x\overline{t}x\overline{t}\right)`$ $`=`$ $`\left(xx\right)\theta \left(t\right)\left(\overline{t}\overline{t}\right)=\mathrm{\Delta }\left(x\right)\mathrm{\Delta }\left(\overline{t}\right)=\mathrm{\Delta }\left(x\overline{t}\right)=\mathrm{\Delta }\left(f\left(\overline{t}\right)\right)`$ and such an $`f`$ is a Hopf algebra isomorphism between $`H_{k,l}`$ and $`H_{k+1,l+1}`$. Thus there are at most two nonisomorphic Hopf algebras of this type: 1. $`H_{a:1}=H_{0,0}`$ with the trivial cocycle and $`𝐆\left(H_{a:1}^{}\right)=\chi \times \phi C_4\times C_2`$, where $`\chi (x)=i`$, $`\chi (y)=\chi (t)=1`$, $`\phi \left(x\right)=\phi \left(y\right)=1`$, $`\phi \left(t\right)=1`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi ^2=\pi \pi =1+\chi ^2+\phi +\chi ^2\phi `$. 2. $`H_{a:y}=H_{0,1}`$ with the cocycle defined by $`\sigma (t,t)=y`$ and $`𝐆\left(H_{a:y}^{}\right)=\chi \times \phi C_4\times C_2`$, where $`\chi (x)=i`$, $`\chi (y)=\chi (t)=1`$, $`\phi \left(x\right)=\phi \left(y\right)=1`$, $`\phi \left(t\right)=1`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi ^2=\pi \pi =1+\chi ^2+\phi +\chi ^2\phi `$. #### Case b) The action is given by $`t`$ $``$ $`x=x^1`$ $`t`$ $``$ $`y=y`$ Then $`te_{i,j}=e_{i,j}`$. Write $`v=\sigma (t,t)=c_{i,j}e_{i,j}`$ . By $`\left(\text{3.2}\right)`$ and $`\left(\text{3.7}\right)`$ (3.10) $`c_{i,j}`$ $`=`$ $`c_{i,j}`$ (3.11) $`c_{i+p,j+q}`$ $`=`$ $`c_{i,j}c_{p,q}`$ Conditions $`\left(\text{3.10}\right)`$ and $`\left(\text{3.11}\right)`$ imply that $`c_{1,0}=\left(1\right)^k`$ and $`c_{0,1}=\left(1\right)^l`$ for $`k,l=0,1`$ and $`\sigma (t,t)={\displaystyle \underset{p,q}{}}\left(1\right)^{kp+lq}e_{p,q}={\displaystyle \left(1\right)^{kp}e_{p,q}\left(1\right)^{lq}e_{p,q}}=x^{2k}y^lk,l=0,1`$ For $`k,l=0,1`$ let $`H_{k,l}`$ be the Hopf algebras with the structures described above with cocycles $`\sigma _{k,l}(t,t)=x^{2k}y^l`$. Define $`f:H_{0,0}`$ $``$ $`H_{1,0}\text{by}`$ $`f\left(e_{r,s}\right)`$ $`=`$ $`e_{r,r+s}`$ $`f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(\left(1+i\right)1+\left(1i\right)x^2\right)\overline{t}={\displaystyle \underset{p=0}{\overset{3}{}}}{\displaystyle \underset{q=0}{\overset{1}{}}}i^{p^2}e_{p,q}\overline{t}`$ and extend it multiplicatively to $`f\left(e_{r,s}\overline{t}\right)`$. Then $`f_{𝐆\left(H_{0,0}\right)}`$ is a group isomorphism $`𝐆\left(H_{0,0}\right)𝐆\left(H_{1,0}\right)`$ with $`f\left(x\right)=x`$, $`f\left(y\right)=x^2y`$ and $`f\left(\overline{t}\right)f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(\left(1+i\right)1+\left(1i\right)x^2\right)\overline{t}\left(\left(1+i\right)1+\left(1i\right)x^2\right)\overline{t}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(\left(1+i\right)1+\left(1i\right)x^2\right)^2\overline{t}^2={\displaystyle \frac{1}{4}}\left(2i1+4x^22i1\right)\sigma _{1,0}(t,t)`$ $`=`$ $`x^2x^2=1=f\left(\overline{t}^2\right)`$ $`f\left(\overline{t}x\right)`$ $`=`$ $`f\left(x^1\overline{t}\right)={\displaystyle \frac{1}{2}}x^1\left(\left(1+i\right)1+\left(1i\right)x^2\right)\overline{t}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(\left(1+i\right)1+\left(1i\right)x^2\right)\overline{t}x=f\left(\overline{t}\right)f\left(x\right)`$ $`\left(ff\right)\mathrm{\Delta }\left(\overline{t}\right)`$ $`=`$ $`\left(ff\right)\left(\theta \left(t\right)\overline{t}\overline{t}\right)=\left(ff\right)\left({\displaystyle \left(1\right)^{q_1p_2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}\right)`$ $`=`$ $`{\displaystyle \left(1\right)^{q_1p_2}e_{p_1,q_1+p_1}i^{p^2}e_{p,q}\overline{t}e_{p_2,q_2+p_2}i^{p^2}e_{p,q}\overline{t}}`$ $`=`$ $`{\displaystyle \left(1\right)^{\left(q_1+p_1\right)p_2}i^{p_1^2+p_2^2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}`$ $`=`$ $`{\displaystyle \left(1\right)^{q_1p_2}i^{p_1^2+2p_1p_2+p_2^2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}`$ $`=`$ $`{\displaystyle \left(1\right)^{q_1p_2}i^{\left(p_1+p_2\right)^2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}`$ $`=`$ $`\left({\displaystyle i^{\left(p_1+p_2\right)^2}e_{p_1,q_1}e_{p_2,q_2}}\right)\left({\displaystyle \left(1\right)^{s_1r_2}e_{r_1,s_1}\overline{t}e_{r_2,s_2}\overline{t}}\right)`$ $`=`$ $`\mathrm{\Delta }\left({\displaystyle i^{p^2}e_{p,q}}\right)\mathrm{\Delta }\left(\overline{t}\right)=\mathrm{\Delta }\left({\displaystyle i^{p^2}e_{p,q}\overline{t}}\right)=\mathrm{\Delta }\left(f\left(\overline{t}\right)\right)`$ and such an $`f`$ is a Hopf algebra isomorphism between $`H_{0,0}`$ and $`H_{1,0}`$. There are at most three nonisomorphic Hopf algebras of this kind: 1. $`H_{b:1}=H_{0,0}`$ with the trivial cocycle and $`𝐆\left(H_{b:1}^{}\right)=\chi \times \phi \times \psi C_2\times C_2\times C_2`$, where $`\chi (x)=1`$, $`\chi (y)=\chi (t)=1`$, $`\phi \left(x\right)=\phi \left(y\right)=1`$, $`\phi \left(t\right)=1`$, $`\psi \left(y\right)=1`$, $`\psi \left(x\right)=\psi \left(t\right)=1`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill 0& \hfill \left(1\right)^p\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi ^2=\pi \pi =1+\chi +\phi +\chi \phi `$. 2. $`H_{b:y}=H_{0,1}`$ with the cocycle defined by $`\sigma (t,t)=y`$ and $`𝐆\left(H_{b:y}^{}\right)=\chi \times \phi C_4\times C_2`$, where $`\chi (x)=1`$, $`\chi (y)=1`$, $`\chi (t)=i`$, $`\phi \left(x\right)=1`$, $`\phi \left(y\right)=\phi \left(t\right)=1`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill 0& \hfill \left(1\right)^p\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi ^2=\pi \pi =1+\chi ^2+\phi +\chi ^2\phi `$. 3. $`H_{b:x^2y}=H_{1,1}`$ with the cocycle defined by $`\sigma (t,t)=x^2y`$ and $`𝐆\left(H_{b:x^2y}^{}\right)=\chi \times \phi C_4\times C_2`$, where $`\chi (x)=1`$, $`\chi (y)=1`$, $`\chi (t)=i`$, $`\phi \left(x\right)=1`$, $`\phi \left(y\right)=\phi \left(t\right)=1`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill 0& \hfill \left(1\right)^p\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi ^2=\pi \pi =1+\chi ^2+\phi +\chi ^2\phi `$. #### Case c) The action is given by $`t`$ $``$ $`x=x`$ $`t`$ $``$ $`y=x^2y`$ Then $`te_{i,j}=e_{i,j+i}`$. Write $`v=\sigma (t,t)=c_{i,j}e_{i,j}`$ . By $`\left(\text{3.2}\right)`$ and $`\left(\text{3.7}\right)`$ (3.12) $`c_{i,j}`$ $`=`$ $`c_{i,i+j}`$ (3.13) $`c_{i+p,j+q}`$ $`=`$ $`\left(1\right)^{ip}c_{i,j}c_{p,q}`$ Conditions $`\left(\text{3.12}\right)`$ and $`\left(\text{3.13}\right)`$ imply that $`c_{1,0}^4=c_{0,1}=1`$, $`c_{2,0}=c_{1,0}^2`$ Thus $`c_{1,0}=i^k`$ for $`k=0,1,2,3`$ and $$\sigma _k(t,t)=\underset{p,q}{}\left(1\right)^{\frac{p\left(p1\right)}{2}}i^{kp}e_{p,q}=x^{1k}\left(\frac{1+i}{2}1+\frac{1i}{2}x^2\right)$$ For $`k=0,1,2,3`$ let $`H_k`$ be the Hopf algebras with the structures described above with cocycles $`\sigma _k`$. Define $`f:H_{k+2}`$ $``$ $`H_k\text{by}`$ $`f\left(e_{p,q}\right)`$ $`=`$ $`e_{p,q}`$ $`f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \underset{p,q}{}}\left(1\right)^{p+q}e_{p,q}\overline{t}=y\overline{t}`$ and extend it multiplicatively to $`f\left(e_{p,q}\overline{t}\right)`$. Then $`f\left(\overline{t}\right)f\left(\overline{t}\right)`$ $`=`$ $`y\overline{t}y\overline{t}=x^2\overline{t}^2=x^2x^{1\left(k2\right)}\left({\displaystyle \frac{1+i}{2}}+{\displaystyle \frac{1i}{2}}x^2\right)`$ $`=`$ $`x^{1k}\left({\displaystyle \frac{1+i}{2}}+{\displaystyle \frac{1i}{2}}x^2\right)=f\left(\overline{t}^2\right)`$ $`f\left(\overline{t}y\right)`$ $`=`$ $`f\left(x^2y\overline{t}\right)=x^2yy\overline{t}=y\overline{t}y=f\left(\overline{t}\right)f\left(y\right)`$ $`\left(ff\right)\mathrm{\Delta }\left(\overline{t}\right)`$ $`=`$ $`\left(ff\right)\left(\theta \left(t\right)\overline{t}\overline{t}\right)=\theta \left(t\right)\left(f\left(\overline{t}\right)f\left(\overline{t}\right)\right)=\theta \left(t\right)\left(y\overline{t}y\overline{t}\right)`$ $`=`$ $`\left(yy\right)\theta \left(t\right)\left(\overline{t}\overline{t}\right)=\mathrm{\Delta }\left(y\right)\mathrm{\Delta }\left(\overline{t}\right)=\mathrm{\Delta }\left(y\overline{t}\right)=\mathrm{\Delta }\left(f\left(\overline{t}\right)\right)`$ and such a $`f`$ is a Hopf algebra isomorphism between $`H_{k+2}`$ and $`H_k`$. Thus there are exactly $`2`$ nonisomorphic Hopf algebras of this type: 1. $`H_{c:\sigma _0}=H_0`$ with cocycle $`\sigma _0`$ defined by $`\sigma (t,t)=\frac{1+i}{2}x+\frac{1i}{2}x^3`$ and $`𝐆\left(H_{c:\sigma _0}^{}\right)=\chi \times \phi C_4\times C_2`$, where $`\chi (x)=1`$, $`\chi (y)=1`$, $`\chi (t)=i`$, $`\phi \left(y\right)=1`$, $`\phi \left(x\right)=\phi \left(t\right)=1`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ with the property $`\pi ^2=\pi \pi =\chi +\chi ^3+\chi \phi +\chi ^3\phi `$. 2. $`H_{c:\sigma _1}=H_1`$ with cocycle $`\sigma _1`$ defined by $`\sigma (t,t)=\frac{1+i}{2}1+\frac{1i}{2}x^2`$ and $`𝐆\left(H_{c:\sigma _1}^{}\right)=\chi \times \phi \times \psi C_2\times C_2\times C_2`$, where $`\chi (y)=1`$, $`\chi (x)=\chi (t)=1`$, $`\phi \left(x\right)=\phi \left(y\right)=1`$, $`\phi \left(t\right)=1`$, $`\psi \left(x\right)=1`$, $`\psi \left(y\right)=\psi \left(t\right)=1`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill \omega & \hfill 0\\ \hfill 0& \hfill \omega \end{array}\right)`$ where $`\omega `$ is a primitive $`8^{th}`$-root of unity, with the property $`\pi ^2=\pi \pi =\psi +\chi \psi +\phi \psi +\chi \phi \psi `$. ### 3.2. Case of $`𝐆\left(H\right)=C_2\times C_2\times C_2`$. We will show that there are at most $`4`$ possible Hopf algebras of this kind. Let $`H`$ be a nontrivial semisimple Hopf algebra of dimension $`16`$ with a subHopfalgebra $`K=k\left(C_2\times C_2\times C_2\right)^{}k\left(C_2\times C_2\times C_2\right)`$. Then $`𝐆\left(H\right)=GC_2\times C_2\times C_2`$. Let $`𝐆\left(H\right)=x\times y\times z`$, where $`\left|x\right|=\left|y\right|=\left|z\right|=2`$. Then the dual basis of $`KK^{}`$ is given by $$e_{p,q,r}=1/8\left(1+\left(1\right)^px\right)\left(1+\left(1\right)^qy\right)\left(1+\left(1\right)^rz\right),\text{ }p,q,r=0,1$$ Then $`\mathrm{\Delta }_H\left(e_{p,q,r}\right)`$ $`=`$ $`\mathrm{\Delta }_K\left(e_{p,q,r}\right)={\displaystyle \underset{\begin{array}{c}p_1+p_2pmod2\\ q_1+q_2qmod2\\ r_1+r_2rmod2\end{array}}{}}e_{p_1,q_1,r_1}e_{p_2,q_2,r_2}`$ $`\mathrm{\Delta }_H\left(\overline{t}\right)`$ $`=`$ $`\theta \left(t\right)\overline{t}\overline{t}`$ where $`\overline{t}=1\mathrm{\#}t.`$ Dualizing $`\left(\text{3.1}\right)`$ we get another extension $$F^{}\stackrel{\pi ^{}}{}H^{}\stackrel{i^{}}{}K^{}$$ and as in \[11, 2.4\], \[12, 2.11\] or \[15, 2.1\], since $`k`$ is algebraically closed, there exist units $`\overline{x}`$, $`\overline{y}`$ and $`\overline{z}H^{}`$, such that $`\overline{x}^2=\overline{y}^2=\overline{z}^2=1_H^{}`$, $`e_{p,q,r},\overline{x}^i\overline{y}^j\overline{z}^k=\delta _{ip}\delta _{jq}\delta _{kr}`$ and $`\alpha =\overline{z}^1\overline{y}^1\overline{z}\overline{y},\beta =\overline{z}^1\overline{x}^1\overline{z}\overline{x},\gamma =\overline{y}^1\overline{x}^1\overline{y}\overline{x}F^{}=k\{e_0,e_1\}`$, where $`\left\{e_r\right\}`$ is a dual basis of $`\left\{t^r\right\}`$, $`r=0,1`$. $`\epsilon \left(\alpha \right)=\epsilon \left(\beta \right)=\epsilon \left(\gamma \right)=1`$ and therefore $`\alpha =e_0+\xi _3e_1,\beta =e_0+\xi _2e_1,\gamma =e_0+\xi _1e_1`$. The right action $`\rho ^{}:F^{}K^{}F^{}`$ is trivial, thus $`F^{}`$ lies in the center of $`H^{}`$. $$\overline{x}=\overline{y}^2\overline{x}=\overline{y}\overline{x}\overline{y}\gamma =\overline{x}\overline{y}\gamma \overline{y}\gamma =\overline{x}\overline{y}^2\gamma ^2=\overline{x}\gamma ^2$$ Thus $`\gamma ^2=1`$ and similarly $`\alpha ^2=\beta ^2=1`$. Therefore $`\xi _1,\xi _2,\xi _3=\pm 1`$ and, since $`H^{}`$ is non-commutative, they cannot be all equal to 1. $`\mathrm{\Delta }_H\left(\overline{t}\right),\overline{x}^i\overline{y}^j\overline{z}^ke_l\overline{x}^p\overline{y}^q\overline{z}^re_s`$ $`=`$ $`\overline{t},\overline{x}^i\overline{y}^j\overline{z}^ke_l\overline{x}^p\overline{y}^q\overline{z}^re_s`$ $`=`$ $`\delta _{ls}\overline{t},\overline{x}^{i+p}\overline{y}^{j+q}\overline{z}^{k+r}\alpha ^{kq}\beta ^{kp}\gamma ^{jp}e_l=\xi _1^{jp}\xi _2^{kp}\xi _3^{kq}\delta _{l1}\delta _{s1}`$ On the other hand $`\mathrm{\Delta }_H\left(\overline{t}\right),\overline{x}^i\overline{y}^j\overline{z}^ke_l\overline{x}^p\overline{y}^q\overline{z}^re_s`$ $`=`$ $`\theta \left(t\right)\overline{t}\overline{t},\overline{x}^i\overline{y}^j\overline{z}^ke_l\overline{x}^p\overline{y}^q\overline{z}^re_s`$ $`=`$ $`\theta \left(t\right),\overline{x}^i\overline{y}^j\overline{z}^k\overline{x}^p\overline{y}^q\overline{z}^r\delta _{k1}\delta _{r1}`$ Therefore $$\theta \left(t\right)=\underset{ijkpqr}{}\xi _1^{jp}\xi _2^{kp}\xi _3^{kq}e_{i,j,k}e_{p,q,r}$$ Action by $`t`$ is a Hopf algebra map and therefore $`tG=G`$ and $`f_t:GG`$ defined by $`f_t\left(g\right)=tg`$ is a group automorphism of order $`2`$. Then, without loss of generality there is only one possibility for such an automorphism: $`t`$ $``$ $`x=y`$ $`t`$ $``$ $`y=x`$ $`t`$ $``$ $`z=z`$ Then $`te_{i,j,k}=e_{j,i,k}`$. Write $`v=\sigma (t,t)=c_{i,j,k}e_{i,j,k}`$ then $`c_{0,0,0}=\epsilon \left(v\right)=1`$ and $`c_{i,j,k}0,`$ since $`v`$ is a unit. By formula $`\left(\text{3.2}\right)`$ (3.14) $$c_{i,j,k}=c_{j,i,k}$$ For $`H`$ to be a bialgebra we need $`\mathrm{\Delta }_H\left(\overline{t}^2\right)=\mathrm{\Delta }_H\left(\overline{t}\right)\mathrm{\Delta }_H\left(\overline{t}\right)`$ $$\mathrm{\Delta }_H\left(\overline{t}^2\right)=\mathrm{\Delta }_H\left(v\right)=\mathrm{\Delta }_K\left(c_{i,j,k}e_{i,j,k}\right)=c_{i+p,j+q,k+r}e_{i,j,k}e_{p,q,r}$$ On the other hand, $`\mathrm{\Delta }_H\left(\overline{t}\right)\mathrm{\Delta }_H\left(\overline{t}\right)=\left(\theta \left(t\right)\overline{t}\overline{t}\right)\left(\theta \left(t\right)\overline{t}\overline{t}\right)`$ $`=`$ $`{\displaystyle \underset{ijpq}{}}\xi _1^{jp}\xi _2^{kp}\xi _3^{kq}e_{ijk}e_{pqr}\left({\displaystyle \underset{ijpq}{}}\xi _1^{jp}\xi _2^{kp}\xi _3^{kq}\left(te_{ijk}\right)\left(te_{pqr}\right)\right)\sigma (t,t)\sigma (t,t)`$ $`=`$ $`{\displaystyle \xi _1^{jp+iq}\xi _2^{kp+kq}\xi _3^{kq+kp}c_{i,j,k}c_{p,q,r}e_{i,j,k}e_{p,q,r}}`$ Therefore (3.15) $$c_{i+p,j+q,k+r}=\xi _1^{jp+iq}\xi _2^{kp+kq}\xi _3^{kq+kp}c_{i,j,k}c_{p,q,r}$$ Conditions $`\left(\text{3.14}\right)`$ and $`\left(\text{3.15}\right)`$ imply that $`c_{1,0,0}^2=c_{0,1,0}^2=c_{0,0,1}^2=c_{1,1,0}^2=c_{1,1,1}^2=1`$ and $`c_{0,1,0}=c_{1,0,0}`$ $`c_{1,1,0}`$ $`=`$ $`\xi _1c_{0,1,0}c_{1,0,0}=\xi _1c_{1,0,0}^2=\xi _1`$ $`c_{1,0,1}`$ $`=`$ $`c_{1,0,0}c_{0,0,1}`$ $`c_{1,0,1}`$ $`=`$ $`c_{0,0,1}c_{1,0,0}\xi _2\xi _3`$ Thus $`\xi _2\xi _3=1`$, that is $`\xi _2=\xi _3`$ and $`c_{1,0,0}=c_{0,1,0}=\omega =\pm 1`$ and $`c_{0,0,1}=\tau =\pm 1`$ and $$\theta \left(t\right)=\underset{ijkpqr}{}\xi _1^{jp}\xi _2^{kp+kq}e_{i,j,k}e_{p,q,r}$$ $`\sigma (t,t)`$ $`=`$ $`e_{0,0,0}+\tau e_{0,0,1}+\xi _1e_{1,1,0}+\xi _1\tau e_{1,1,1}+\omega \left(e_{1,0,0}+e_{0,1,0}+\tau e_{1,0,1}+\tau e_{0,1,1}\right)`$ $`=`$ $`{\displaystyle \omega ^{p+q}e_{pqr}\xi _1^{pq}e_{pqr}\tau ^re_{pqr}}=\left(xy\right)^{\delta _{\omega ,1}}\left(1+x+yxy\right)^{\delta _{\xi _1,1}}z^{\delta _{\tau ,1}}`$ For $`\xi _1,\xi _2,\tau ,\omega =\pm 1`$ let $`H_{d:\xi _1,\xi _2,\tau ,\omega }`$ be the Hopf algebras with the structures described above with cocycles $`\sigma _{\xi _1,\tau ,\omega }`$. Then $`\sigma _{\xi _1,\tau ,1}(t,t)=xy\sigma _{\xi _1,\tau ,1}(t,t)`$. Define $`f:H_{d:\xi _1,\xi _2,\tau ,1}`$ $``$ $`H_{d:\xi _1,\xi _2,\tau ,1}\text{by}`$ $`f\left(e_{p,q,r}\right)`$ $`=`$ $`e_{p,q,r}`$ $`f\left(\overline{t}\right)`$ $`=`$ $`x\overline{t}`$ and extend it multiplicatively to $`f\left(e_{p,q,r}\overline{t}\right)`$. Then $`f\left(\overline{t}\right)f\left(\overline{t}\right)`$ $`=`$ $`x\overline{t}x\overline{t}=xy\overline{t}^2=xy\sigma _{\xi _1,\tau ,1}(t,t)`$ $`=`$ $`\sigma _{\xi _1,\tau ,1}(t,t)=f\left(\sigma _{\xi _1,\tau ,1}(t,t)\right)=f\left(\overline{t}^2\right)`$ $`f\left(\overline{t}x\right)`$ $`=`$ $`f\left(y\overline{t}\right)=yx\overline{t}=x\overline{t}x=f\left(\overline{t}\right)f\left(x\right)`$ $`f\left(\overline{t}y\right)`$ $`=`$ $`f\left(x\overline{t}\right)=x^2\overline{t}=x\overline{t}y=f\left(\overline{t}\right)f\left(y\right)`$ $`\left(ff\right)\mathrm{\Delta }\left(\overline{t}\right)`$ $`=`$ $`\left(ff\right)\left(\theta \left(t\right)\overline{t}\overline{t}\right)=\theta \left(t\right)\left(f\left(\overline{t}\right)f\left(\overline{t}\right)\right)=\theta \left(t\right)\left(x\overline{t}x\overline{t}\right)`$ $`=`$ $`\left(xx\right)\theta \left(t\right)\left(\overline{t}\overline{t}\right)=\mathrm{\Delta }\left(x\right)\mathrm{\Delta }\left(\overline{t}\right)=\mathrm{\Delta }\left(x\overline{t}\right)=\mathrm{\Delta }\left(f\left(\overline{t}\right)\right)`$ and such a $`f`$ is a Hopf algebra isomorphism between $`H_{d:\xi _1,\xi _2,\tau ,1}`$ and $`H_{d:\xi _1,\xi _2,\tau ,1}`$. Define $`f^{}:H_{d:1,1,\tau ,1}`$ $``$ $`H_{d:1,1,\tau ,1}\text{by}`$ $`f^{}\left(e_{p,q,r}\right)`$ $`=`$ $`e_{p+r,q+r,r}`$ $`f^{}\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(1+z+iyiyz\right)\overline{t}={\displaystyle i^{r^2}\left(1\right)^{qr}e_{p,q,r}\overline{t}}`$ and extend it multiplicatively to $`f^{}\left(e_{p,q,r}\overline{t}\right)`$. Then $`f^{}_{𝐆\left(H_{d:1,1,\tau ,1}\right)}`$ is a group isomorphism $`𝐆\left(H_{d:1,1,\tau ,1}\right)𝐆\left(H_{d:1,1,\tau ,1}\right)`$ with $`f^{}\left(x\right)=xz`$, $`f^{}\left(y\right)=yz`$, $`f^{}\left(z\right)=z`$ and $`f^{}\left(\overline{t}\right)f^{}\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(\left(1+z\right)+iy\left(1z\right)\right)\overline{t}\left(\left(1+z\right)+iy\left(1z\right)\right)\overline{t}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(\left(1+z\right)+iy\left(1z\right)\right)\left(\left(1+z\right)+ix\left(1z\right)\right)\overline{t}^2`$ $`=`$ $`{\displaystyle \frac{1}{8}}\left(2+2zxy\left(22z\right)\right)\left(1+x+yxy\right)z^{\delta _{\tau ,1}}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(\left(1xy\right)+z\left(1+xy\right)\right)\left(\left(1xy\right)+x\left(1+xy\right)\right)z^{\delta _{\tau ,1}}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(22xy+xz\left(2+2xy\right)\right)z^{\delta _{\tau ,1}}={\displaystyle \frac{1}{2}}\left(1+xz+yzxy\right)z^{\delta _{\tau ,1}}`$ $`=`$ $`f\left({\displaystyle \frac{1}{2}}\left(1+x+yxy\right)z^{\delta _{\tau ,1}}\right)=f^{}\left(\overline{t}^2\right)`$ $`f^{}\left(\overline{t}x\right)`$ $`=`$ $`f^{}\left(y\overline{t}\right)={\displaystyle \frac{1}{2}}yz\left(1+z+iyiyz\right)\overline{t}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(1+z+iyiyz\right)\overline{t}xz=f^{}\left(\overline{t}\right)f^{}\left(x\right)`$ $`f^{}\left(\overline{t}y\right)`$ $`=`$ $`f^{}\left(x\overline{t}\right)={\displaystyle \frac{1}{2}}xz\left(1+z+iyiyz\right)\overline{t}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(1+z+iyiyz\right)\overline{t}yz=f^{}\left(\overline{t}\right)f^{}\left(y\right)`$ $`\left(f^{}f^{}\right)\mathrm{\Delta }\left(\overline{t}\right)`$ $`=`$ $`\left(f^{}f^{}\right)\left(\theta \left(t\right)\overline{t}\overline{t}\right)`$ $`=`$ $`\left(f^{}f^{}\right)\left({\displaystyle \left(1\right)^{bp}\left(1\right)^{cp+cq}e_{a,b,c}\overline{t}e_{p,q,r}\overline{t}}\right)`$ $`=`$ $`\left({\displaystyle \left(1\right)^{bp}\left(1\right)^{c\left(p+q\right)}e_{a+c,b+c,c}e_{p+r,q+r,r}}\right)`$ $`\times \left({\displaystyle i^{n^2}\left(1\right)^{mn}e_{l,m,n}\overline{t}i^{n^2}\left(1\right)^{mn}e_{l,m,n}\overline{t}}\right)`$ $`=`$ $`\left({\displaystyle \left(1\right)^{\left(b+c\right)\left(p+r\right)}\left(1\right)^{c\left(p+q\right)}e_{a,b,c}e_{p,q,r}}\right)`$ $`\times \left({\displaystyle i^{n^2}\left(1\right)^{mn}e_{l,m,n}\overline{t}i^{n^2}\left(1\right)^{mn}e_{l,m,n}\overline{t}}\right)`$ $`=`$ $`{\displaystyle \left(1\right)^{bp+cp+br+cr}\left(1\right)^{cp+cq}i^{c^2}\left(1\right)^{bc}i^{r^2}\left(1\right)^{qr}e_{a,b,c}\overline{t}e_{p,q,r}\overline{t}}`$ $`=`$ $`{\displaystyle \left(1\right)^{bp+br+cq+bc+qr}\left(1\right)^{cr}i^{c^2}i^{r^2}e_{a,b,c}\overline{t}e_{p,q,r}\overline{t}}`$ $`=`$ $`{\displaystyle \left(1\right)^{bp+br+cq+bc+qr}\left(1\right)^{cr}i^{c^2}i^{r^2}e_{a,b,c}\overline{t}e_{p,q,r}\overline{t}}`$ $`=`$ $`{\displaystyle i^{\left(c+r\right)^2}\left(1\right)^{\left(b+q\right)\left(c+r\right)}\left(1\right)^{bp}e_{a,b,c}\overline{t}e_{p,q,r}\overline{t}}`$ $`=`$ $`{\displaystyle i^{n^2}\left(1\right)^{mn}\underset{\begin{array}{c}l_1+l_2=l\\ m_1+m_2=m\\ n_1+n_2=n\end{array}}{}e_{l_1,m_1,n_1}e_{l_2,m_2,n_2}}`$ $`\times {\displaystyle }(1)^{bp}e_{a,b,c}\overline{t}e_{p,q,r}\overline{t}`$ $`=`$ $`{\displaystyle i^{n^2}\left(1\right)^{mn}\mathrm{\Delta }\left(e_{l,m,n}\right)\mathrm{\Delta }\left(\overline{t}\right)}`$ $`=`$ $`\mathrm{\Delta }\left({\displaystyle i^{n^2}\left(1\right)^{mn}e_{l,m,n}\overline{t}}\right)=\mathrm{\Delta }f^{}\left(\overline{t}\right)`$ and such an $`f^{}`$ is a Hopf algebra isomorphism between $`H_{d:1,1,\tau ,1}`$ and $`H_{d:1,1,\tau ,1}`$. Thus we may assume that $`\omega =1`$ and $`\xi _2=1`$. Therefore there are at most four nonisomorphic Hopf algebras $`H_{d:\xi _1,\tau }`$ of this kind, $`H_{d:1,1}`$, $`H_{d:1,1}`$, $`H_{d:1,1}`$ and $`H_{d:1,1}`$: 1. $`H_{d:1,1}`$ with the trivial cocycle and $`𝐆\left(H_{d:1,1}^{}\right)=\chi \times \phi \times \psi C_2\times C_2\times C_2`$, where $`\chi \left(x\right)=\chi \left(y\right)=\chi \left(z\right)=1`$, $`\chi \left(t\right)=1`$, $`\phi (x)=\phi (y)=1`$, $`\phi (z)=\phi (t)=1`$, $`\psi \left(z\right)=1`$, $`\psi \left(x\right)=\psi \left(y\right)=\psi \left(t\right)=1`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (z)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi ^2=\pi \pi =1+\chi +\phi +\chi \phi `$. 2. $`H_{d:1,1}`$ with the cocycle defined by $`\sigma (t,t)=z`$ and $`𝐆\left(H_{d:1,1}^{}\right)=\chi \times \phi C_4\times C_2`$, where $`\chi (x)=\chi (y)=1`$, $`\chi (z)=1`$, $`\chi (t)=i`$, $`\phi \left(x\right)=\phi \left(y\right)=1`$, $`\phi \left(t\right)=\phi \left(z\right)=1`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (z)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi ^2=\pi \pi =1+\chi ^2+\phi +\chi ^2\phi `$. 3. $`H_{d:1,1}`$ with the cocycle defined by $`\sigma (t,t)=1+x+yxy`$ and $`𝐆\left(H_{d:1,1}^{}\right)=\chi \times \phi \times \psi C_2\times C_2\times C_2`$, where $`\chi \left(x\right)=\chi \left(y\right)=\chi \left(z\right)=1`$, $`\chi \left(t\right)=1`$, $`\phi (x)=\phi (y)=1`$, $`\phi (z)=1`$, $`\phi (t)=i`$, $`\psi \left(z\right)=1`$, $`\psi \left(x\right)=\psi \left(y\right)=\psi \left(t\right)=1`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (z)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi ^2=\pi \pi =1+\chi +\phi +\chi \phi `$. 4. $`H_{d:1,1}`$ with the cocycle defined by $`\sigma (t,t)=\left(1+x+yxy\right)z`$ and $`𝐆\left(H_{d:1,1}^{}\right)=\chi \times \phi C_4\times C_2`$, where $`\chi \left(x\right)=\chi \left(y\right)=1`$, $`\chi \left(z\right)=1`$, $`\chi \left(t\right)=i`$, $`\phi (x)=\phi (y)=1`$, $`\phi (z)=1`$, $`\phi (t)=i`$. There is a degree $`2`$ irreducible representation defined by $`\pi (x)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (y)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (z)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi (t)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi ^2=\pi \pi =1+\chi ^2+\phi +\chi ^2\phi `$. ### 3.3. Case of $`G=D_8`$. Let $`G=D_8=x,yx^4=y^2=1,yx=x^1y`$. Let $`\left\{e_{pq}\right\}_{p=0,1,2,3;q=0,1}`$ be the basis of $`K`$, dual to the basis $`\left\{x^py^q\right\}_{p=0,1,2,3;q=0,1}`$ of $`K^{}=kD_8`$. Then $$\mathrm{\Delta }_H\left(e_{pq}\right)=\mathrm{\Delta }_K\left(e_{pq}\right)=\underset{\begin{array}{c}p_1+p_2+2q_1p_2pmod4\\ q_1+q_2qmod2\end{array}}{}e_{p_1q_1}e_{p_2q_2}$$ and it is easy to check that elements $`X`$ $`=`$ $`{\displaystyle \underset{pq}{}}\left(1\right)^pe_{pq}`$ $`Y`$ $`=`$ $`{\displaystyle \underset{pq}{}}\left(1\right)^qe_{pq}`$ are grouplike of order $`2`$. For $`\overline{t}=1\mathrm{\#}t`$ $$\mathrm{\Delta }_H\left(\overline{t}\right)=\theta \left(t\right)\overline{t}\overline{t}$$ Dualizing $`\left(\text{3.1}\right)`$ we get another extension $$F^{}\stackrel{\pi ^{}}{}H^{}\stackrel{i^{}}{}K^{}$$ and as in \[11, 2.4\], \[12, 2.11\] or \[15, 2.1\], since $`k`$ is algebraically closed, there exist units $`\overline{x}`$ and $`\overline{y}H^{}`$, such that $`\overline{x}^4=\overline{y}^2=1_H^{}`$, $`e_{pq},\overline{x}^i\overline{y}^j=\delta _{ip}\delta _{jq}`$ and $`\alpha =\overline{y}\overline{x}^2\overline{y}\overline{x}^2F^{}=k\{e_0,e_1\}`$, where $`\left\{e_r\right\}`$ is a dual basis of $`\left\{t^r\right\}`$, $`r=0,1`$. The right action $`\rho ^{}:F^{}K^{}F^{}`$ is trivial, thus $`F^{}`$ lies in the center of $`H^{}`$. $$\overline{x}^2=\overline{y}^2\overline{x}^2=\overline{y}\overline{x}^2\overline{y}\alpha =\overline{x}^2\overline{y}\alpha \overline{y}\alpha =\overline{x}^2\overline{y}^2\alpha ^2=\overline{x}^2\alpha ^2$$ Thus $`\alpha ^2=1`$. Consider $`\beta =\overline{y}\overline{x}\overline{y}\overline{x}F^{}=k\{e_0,e_1\}`$. $`\epsilon \left(\beta \right)=\epsilon \left(\overline{y}^1\overline{x}^1\overline{y}\overline{x}\right)=1`$ and therefore $`\beta =e_0+\xi e_1`$. Moreover, $`\overline{y}\overline{x}\overline{y}\overline{x}^1=\beta \overline{x}^2`$ and $$\overline{x}=\overline{y}^2\overline{x}=\overline{y}\beta \overline{x}^2\overline{x}\overline{y}=\overline{y}\beta \overline{x}^2\overline{y}\beta \overline{x}^2\overline{x}=\overline{y}\overline{x}^2\overline{y}\overline{x}^3\beta ^2=\overline{y}^2\alpha \overline{x}^2\overline{x}^3\beta ^2=\overline{x}\alpha \beta ^2$$ Thus $`\beta ^2=\alpha ^1=\alpha `$, implying $`\beta ^4=1`$ and $`\xi =\pm 1`$ or $`\pm i`$. $`\mathrm{\Delta }_H\left(\overline{t}\right),\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r`$ $`=`$ $`\overline{t},\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r=\delta _{kr}\overline{t},\overline{x}^{i+p}\beta ^{jp}\overline{x}^{2jp}\overline{y}^{j+q}e_k`$ $`=`$ $`\delta _{kr}\overline{t},\overline{x}^{i+p+2jp}\overline{y}^{j+q}\beta ^{jp}e_k=\xi ^{jp}\delta _{k1}\delta _{r1}`$ On the other hand $$\mathrm{\Delta }_H\left(\overline{t}\right),\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r=\theta \left(t\right)\overline{t}\overline{t},\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r=\theta \left(t\right),\overline{x}^i\overline{y}^j\overline{x}^p\overline{y}^q\delta _{k1}\delta _{r1}$$ Therefore $$\theta \left(t\right)=\underset{ijpq}{}\xi ^{jp}e_{ij}e_{pq}$$ It is easy to check that if $`\xi =\pm i`$ then $`1,X,Y`$ and $`XY`$ are the only grouplikes of $`H`$. Write $`v=\sigma (t,t)=c_{i,j}e_{i,j}`$ then $`c_{0,0}=\epsilon \left(v\right)=1`$ and $`c_{i,j}0,`$ since $`v`$ is a unit and $$\mathrm{\Delta }_H\left(\overline{t}^2\right)=\mathrm{\Delta }_H\left(v\right)=\mathrm{\Delta }_K\left(c_{i,j}e_{i,j}\right)=c_{p+r+2rq,q+s}e_{p,q}e_{r,s}$$ On the other hand, if we write $$te_{p,q}=e_{\alpha _1(p,q),\alpha _2(p,q)}$$ $`\mathrm{\Delta }\left(\overline{t}\right)\mathrm{\Delta }\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \underset{pqrs}{}}\xi ^{qr}e_{p,q}\overline{t}e_{r,s}\overline{t}{\displaystyle \underset{pqrs}{}}\xi ^{qr}e_{p,q}\overline{t}e_{r,s}\overline{t}`$ $`=`$ $`{\displaystyle \underset{pqrs}{}}\xi ^{qr}e_{p,q}e_{r,s}{\displaystyle \underset{pqrs}{}}\xi ^{qr}e_{\alpha _1(p,q),\alpha _2(p,q)}\overline{t}^2e_{\alpha _1(r,s),\alpha _2(r,s)}\overline{t}^2`$ $`=`$ $`{\displaystyle \underset{pqrs}{}}\xi ^{qr}e_{p,q}e_{r,s}{\displaystyle \underset{pqrs}{}}\xi ^{\alpha _2(p,q)\alpha _1(r,s)}e_{p,q}\overline{t}^2e_{r,s}\overline{t}^2`$ $`=`$ $`{\displaystyle \underset{pqrs}{}}\xi ^{qr+\alpha _2(p,q)\alpha _1(r,s)}c_{pq}c_{rs}e_{p,q}e_{r,s}`$ Thus for $`H`$ to be a bialgebra we should have (3.16) $$c_{p+r+2rq,q+s}=\xi ^{qr+\alpha _2(p,q)\alpha _1(r,s)}c_{pq}c_{rs}$$ Action by $`t`$ is a Hopf algebra map and therefore it induces a group automorphism $`f_t:GG`$ defined by $`e_{p,q},f_t\left(g\right)=te_{p,q},g`$, which has order $`2`$. $`f_t\left(x\right)=x`$ or $`x^1`$ since order of $`x`$ is $`4`$. If $`f_t\left(x\right)=x`$ then in order for $`f_t`$ to be of order $`2`$ we should have $`f_t\left(y\right)=x^2y`$. If $`f_t\left(x\right)=x^1`$ then renaming generators we are down to two choices for $`f_t\left(y\right)`$, namely $`f_t\left(y\right)=y`$ or $`xy`$. Thus there are three possibilities for the action of $`t`$; we consider them below: #### Case A) The action is given by $`te_{p,q}=e_{p+2q,q}`$, corresponding to $`f_t\left(x\right)`$ $`=`$ $`x`$ $`f_t\left(y\right)`$ $`=`$ $`x^2y`$ Then $`X`$ and $`Y`$ are central grouplikes of $`H`$. Write $`v=\sigma (t,t)=c_{p,q}e_{p,q}`$. By $`\left(\text{3.2}\right)`$ and $`\left(\text{3.16}\right)`$ (3.17) $`c_{p,q}`$ $`=`$ $`c_{p+2q,q}`$ (3.18) $`c_{p+r+2rq,q+s}`$ $`=`$ $`\xi ^{qr+q\left(r+2s\right)}c_{p,q}c_{r,s}\xi ^{2q\left(r+s\right)}c_{p,q}c_{r,s}`$ Conditions $`\left(\text{3.17}\right)`$ and $`\left(\text{3.18}\right)`$ imply that $`\xi ^2c_{0,1}c_{1,0}`$ $`=`$ $`c_{3,1}=c_{1,1}=c_{1,0}c_{0,1}`$ $`\xi ^2c_{0,1}c_{0,1}`$ $`=`$ $`c_{0,0}=1`$ $`c_{1,0}c_{1,0}`$ $`=`$ $`c_{2,0}`$ $`c_{2,0}c_{0,1}`$ $`=`$ $`c_{2,1}=c_{0,1}`$ Thus $`\xi ^2=1`$, $`c_{2,0}=1`$ and $`c_{1,0}^2=c_{0,1}^2=1.`$ Therefore $`c_{1,0}=\left(1\right)^k`$ and $`c_{0,1}=\left(1\right)^l`$ for $`k,l=0,1`$ and (3.19) $$\sigma (t,t)=\left(1\right)^{kp}\left(1\right)^{lq}e_{p,q}=\left(1\right)^{kp}e_{p,q}\left(1\right)^{ls}e_{r,s}=X^kY^l,k,l=0,1$$ If $`\xi =1`$ then $`\overline{t}`$ is a grouplike of $`H`$, if $`\xi =1`$ then $`i^pe_{p,q}\overline{t}`$ is a grouplike of $`H`$. In both cases $`𝐆\left(H\right)`$ is abelian of order $`8`$ and $`H`$ was described in Section 3.1 or Section 3.2. #### Case B) The action is given by $`te_{p,q}=e_{p,q}`$, corresponding to $`f_t\left(x\right)`$ $`=`$ $`x^1`$ $`f_t\left(y\right)`$ $`=`$ $`y`$ Then $`X`$ and $`Y`$ are central grouplikes of $`H`$. Write $`v=\sigma (t,t)=c_{p,q}e_{p,q}`$ . By $`\left(\text{3.2}\right)`$ and $`\left(\text{3.16}\right)`$ (3.20) $`c_{p,q}`$ $`=`$ $`c_{p,q}`$ (3.21) $`c_{p+r+2rq,q+s}`$ $`=`$ $`\xi ^{qrqr}c_{p,q}c_{r,s}=c_{p,q}c_{r,s}`$ Conditions $`\left(\text{3.20}\right)`$ and $`\left(\text{3.21}\right)`$ imply that $`c_{1,0}=\left(1\right)^k`$ and $`c_{0,1}=\left(1\right)^l`$ for $`k,l=0,1`$ and (3.22) $$\sigma (t,t)=\left(1\right)^{kp}\left(1\right)^{lq}e_{p,q}=\left(1\right)^{kp}e_{p,q}\left(1\right)^{ls}e_{r,s}=X^kY^l,k,l=0,1$$ If $`\xi =1`$ then $`\overline{t}`$ is a grouplike of $`H`$, if $`\xi =1`$ then $`i^pe_{p,q}\overline{t}`$ is a grouplike of $`H`$. In both cases $`𝐆\left(H\right)`$ is abelian of order $`8`$ and $`H`$ was described in Section 3.1 or Section 3.2. So now we will consider only $`\xi =\pm i`$. For $`k,l=0,1`$ let $`H_{\xi ,X^kY^l}`$ be the Hopf algebras with the structures described above with cocycles $`\sigma _{k,l}(t,t)=X^kY^l`$. Define $`f:H_{\xi ,X^kY^l}`$ $``$ $`H_{\xi ,X^kY^l}\text{by}`$ $`f\left(e_{r,s}\right)`$ $`=`$ $`e_{r,s}`$ $`f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle i^pe_{p,q}\overline{t}}`$ and extend it multiplicatively to $`f\left(e_{r,s}\overline{t}\right)`$. Then $`f\left(\overline{t}\right)f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle i^pe_{p,q}\overline{t}i^pe_{p,q}\overline{t}}={\displaystyle i^pe_{p,q}i^pe_{p,q}\overline{t}^2}=\overline{t}^2=f\left(\overline{t}^2\right)`$ $`f\left(\overline{t}e_{r,s}\right)`$ $`=`$ $`f\left(e_{r,s}\overline{t}\right)=e_{r,s}{\displaystyle i^pe_{p,q}\overline{t}}={\displaystyle i^pe_{p,q}\overline{t}e_{r,s}}=f\left(\overline{t}\right)f\left(e_{r,s}\right)`$ $`\mathrm{\Delta }\left(f\left(\overline{t}\right)\right)`$ $`=`$ $`\mathrm{\Delta }\left({\displaystyle i^pe_{p,q}\overline{t}}\right)=\mathrm{\Delta }\left({\displaystyle i^pe_{p,q}}\right)\mathrm{\Delta }\left(\overline{t}\right)=`$ $`=`$ $`\left({\displaystyle i^{p_1+p_2+2q_1p_2}e_{p_1,q_1}e_{p_2,q_2}}\right)\left({\displaystyle \xi ^{s_1r_2}e_{r_1,s_1}\overline{t}e_{r_2,s_2}\overline{t}}\right)`$ $`=`$ $`\left({\displaystyle i^{p_1+p_2+2q_1p_2}\xi ^{q_1p_2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}\right)`$ $`=`$ $`{\displaystyle i^{p_1+p_2}\left(\xi \right)^{q_1p_2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}`$ $`=`$ $`{\displaystyle \left(\xi \right)^{q_1p_2}e_{p_1,q_1}f\left(\overline{t}\right)e_{p_2,q_2}f\left(\overline{t}\right)}`$ $`=`$ $`\left(ff\right)\left({\displaystyle \left(\xi \right)^{q_1p_2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}\right)=\left(ff\right)\mathrm{\Delta }\left(\overline{t}\right)`$ and such an $`f`$ is a Hopf algebra isomorphism between $`H_{\xi ,X^kY^l}`$ and $`H_{\xi ,X^kY^l}`$. Thus we may assume that $`\xi =i`$ and write $`H_{i,X^kY^l}=H_{X^kY^l}`$. Define $`f^{}:H_{X^kY}`$ $``$ $`H_{X^k}\text{by}`$ $`f^{}\left(e_{r,s}\right)`$ $`=`$ $`e_{r+2s,s}`$ $`f^{}\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle i^qe_{p,q}\overline{t}}=\left({\displaystyle \frac{1+i}{2}}1+{\displaystyle \frac{1i}{2}}Y\right)\overline{t}`$ and extend it multiplicatively to $`f^{}\left(e_{r,s}\overline{t}\right)`$. Note that restriction $`f^{}|_{\left(kD_8\right)^{}}`$ corresponds to the group automorphism $`f_t`$ described in Case A) and $`f^{}\left(X\right)=X,f^{}\left(Y\right)=Y`$. Then $`f^{}\left(\overline{t}\right)f^{}\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle i^qe_{p,q}\overline{t}i^qe_{p,q}\overline{t}}={\displaystyle i^qe_{p,q}i^qe_{p,q}\overline{t}^2}`$ $`=`$ $`{\displaystyle (1)^qe_{p,q}X^k}=YX^k=f^{}\left(YX^k\right)=f^{}\left(\overline{t}^2\right)`$ $`f^{}\left(\overline{t}e_{r,s}\right)`$ $`=`$ $`f^{}\left(e_{r,s}\overline{t}\right)=e_{r+2s,s}{\displaystyle i^qe_{p,q}\overline{t}}={\displaystyle i^qe_{p,q}\overline{t}e_{r2s,s}}=f^{}\left(\overline{t}\right)f^{}\left(e_{r,s}\right)`$ There are at most two nonisomorphic Hopf algebras of this kind: 1. $`H_{B:1}`$ with trivial cocycle and $`𝐆\left(H_{B:1}^{}\right)=\chi ,\phi D_8`$, where $`\phi (e_{r,s})=\delta _{r,2}\delta _{s,0}`$, $`\phi \left(\overline{t}\right)=1`$, $`\chi (e_{r,s})=\delta _{r,2}\delta _{s,1}`$, $`\chi (\overline{t})=1`$. There are two degree $`2`$ irreducible representations defined by $`\pi _1\left(e_{1,0}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _1\left(e_{3,0}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _1\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2\left(e_{1,1}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _2\left(e_{3,1}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _2\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi _k^2=\pi _k\pi _k=1+\chi ^2+\phi +\chi ^2\phi `$. 2. $`H_{B:X}`$ with the cocycle defined by $`\sigma _X(t,t)=X`$ and $`𝐆\left(H_{B:X}^{}\right)=\chi ,\phi D_8`$, where $`\chi (e_{r,s})=\delta _{r,2}\delta _{s,1}`$, $`\chi (\overline{t})=1`$, $`\phi (e_{r,s})=\delta _{r,2}\delta _{s,0}`$, $`\phi \left(\overline{t}\right)=1`$. There are two degree $`2`$ irreducible representations defined by $`\pi _1\left(e_{1,0}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _1\left(e_{3,0}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _1\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2\left(e_{1,1}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _2\left(e_{3,1}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _2\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi _k^2=\pi _k\pi _k=1+\chi ^2+\phi +\chi ^2\phi `$. #### Case C) The action is given by $`te_{p,q}=e_{p+q,q}`$, corresponding to $`f_t\left(x\right)`$ $`=`$ $`x^1`$ $`f_t\left(y\right)`$ $`=`$ $`xy`$ Then $`Y`$ is a central grouplike of $`H`$. Write $`v=\sigma (t,t)=c_{p,q}e_{p,q}`$ . By $`\left(\text{3.2}\right)`$ and $`\left(\text{3.16}\right)`$ (3.23) $`c_{p,q}`$ $`=`$ $`c_{p+q,q}`$ (3.24) $`c_{p+r+2rq,q+s}`$ $`=`$ $`\xi ^{qr+q\left(r+s\right)}c_{p,q}c_{r,s}=\xi ^{qs}c_{p,q}c_{r,s}`$ Conditions $`\left(\text{3.23}\right)`$ and $`\left(\text{3.24}\right)`$ imply that $`c_{0,1}`$ $`=`$ $`c_{1,1}`$ $`c_{2,1}`$ $`=`$ $`c_{3,1}`$ $`c_{1,0}`$ $`=`$ $`c_{3,0}`$ $`c_{1,0}c_{0,1}`$ $`=`$ $`c_{1,1}=c_{0,1}`$ $`c_{0,1}c_{1,0}`$ $`=`$ $`c_{3,1}`$ $`\xi c_{0,1}c_{0,1}`$ $`=`$ $`c_{0,0}=1`$ Thus $`c_{1,0}=1`$ and $`c_{0,1}=\omega ^k`$, where $`\omega `$ is a primitive $`8`$-th root of $`1`$ and $`\xi =\omega ^{2k}`$. Therefore $$\sigma _k(t,t)=\omega ^{kq}e_{p,q}=\frac{1+Y}{2}+\frac{\omega ^k\left(1Y\right)}{2},k=0,\mathrm{},7$$ For $`k=0,\mathrm{},7`$ let $`H_k`$ be the Hopf algebra with the structure described above with cocycle $`\sigma _k(t,t)`$. Define $`f:H_{k+2}`$ $``$ $`H_k\text{by}`$ $`f\left(e_{p,q}\right)`$ $`=`$ $`e_{p,q}`$ $`f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \underset{p,q}{}}i^pe_{p,q}\overline{t}`$ and extend it multiplicatively to $`f\left(e_{p,q}\overline{t}\right)`$. Then $`f\left(\overline{t}\right)f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle i^pe_{p,q}\overline{t}i^pe_{p,q}\overline{t}}={\displaystyle i^pe_{p,q}i^pe_{p+q,q}\overline{t}^2}`$ $`=`$ $`{\displaystyle i^pi^{p+q}e_{p,q}\sigma _k(t,t)}={\displaystyle i^qe_{p,q}\omega ^{kq}e_{p,q}}`$ $`=`$ $`{\displaystyle \omega ^{kq+2q}e_{p,q}}=\sigma _{k+2}(t,t)=f\left(\sigma _{k+2}(t,t)\right)=f\left(\overline{t}^2\right)`$ $`f\left(\overline{t}e_{p,q}\right)`$ $`=`$ $`f\left(e_{p+q,q}\overline{t}\right)=e_{p+q,q}{\displaystyle i^pe_{p,q}\overline{t}}={\displaystyle i^pe_{p,q}\overline{t}e_{p,q}}=f\left(\overline{t}\right)f\left(e_{p,q}\right)`$ $`\mathrm{\Delta }\left(f\left(\overline{t}\right)\right)`$ $`=`$ $`\mathrm{\Delta }\left({\displaystyle i^pe_{p,q}\overline{t}}\right)=\mathrm{\Delta }\left({\displaystyle i^pe_{p,q}}\right)\mathrm{\Delta }\left(\overline{t}\right)=`$ $`=`$ $`\left({\displaystyle i^{p_1+p_2+2q_1p_2}e_{p_1,q_1}e_{p_2,q_2}}\right)\left({\displaystyle \omega ^{2ks_1r_2}e_{r_1,s_1}\overline{t}e_{r_2,s_2}\overline{t}}\right)`$ $`=`$ $`\left({\displaystyle i^{p_1+p_2+2q_1p_2}\omega ^{2kq_1p_2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}\right)`$ $`=`$ $`{\displaystyle i^{p_1+p_2}\omega ^{2\left(k+2\right)q_1p_2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}`$ $`=`$ $`{\displaystyle \omega ^{2\left(k+2\right)q_1p_2}e_{p_1,q_1}f\left(\overline{t}\right)e_{p_2,q_2}f\left(\overline{t}\right)}`$ $`=`$ $`\left(ff\right)\left({\displaystyle \omega ^{2\left(k+2\right)q_1p_2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}\right)=\left(ff\right)\mathrm{\Delta }\left(\overline{t}\right)`$ and such a $`f`$ is a Hopf algebra isomorphism between $`H_k`$ and $`H_{k+2}`$. Thus there are exactly two nonisomorphic Hopf algebras of this type: 1. $`H_{C:1}=H_0`$ with a trivial cocycle and $`\xi =1`$. Then $`𝐆\left(H_{C:1}\right)=X\overline{t},XD_8`$ and $`𝐆\left(H_{C:1}^{}\right)=\chi \times \phi C_2\times C_2`$, where $`\chi (e_{p,q})=\delta _{p,2}\delta _{q,0}`$, $`\chi (\overline{t})=1`$, $`\phi (e_{p,q})=\delta _{p,0}\delta _{q,0}`$, $`\phi \left(\overline{t}\right)=1`$. There are three degree $`2`$ irreducible representations defined by $`\pi _1\left(e_{0,1}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _1\left(e_{1,1}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _1\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2\left(e_{1,0}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _2\left(e_{3,0}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _2\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _3\left(e_{2,1}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _3\left(e_{3,1}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _3\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi _2^2=\pi _2\pi _2=1+\chi +\phi +\chi \phi `$, $`\pi _1^2=\pi _3^2=1+\phi +\pi _2`$. 2. $`H_{C:\sigma _1}`$ with cocycle $`\sigma _1`$ defined by $`\sigma _1(t,t)=\omega ^qe_{p,q}`$ and $`\xi =\omega ^2`$, where $`\omega `$ is a primitive $`8`$-th root of $`1`$. Then $`𝐆\left(H_{C:\sigma _1}\right)=X\times YC_2\times C_2`$ and $`𝐆\left(H_{C:\sigma _1}^{}\right)=\chi \times \phi C_2\times C_2`$, where $`\chi (e_{p,q})=\delta _{p,2}\delta _{q,0}`$, $`\chi (\overline{t})=1`$, $`\phi (e_{p,q})=\delta _{p,0}\delta _{q,0}`$, $`\phi \left(\overline{t}\right)=1`$. There are three degree $`2`$ irreducible representations defined by $`\pi _1\left(e_{0,1}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _1\left(e_{1,1}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _1\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill \sqrt{\omega }\\ \hfill \sqrt{\omega }& \hfill 0\end{array}\right)`$ $`\pi _2\left(e_{1,0}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _2\left(e_{3,0}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _2\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _3\left(e_{2,1}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _3\left(e_{3,1}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _3\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill \sqrt{\omega }\\ \hfill \sqrt{\omega }& \hfill 0\end{array}\right)`$ with the property $`\pi _2^2=\pi _2\pi _2=1+\chi +\phi +\chi \phi `$, $`\pi _1^2=\pi _3^2=1+\phi +\pi _2`$. ### 3.4. Case of $`G=Q_8`$. Let $`G=Q_8=x,yx^4=1,y^2=x^2,yx=x^1y`$. Let $`\left\{e_{pq}\right\}_{p=0,1,2,3;q=0,1}`$ be the basis of $`K`$, dual to the basis $`\left\{x^py^q\right\}_{p=0,1,2,3;q=0,1}`$ of $`K^{}=kQ_8`$. Then $$\mathrm{\Delta }_H\left(e_{pq}\right)=\mathrm{\Delta }_K\left(e_{pq}\right)=\underset{\begin{array}{c}p_1+p_2+2q_1\left(p_2+q_2\right)pmod4\\ q_1+q_2qmod2\end{array}}{}e_{p_1q_1}e_{p_2q_2}$$ and it is easy to check that elements $`X`$ $`=`$ $`{\displaystyle \underset{pq}{}}\left(1\right)^pe_{pq}`$ $`Y`$ $`=`$ $`{\displaystyle \underset{pq}{}}\left(1\right)^qe_{pq}`$ are grouplike of order $`2`$. For $`\overline{t}=1\mathrm{\#}t`$ $$\mathrm{\Delta }_H\left(\overline{t}\right)=\theta \left(t\right)\overline{t}\overline{t}$$ Dualizing $`\left(\text{3.1}\right)`$ we get another extension $$F^{}\stackrel{\pi ^{}}{}H^{}\stackrel{i^{}}{}K^{}$$ and as in \[11, 2.4\], \[12, 2.11\] or \[15, 2.1\], since $`k`$ is algebraically closed, there exist units $`\overline{x}`$ and $`\overline{y}H^{}`$, such that such that $`\overline{x}^4=1_H^{}`$, $`\overline{y}^2=\overline{x}^2`$, $`e_{pq},\overline{x}^i\overline{y}^j=\delta _{ip}\delta _{jq}`$ and $`\alpha =\overline{x}\overline{y}\overline{x}\overline{y}^1F^{}=k\{e_0,e_1\}`$, where $`\left\{e_r\right\}`$ is a dual basis of $`\left\{t^r\right\}`$, $`r=0,1`$. $`\epsilon \left(\alpha \right)=\epsilon \left(\overline{x}\overline{y}\overline{x}\overline{y}^1\right)=1`$ and therefore $`\alpha =e_0+\xi e_1`$. The right action $`\rho ^{}:F^{}K^{}F^{}`$ is trivial, thus $`F^{}`$ lies in the center of $`H^{}`$. Moreover, $`\overline{x}^2=\overline{y}^2`$ also lies in the center of $`H^{}`$. Then $`\overline{x}\overline{y}\overline{x}^1\overline{y}^1`$ $`=`$ $`\overline{x}\overline{y}\overline{x}^3\overline{y}^1=\overline{x}\overline{y}\overline{x}\overline{y}^1\overline{x}^2=\alpha \overline{x}^2`$ $`\overline{x}^3`$ $`=`$ $`\overline{x}\overline{x}^2=\overline{x}\overline{y}^2=\alpha \overline{x}^2\overline{y}\overline{x}\overline{y}=\alpha \overline{x}^2\overline{y}\alpha \overline{x}^2\overline{y}\overline{x}=\alpha ^2\overline{x}^4\overline{y}^2\overline{x}=\alpha ^2\overline{x}^3`$ Thus $`\alpha ^2=1`$ and $`\xi =\pm 1`$. $`\mathrm{\Delta }_H\left(\overline{t}\right),\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r`$ $`=`$ $`\overline{t},\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r=\delta _{kr}\overline{t},\overline{x}^{i+p}\left(\alpha \overline{x}^2\right)^{jp}\overline{y}^j\overline{y}^qe_k`$ $`=`$ $`\delta _{kr}\overline{t},\overline{x}^{i+p+2jp+2jq}\overline{y}^{j+q2jq}\alpha ^{jp}e_k=\xi ^{jp}\delta _{k1}\delta _{r1}`$ On the other hand $$\mathrm{\Delta }_H\left(\overline{t}\right),\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r=\theta \left(t\right)\overline{t}\overline{t},\overline{x}^i\overline{y}^je_k\overline{x}^p\overline{y}^qe_r=\theta \left(t\right),\overline{x}^i\overline{y}^j\overline{x}^p\overline{y}^q\delta _{k1}\delta _{r1}$$ Therefore $$\theta \left(t\right)=\underset{ijpq}{}\xi ^{jp}e_{ij}e_{pq}$$ If $`\xi =1`$ then $`\overline{t}`$ is a grouplike of $`H`$, if $`\xi =1`$ then $`i^{p+q^2}e_{p,q}\overline{t}`$ is a grouplike of $`H`$. Thus $`𝐆\left(H\right)`$ has always order $`8`$. Write $`v=\sigma (t,t)=c_{i,j}e_{i,j}`$ then $`c_{0,0}=\epsilon \left(v\right)=1`$ and $`c_{i,j}0,`$ since $`v`$ is a unit and $$\mathrm{\Delta }_H\left(\overline{t}^2\right)=\mathrm{\Delta }_H\left(v\right)=\mathrm{\Delta }_K\left(c_{i,j}e_{i,j}\right)=c_{p+r+2rq+2sq,q+s}e_{p,q}e_{r,s}$$ Action by $`t`$ is a Hopf algebra map and therefore it induces a group automorphism $`f_t:GG`$ defined by $`e_{p,q},f_t\left(g\right)=te_{p,q},g`$, which has order $`2`$. Renaming generators we are down to two choices for $`f_t`$; we consider them below: #### Case D) The action is given by $`te_{i,j}=e_{i+2j,j}`$, corresponding to $`f_t\left(x\right)`$ $`=`$ $`x`$ $`f_t\left(y\right)`$ $`=`$ $`x^2y`$ Then $`X`$ and $`Y`$ are central grouplikes of $`H`$. Thus $`𝐆\left(H\right)`$ is abelian of order $`8`$ and $`H`$ was described in Section 3.1 or Section 3.2. #### Case E) The action is given by $`te_{i,j}=e_{i+j,j}`$, corresponding to $`f_t\left(x\right)`$ $`=`$ $`x^1`$ $`f_t\left(y\right)`$ $`=`$ $`xy`$ Then $`Y`$ is a central grouplike of $`H`$. Write $`v=\sigma (t,t)=c_{i,j}e_{i,j}`$ . By $`\left(\text{3.2}\right)`$ (3.25) $$c_{i,j}=c_{i+j,j}$$ On the other hand, for $`H`$ to be a bialgebra $`\mathrm{\Delta }_H\left(\overline{t}^2\right)`$ $`=`$ $`\mathrm{\Delta }_H\left(\overline{t}\right)\mathrm{\Delta }_H\left(\overline{t}\right)=\left(\theta \left(t\right)\overline{t}\overline{t}\right)\left(\theta \left(t\right)\overline{t}\overline{t}\right)`$ $`=`$ $`\left({\displaystyle \underset{pqrs}{}}\xi ^{rq}e_{pq}e_{rs}\right)\left({\displaystyle \underset{pqrs}{}}\xi ^{rq}\left(te_{pq}\right)\left(te_{rs}\right)\right)\sigma (t,t)\sigma (t,t)`$ $`=`$ $`{\displaystyle \xi ^{rq}\xi ^{\left(r+s\right)q}c_{p,q}c_{r,s}e_{p,q}e_{r,s}}={\displaystyle \xi ^{qs}c_{p,q}c_{r,s}e_{p,q}e_{r,s}}`$ Therefore (3.26) $$c_{p+r+2\left(r+s\right)q,q+s}=\xi ^{qs}c_{p,q}c_{r,s}$$ Conditions $`\left(\text{3.25}\right)`$ and $`\left(\text{3.26}\right)`$ imply that $`c_{0,1}`$ $`=`$ $`c_{1,1}`$ $`c_{2,1}`$ $`=`$ $`c_{3,1}`$ $`c_{1,0}`$ $`=`$ $`c_{3,0}`$ $`c_{1,0}c_{0,1}`$ $`=`$ $`c_{1,1}=c_{0,1}`$ $`c_{0,1}c_{1,0}`$ $`=`$ $`c_{3,1}`$ $`\xi c_{0,1}c_{0,1}`$ $`=`$ $`c_{2,0}=c_{1,0}c_{1,0}`$ Thus $`c_{1,0}=1`$ and $`c_{0,1}=i^k`$, where $`\xi =i^{2k}`$ and $`k=0,1,2,3`$. Therefore $$\sigma _k(t,t)=i^{kq}e_{p,q}=\frac{1+Y}{2}+\frac{i^k\left(1Y\right)}{2}$$ Let $`H_k`$ be the Hopf algebra with the structure described above with cocycle $`\sigma _k(t,t)`$. Define $`f:H_k`$ $``$ $`H_{k+1}\text{by}`$ $`f\left(e_{p,q}\right)`$ $`=`$ $`e_{p,q}`$ $`f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \underset{p,q}{}}i^{p+q^2}e_{p,q}\overline{t}`$ and extend it multiplicatively to $`f\left(e_{p,q}\overline{t}\right)`$. Then $`f\left(\overline{t}\right)f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle i^{p+q^2}e_{p,q}\overline{t}i^{p+q^2}e_{p,q}\overline{t}}={\displaystyle i^{p+q^2}e_{p,q}i^{p+q^2}e_{p+q,q}\overline{t}^2}`$ $`=`$ $`{\displaystyle i^{p+q^2}i^{p+q+q^2}e_{p,q}\sigma _{k+1}(t,t)}={\displaystyle i^{3q}e_{p,q}i^{\left(k+1\right)q}e_{p,q}}`$ $`=`$ $`{\displaystyle i^{\left(k+1\right)q+3q}e_{p,q}}={\displaystyle i^{kq}e_{p,q}}=\sigma _k(t,t)=f\left(\sigma _k(t,t)\right)=f\left(\overline{t}^2\right)`$ $`f\left(\overline{t}e_{p,q}\right)`$ $`=`$ $`f\left(e_{p+q,q}\overline{t}\right)=e_{p+q,q}{\displaystyle i^{p+q^2}e_{p,q}\overline{t}}={\displaystyle i^{p+q^2}e_{p,q}\overline{t}e_{p,q}}=f\left(\overline{t}\right)f\left(e_{p,q}\right)`$ $`\mathrm{\Delta }f\left(\overline{t}\right)`$ $`=`$ $`\mathrm{\Delta }\left({\displaystyle i^{p+q^2}e_{p,q}\overline{t}}\right)=\mathrm{\Delta }\left({\displaystyle i^{p+q^2}e_{p,q}}\right)\mathrm{\Delta }\left(\overline{t}\right)=`$ $`=`$ $`\left({\displaystyle i^{p_1+p_2+2q_1\left(p_2+q_2\right)+\left(q_1+q_2\right)^2}e_{p_1,q_1}e_{p_2,q_2}}\right)\left({\displaystyle i^{2\left(k+1\right)s_1r_2}e_{r_1,s_1}\overline{t}e_{r_2,s_2}\overline{t}}\right)`$ $`=`$ $`\left({\displaystyle i^{p_1+p_2+2q_1\left(p_2+q_2\right)+\left(q_1+q_2\right)^2+2\left(k+1\right)q_1p_2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}\right)`$ $`=`$ $`{\displaystyle i^{p_1+p_2+q_1^2+q_2^2}i^{2kq_1p_2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}`$ $`=`$ $`{\displaystyle i^{2kq_1p_2}e_{p_1,q_1}f\left(\overline{t}\right)e_{p_2,q_2}f\left(\overline{t}\right)}`$ $`=`$ $`\left(ff\right)\left({\displaystyle i^{2kq_1p_2}e_{p_1,q_1}\overline{t}e_{p_2,q_2}\overline{t}}\right)=\left(ff\right)\mathrm{\Delta }\left(\overline{t}\right)`$ and such an $`f`$ is a Hopf algebra isomorphism between $`H_k`$ and $`H_{k+1}`$. Thus there is exactly one Hopf algebra of this type: $`H_E=H_0`$ with a trivial cocycle and $`\xi =1`$. Then $`𝐆\left(H_E\right)=X\overline{t},XD_8`$ and $`𝐆\left(H_E^{}\right)=\chi \times \phi C_2\times C_2`$, where $`\chi (e_{p,q})=\delta _{p,2}\delta _{q,0}`$, $`\chi (\overline{t})=1`$, $`\phi (e_{p,q})=\delta _{p,0}\delta _{q,0}`$, $`\phi \left(\overline{t}\right)=1`$. There are three degree $`2`$ irreducible representations defined by $`\pi _1\left(e_{0,1}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _1\left(e_{1,1}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _1\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2\left(e_{1,0}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _2\left(e_{3,0}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _2\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _3\left(e_{2,1}\right)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 0\end{array}\right)`$ $`\pi _3\left(e_{3,1}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _3\left(\overline{t}\right)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi _2^2=\pi _2\pi _2=1+\chi +\phi +\chi \phi `$, $`\pi _1^2=\pi _3^2=\chi +\chi \phi +\pi _2`$. ### 3.5. Summary ###### Proposition 3.1. Let $`H`$ be a nontrivial semisimple Hopf algebra of dimension $`16`$. Then $`𝐆\left(H\right)`$ is abelian of order $`8`$ if and only if $`𝐆\left(H^{}\right)`$ is abelian of order $`8`$. ###### Proof. All nontrivial Hopf algebras with abelian groups of grouplikes were described in Sections 3.1 and 3.2 and their duals have abelian groups of grouplikes. ∎ ###### Proposition 3.2. There are exactly $`7`$ nonisomorphic nontrivial semisimple Hopf algebras of dimension $`16`$ with $`𝐆\left(H\right)C_4\times C_2`$. ###### Proof. All nontrivial Hopf algebras with $`𝐆\left(H\right)C_4\times C_2`$ were described in Section 3.1. There are at most $`7`$ nonisomorphic Hopf algebras with $`𝐆\left(H\right)C_4\times C_2`$, namely $`H_{a:1}`$, $`H_{a:y}`$, $`H_{b:1}`$, $`H_{b:y}`$, $`H_{b:x^2y}`$, $`H_{c:\sigma _0}`$, $`H_{c:\sigma _1}`$. Assume $`f`$ is a Hopf algebra isomorphism between Hopf algebras $`H_1`$ and $`H_2`$ with $`𝐆\left(H_1\right)𝐆\left(H_2\right)C_4\times C_2`$. Then we get a group isomorphism $$f_{𝐆\left(H_1\right)}:𝐆\left(H_1\right)𝐆\left(H_2\right)$$ Write $`𝐆\left(H_1\right)𝐆\left(H_2\right)=x\times y`$, where $`\left|x\right|=4`$ and $`\left|y\right|=2`$. Then the dual basis of $`k𝐆\left(H_1\right)k𝐆\left(H_2\right)`$ is given by $$e_{pq}=1/8\left(1+i^px+i^{2p}x^2+i^{3p}x^3\right)\left(1+\left(1\right)^qy\right),\text{ }p=0,1,2,3;q=0,1$$ Write $`f\left(e_{p,q}\right)`$ $`=`$ $`e_{\alpha _1(p,q),\alpha _2(p,q)}`$ $`f^1\left(e_{p,q}\right)`$ $`=`$ $`e_{\beta _1(p,q),\beta _2(p,q)}`$ where $`\alpha _1(p,q),\beta _1(p,q)\{0,1,2,3\}`$ and $`\alpha _2(p,q),\beta _2(p,q)\{0,1\}`$. Write $`\left\{e_{pq}\overline{t^r}\right\}_{p=0,1,2,3;q=0,1;r=0,1}`$ and $`\left\{e_{pq}\overline{T^r}\right\}_{p=0,1,2,3;q=0,1;r=0,1}`$ for the bases of $`H_1`$ and $`H_2`$, respectively. Write $$f\left(\overline{t}\right)=\underset{p,q,r}{}\lambda _{p,q,r}e_{pq}\overline{T^r}$$ Then $`\mathrm{\Delta }f\left(\overline{t}\right)`$ $`=`$ $`\mathrm{\Delta }\left({\displaystyle \underset{p,q,r}{}}\lambda _{p,q,r}e_{pq}\overline{T^r}\right)={\displaystyle \underset{p,q}{}}\lambda _{p,q,0}\mathrm{\Delta }\left(e_{pq}\right)+{\displaystyle \underset{p,q}{}}\lambda _{p,q,1}\mathrm{\Delta }\left(e_{pq}\right)\mathrm{\Delta }\left(\overline{T}\right)`$ $`=`$ $`{\displaystyle \lambda _{p_1+p_2,q_1+q_2,0}e_{p_1q_1}e_{p_2q_2}}`$ $`+\left({\displaystyle \lambda _{p_1+p_2,q_1+q_2,1}e_{p_1q_1}e_{p_2q_2}}\right)\left({\displaystyle \left(1\right)^{bc}e_{ab}\overline{T}e_{cd}\overline{T}}\right)`$ $`=`$ $`{\displaystyle \lambda _{p_1+p_2,q_1+q_2,0}e_{p_1q_1}e_{p_2q_2}}+{\displaystyle \left(1\right)^{p_2q_1}\lambda _{p_1+p_2,q_1+q_2,1}e_{p_1q_1}\overline{T}e_{p_2q_2}\overline{T}}`$ $`\left(ff\right)\mathrm{\Delta }\left(\overline{t}\right)`$ $`=`$ $`\left(ff\right)\left({\displaystyle \left(1\right)^{p_2q_1}e_{p_1q_1}\overline{T}e_{p_2q_2}\overline{T}}\right)`$ $`=`$ $`{\displaystyle \left(1\right)^{p_2q_1}f\left(e_{p_1q_1}\right)\underset{p,q,r}{}\lambda _{p,q,r}e_{pq}\overline{T^r}f\left(e_{p_2q_2}\right)\underset{p,q,r}{}\lambda _{p,q,r}e_{pq}\overline{T^r}}`$ $`=`$ $`{\displaystyle \left(1\right)^{\beta _1(p_2,q_2)\beta _2(p_1,q_1)}\lambda _{p_1,q_1,r_1}\lambda _{p_2,q_2,r_2}e_{p_1q_1}\overline{T^{r_1}}e_{p_2q_2}\overline{T^{r_2}}}`$ Since $`f`$ is a coalgebra map, $$\mathrm{\Delta }f\left(\overline{t}\right)=\left(ff\right)\mathrm{\Delta }\left(\overline{t}\right)$$ and therefore $`\lambda _{p_1,q_1,0}\lambda _{p_2,q_2,1}=0`$ for all $`p_1,p_2\{0,1,2,3\}`$, $`q_1,q_2\{0,1\}`$. Thus either $`\lambda _{p,q,0}=0`$ for all $`p\{0,1,2,3\}`$, $`q\{0,1\}`$ or $`\lambda _{p,q,1}=0`$ for all $`p\{0,1,2,3\}`$, $`q\{0,1\}`$. In the latter case $`f\left(\overline{t}\right)=\lambda _{p,q,0}e_{pq}k𝐆\left(H_2\right)`$, which contradicts the bijectivity of $`f`$. Therefore $`\lambda _{p,q,0}=0`$ for all $`p\{0,1,2,3\}`$, $`q\{0,1\}`$. Write $`\lambda _{p,q}=\lambda _{p,q,1}`$. Then $$f\left(\overline{t}\right)=\underset{p,q}{}\lambda _{p,q}e_{pq}\overline{T}$$ and so, applying $`\epsilon `$, also $$\lambda _{0,0}=\epsilon \left(\overline{t}\right)=1$$ Moreover, since $`\left(1\right)^{p_2q_1}\lambda _{p_1+p_2,q_1+q_2}e_{p_1q_1}\overline{T}e_{p_2q_2}\overline{T}`$ $`=\left(1\right)^{\beta _1(p_2,q_2)\beta _2(p_1,q_1)}\lambda _{p_1,q_1}\lambda _{p_2,q_2}e_{p_1q_1}\overline{T}e_{p_2q_2}\overline{T}`$ we get (3.27) $$\lambda _{p_1+p_2,q_1+q_2}=\left(1\right)^{p_2q_1}\left(1\right)^{\beta _1(p_2,q_2)\beta _2(p_1,q_1)}\lambda _{p_1,q_1}\lambda _{p_2,q_2}$$ for any $`p_1,p_2\{0,1,2,3\}`$, $`q_1,q_2\{0,1\}`$. Let $`uk𝐆\left(H_1\right)`$. Then $`f\left(t_1u\right)f\left(\overline{t}\right)`$ $`=`$ $`f\left(\left(t_1u\right)\overline{t}\right)=f\left(\overline{t}u\right)=f\left(\overline{t}\right)f\left(u\right)={\displaystyle \lambda _{p,q}e_{p,q}\overline{T}f\left(u\right)}`$ $`=`$ $`\left(t_2f\left(u\right)\right){\displaystyle \lambda _{p,q}e_{p,q}\overline{T}}=\left(t_2f\left(u\right)\right)f\left(\overline{t}\right)`$ Thus, since $`\overline{t}`$ is a unit ($`\overline{t}^2=\sigma (t,t)`$ is a unit), (3.28) $$f\left(t_1u\right)=t_2f\left(u\right)$$ Let’s show that Hopf algebras from types $`H_a`$, $`H_b`$ and $`H_c`$ cannot be isomorphic to each other. $`K_0\left(H_c\right)K_0\left(H_a\right)`$ or $`K_0\left(H_b\right)`$, thus $`H_cH_a`$ or $`H_b`$. If $`f:H_aH_b`$ then by formula (3.28) $$f\left(x\right)f\left(y\right)=f\left(xy\right)=f\left(t_1x\right)=t_2f\left(x\right)=f\left(x\right)^1$$ and therefore $`f\left(y\right)=f\left(x^2\right)`$, which is impossible if $`f`$ is an isomorphism. $`H_{b:1}H_{b:y}`$ or $`H_{b:x^2y}`$ and $`H_{c:\sigma _0}H_{c:\sigma _1}`$ since their duals have nonisomorphic groups of grouplikes. Thus there are at least $`5`$ nonisomorphic Hopf algebras with $`𝐆\left(H\right)C_4\times C_2`$, namely $`H_{a:1}`$, $`H_{b:1}`$, $`H_{b:y}`$, $`H_{c:\sigma _0}`$ and $`H_{c:\sigma _1}`$. Now we prove that $`H_{a:1}H_{a:y}`$ and $`H_{b:y}H_{b:x^2y}`$. If $`f`$ is a Hopf algebra isomorphism as before, then since $`f_{𝐆\left(H_1\right)}`$ is a group isomorphism $`f\left(x\right)\{x,x^1,xy,x^1y\}`$, $`f\left(y\right)\{y,x^2y\}`$ and $`f\left(x^2\right)=x^2`$. If $`f\left(x\right)=x^{2k+1}y^l`$ and $`f\left(y\right)=y`$, where $`k,l=0,1`$ then $$f^1\left(e_{p,q}\right)=f\left(e_{p,q}\right)=e_{\left(2k+1\right)p+2lq,q}$$ and by formula (3.27) $$\lambda _{p_1+p_2,q_1+q_2}=\left(1\right)^{p_2q_1}\left(1\right)^{\left(\left(2k+1\right)p_2+2lq_2\right)q_1}\lambda _{p_1,q_1}\lambda _{p_2,q_2}=\lambda _{p_1,q_1}\lambda _{p_2,q_2}$$ If $`f\left(x\right)=x^{2k+1}y^l`$ and $`f\left(y\right)=x^2y`$, where $`k,l=0,1`$ then $`f\left(e_{p,q}\right)`$ $`=`$ $`e_{\left(2k+2l+1\right)p+2lq,p+q}`$ $`f^1\left(e_{p,q}\right)`$ $`=`$ $`e_{\left(2k+1\right)p+2lq,p+q}`$ and by formula (3.27) $$\lambda _{p_1+p_2,q_1+q_2}=\left(1\right)^{p_2q_1}\left(1\right)^{\left(\left(2k+1\right)p_2+2lq_2\right)\left(q_1+p_1\right)}\lambda _{p_1,q_1}\lambda _{p_2,q_2}=\left(1\right)^{p_1p_2}\lambda _{p_1,q_1}\lambda _{p_2,q_2}$$ Now assume $`f`$ is a Hopf algebra isomorphism $$f:H_{a:1}H_{a:y}$$ If $`f\left(y\right)=x^2y`$ then by formula (3.28) $$f\left(x\right)y=t_2f\left(x\right)=f\left(t_1x\right)=f\left(xy\right)=f\left(x\right)f\left(y\right)=f\left(x\right)x^2y$$ that is $`x^2=1`$, which contradicts the fact that $`\left|x\right|=4`$. Thus $`f\left(y\right)=y`$ and therefore $$\lambda _{p_1+p_2,q_1+q_2}=\lambda _{p_1,q_1}\lambda _{p_2,q_2}$$ and thus $$\lambda _{1,0}^4=\lambda _{2,0}^2=\lambda _{0,1}^2=\lambda _{0,0}=1$$ Then $`f\left(\overline{t}\right)f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \lambda _{p,q}e_{pq}\overline{T}\lambda _{r,s}e_{rs}\overline{T}}`$ $`=`$ $`{\displaystyle \lambda _{p,q}e_{pq}\lambda _{r,s}e_{r+2s,s}\overline{T}^2}={\displaystyle \lambda _{p,q}\lambda _{p+2q,q}e_{pq}\sigma _{H_{a:y}}(t,t)}`$ $`=`$ $`{\displaystyle \lambda _{2\left(p+q\right),0}e_{pq}\sigma _{H_{a:y}}(t,t)}={\displaystyle \lambda _{2,0}^{p+q}e_{pq}\sigma _{H_{a:y}}(t,t)}`$ If $`\lambda _{2,0}=1`$, $`f\left(\overline{t}\right)f\left(\overline{t}\right)=\sigma _{H_{a:y}}(t,t)=yf\left(\overline{t}^2\right)=1`$. If $`\lambda _{2,0}=1`$, $`f\left(\overline{t}\right)f\left(\overline{t}\right)=x^2y\sigma _{H_{a:y}}(t,t)=x^2f\left(\overline{t}^2\right)=1`$. Therefore, there is no Hopf algebra isomorphism between $`H_{a:1}`$ and $`H_{a:y}`$. Now assume $`f`$ is a Hopf algebra isomorphism $$f:H_{b:y}H_{b:x^2y}$$ Then $`f\left(\overline{t}\right)f\left(\overline{t}\right)`$ $`=`$ $`{\displaystyle \lambda _{p,q}e_{pq}\overline{T}\lambda _{r,s}e_{rs}\overline{T}}`$ $`=`$ $`{\displaystyle \lambda _{p,q}e_{pq}\lambda _{r,s}e_{r,s}\overline{T}^2}={\displaystyle \lambda _{p,q}\lambda _{p,q}e_{pq}\sigma _{H_{b:x^2y}}(t,t)}`$ $`f\left(y\right)=y`$ is not possible, since then we have $$\lambda _{p,q}\lambda _{p,q}=\lambda _{0,0}=1$$ and thus $$f\left(\overline{t}\right)f\left(\overline{t}\right)=\left(e_{pq}\right)\sigma _{H_{b:x^2y}}(t,t)=\sigma _{H_{b:x^2y}}(t,t)=x^2yy=f\left(y\right)=f\left(\overline{t}^2\right).$$ $`f\left(y\right)=x^2y`$ is not possible, since then we get $$\lambda _{p_1+p_2,q_1+q_2}=\left(1\right)^{p_1p_2}\lambda _{p_1,q_1}\lambda _{p_2,q_2}$$ so $$\lambda _{p,q}\lambda _{p,q}=\left(1\right)^{p^2}\lambda _{0,0}=\left(1\right)^{p^2}$$ and $$f\left(\overline{t}\right)f\left(\overline{t}\right)=\left(\left(1\right)^{p^2}e_{pq}\right)\sigma _{H_{b:x^2y}}(t,t)=x^2x^2y=y=f\left(x^2y\right)f\left(y\right)=f\left(\overline{t}^2\right).$$ Therefore, there is no Hopf algebra isomorphism between $`H_{b:y}`$ and $`H_{b:x^2y}`$. Thus there are exactly $`7`$ nonisomorphic Hopf algebras with $`𝐆\left(H\right)C_4\times C_2`$, namely $`H_{a:1}`$, $`H_{a:y}`$, $`H_{b:1}`$, $`H_{b:y}`$, $`H_{b:x^2y}`$, $`H_{c:\sigma _0}`$ and $`H_{c:\sigma _1}`$. ∎ ###### Proposition 3.3. There are at least $`2`$ and at most $`4`$ nonisomorphic nontrivial semisimple Hopf algebras of dimension $`16`$ with $`𝐆\left(H\right)C_2\times C_2\times C_2`$. ###### Proof. All nontrivial Hopf algebras with $`𝐆\left(H\right)C_2\times C_2\times C_2`$ were described in Section 3.2. There are at most $`4`$ nonisomorphic Hopf algebras with $`𝐆\left(H\right)C_2\times C_2\times C_2`$, namely $`H_{d:1,1}`$, $`H_{d:1,1}`$, $`H_{d:1,1}`$ and $`H_{d:1,1}`$. At least two of them are not isomorphic, since $`𝐆\left(H_{d:1,1}^{}\right)𝐆\left(H_{d:1,1}^{}\right)`$. ∎ ###### Proposition 3.4. There are exactly $`2`$ nonisomorphic nontrivial semisimple Hopf algebras of dimension $`16`$ with a commutative subHopfalgebra of dimension $`8`$ and nonabelian $`𝐆\left(H^{}\right)`$. In this case $`𝐆\left(H^{}\right)D_8`$, $`𝐆\left(H\right)C_2\times C_2`$ and $`H^{}`$ also has a commutative subHopfalgebra of dimension $`8`$. ###### Proof. All nontrivial Hopf algebras with a commutative subHopfalgebra of dimension $`8`$ and nonabelian $`𝐆\left(H^{}\right)`$ were described in Section 3.3, Case B). There are at most $`2`$ of them, namely $`H_{B:1}`$ and $`H_{B:X}`$. Let’s compute all the possible $`8`$-dimensional Hopf quotients of $`H_B`$. There is a one-to-one correspondence between hereditary subrings of $`K_0(H)`$ and Hopf quotients of $`H`$ (see \[22, Theorem 6\] or \[24, Proposition 3.11\]). Thus $`H_B`$ has $`3`$ quotients of dimension $`8`$ corresponding to the hereditary subrings $`R_1=\{a1+b\phi +c\chi ^2+d\chi ^2\phi +e\pi _1K_0(H):a,b,c,d,e\}`$, $`R_2=\{a1+b\phi +c\chi ^2+d\chi ^2\phi +e\pi _2K_0(H):a,b,c,d,e\}`$ and $`R_3=\{_{i=0}^4_{j=0}^2a_{i,j}\chi ^i\phi ^jK_0(H):a_{i,j}\}`$. They are obtained by factoring modulo normal ideals $`\left(Y1\right)H`$, $`\left(X1\right)H`$ and $`\left(XY1\right)H`$, where $`X`$, $`Y`$ and $`XY`$ are central grouplikes of $`H`$. It is easy to see that $`H/\left(Y1\right)H`$ is cocommutative (in fact, $`H_{B:1}/\left(Y1\right)H_{B:1}kD_8`$ and $`H_{B:X}/\left(Y1\right)H_{B:X}kQ_8`$), $`H/\left(X1\right)H`$ is commutative (therefore $`H/\left(X1\right)H\left(kD_8\right)^{}`$ since $`𝐆\left(H^{}\right)D_8`$) and $`H/\left(XY1\right)H`$ is neither commutative nor cocommutative (therefore $`H/\left(X1\right)HH_8`$). Therefore we see that $`H_{B:1}H_{B:X}`$ since they have different sets of quotients. Both $`H_{B:1}`$ and $`H_{B:X}`$ have cocommutative Hopf quotients of dimension $`8`$, $`kD_8`$ and $`kQ_8`$ respectively. Thus their duals were described in Section 3.4. In particular, $`H_{B:1}H_{C:1}^{}`$, $`H_{B:X}H_E^{}`$ and $`𝐆\left(H_{C:1}^{}\right)𝐆\left(H_E^{}\right)C_2\times C_2`$. ∎ ## 4. Nonabelian groups of order 16. There are nine nonabelian groups of order $`16`$ (see ). The first four of them are of exponent $`8`$, the last five of exponent $`4`$ (we denote the quaternion group of order $`8`$ by $`Q_8`$ and the quasiquaternion group of order $`16`$ by $`Q_{16}`$): 1. $`G_1=a,b:a^8=b^2=1,ba=a^5b`$ $`𝐆\left(\left(kG_1\right)^{}\right)=\chi \times \phi C_4\times C_2`$, where $`\chi (a)=i`$, $`\chi (b)=1`$, $`\phi \left(a\right)=1`$, $`\phi \left(b\right)=1`$. Degree $`2`$ irreducible representations of $`G_1`$ are defined by $`\pi _1(a)=\left(\begin{array}{cc}\hfill \omega & \hfill 0\\ \hfill 0& \hfill \omega \end{array}\right)`$ $`\pi _1(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2(a)=\left(\begin{array}{cc}\hfill \omega ^3& \hfill 0\\ \hfill 0& \hfill \omega ^3\end{array}\right)`$ $`\pi _2(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ where $`\omega `$ is a primitive $`8^{th}`$-root of unity and $`\pi _1^2=\chi +\chi ^3+\chi \phi +\chi ^3\phi =\pi _2^2.`$ 2. $`G_2=a,b:a^8=b^2=1,ba=a^3b`$ $`𝐆\left(\left(kG_2\right)^{}\right)=\chi \times \phi C_2\times C_2`$, where $`\chi (a)=1`$, $`\chi (b)=1`$, $`\phi \left(a\right)=1`$, $`\phi \left(b\right)=1`$. Degree $`2`$ irreducible representations of $`G_2`$ are defined by $`\pi _1(a)=\left(\begin{array}{cc}\hfill \omega & \hfill 0\\ \hfill 0& \hfill \omega ^3\end{array}\right)`$ $`\pi _2(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _3(a)=\left(\begin{array}{cc}\hfill \omega ^5& \hfill 0\\ \hfill 0& \hfill \omega ^7\end{array}\right)`$ $`\pi _1(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _3(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ where $`\omega `$ is a primitive $`8^{th}`$-root of unity, and representations satisfy the properties $`\pi _1^2=\chi +\chi \phi +\pi _2=\pi _3^2`$ $`\chi \pi _1=\pi _3`$ $`\chi \pi _3=\pi _1`$ $`\pi _2^2=1+\chi +\phi +\chi \phi `$ $`\phi \pi _1=\pi _1`$ $`\phi \pi _3=\pi _3`$ 3. $`G_3=a,b:a^8=b^2=1,ba=a^1b=D_{16}`$, the dihedral group $`𝐆\left(\left(kG_3\right)^{}\right)=\chi \times \phi C_2\times C_2`$, where $`\chi (a)=1`$, $`\chi (b)=1`$, $`\phi \left(a\right)=1`$, $`\phi \left(b\right)=1`$. Degree $`2`$ irreducible representations of $`G_3`$ are defined by $`\pi _1(a)=\left(\begin{array}{cc}\hfill \omega & \hfill 0\\ \hfill 0& \hfill \omega ^7\end{array}\right)`$ $`\pi _2(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _3(a)=\left(\begin{array}{cc}\hfill \omega ^3& \hfill 0\\ \hfill 0& \hfill \omega ^5\end{array}\right)`$ $`\pi _1(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _3(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ where $`\omega `$ is a primitive $`8^{th}`$-root of unity, and representations satisfy the properties $`\pi _1^2=1+\phi +\pi _2=\pi _3^2`$ $`\chi \pi _1=\pi _3`$ $`\chi \pi _3=\pi _1`$ $`\pi _2^2=1+\chi +\phi +\chi \phi `$ $`\phi \pi _1=\pi _1`$ $`\phi \pi _3=\pi _3`$ 4. $`G_4=a,b:a^8=1,b^2=a^4,ba=a^1b=Q_{16}`$, the quasiquaternion group $`𝐆\left(\left(kG_4\right)^{}\right)=\chi \times \phi C_2\times C_2`$, where $`\chi (a)=1`$, $`\chi (b)=1`$, $`\phi \left(a\right)=1`$, $`\phi \left(b\right)=1`$. Degree $`2`$ irreducible representations of $`G_4`$ are defined by $`\pi _1(a)=\left(\begin{array}{cc}\hfill \omega & \hfill 0\\ \hfill 0& \hfill \omega ^7\end{array}\right)`$ $`\pi _2(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _3(a)=\left(\begin{array}{cc}\hfill \omega ^3& \hfill 0\\ \hfill 0& \hfill \omega ^5\end{array}\right)`$ $`\pi _1(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _3(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ where $`\omega `$ is a primitive $`8^{th}`$-root of unity, and representations satisfy the properties $`\pi _1^2=1+\phi +\pi _2=\pi _3^2`$ $`\chi \pi _1=\pi _3`$ $`\chi \pi _3=\pi _1`$ $`\pi _2^2=1+\chi +\phi +\chi \phi `$ $`\phi \pi _1=\pi _1`$ $`\phi \pi _3=\pi _3`$ 5. $`G_5=a,b:a^4=b^4=1,ba=a^1b`$ $`𝐆\left(\left(kG_5\right)^{}\right)=\chi \times \phi C_4\times C_2`$, where $`\chi (a)=1`$, $`\chi (b)=i`$, $`\phi \left(a\right)=1`$, $`\phi \left(b\right)=1`$. Degree $`2`$ irreducible representations of $`G_5`$ are defined by $`\pi _1(a)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _1(b)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _2(a)=\left(\begin{array}{cc}\hfill 0& \hfill i\\ \hfill i& \hfill 0\end{array}\right)`$ $`\pi _2(b)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ with the property $`\pi _1^2=1+\chi ^2+\phi +\chi ^2\phi =\pi _2^2.`$ 6. $`G_6=a,b,c:a^4=b^2=c^2=1,bab=ac`$ $`𝐆\left(\left(kG_6\right)^{}\right)=\chi \times \phi C_4\times C_2`$, where $`\chi (a)=i`$, $`\chi (b)=\chi (c)=1`$, $`\phi \left(a\right)=\phi \left(c\right)=1`$, $`\phi \left(b\right)=1`$. Degree $`2`$ irreducible representations of $`G_6`$ are defined by $`\pi _1(a)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _1(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _1(c)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _2(a)=\left(\begin{array}{cc}\hfill 0& \hfill i\\ \hfill i& \hfill 0\end{array}\right)`$ $`\pi _2(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2(c)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ with the property $`\pi _1^2=1+\chi ^2+\phi +\chi ^2\phi =\pi _2^2.`$ 7. $`G_7=a,b,c:a^4=b^2=c^2=1,cbc=a^2b`$ $`𝐆\left(\left(kG_7\right)^{}\right)=\chi \times \phi \times \psi C_2\times C_2\times C_2`$, where $`\chi (a)=1`$, $`\chi (b)=\chi (c)=1`$, $`\phi \left(a\right)=\phi \left(b\right)=1`$, $`\phi \left(c\right)=1`$, $`\psi \left(a\right)=\psi \left(b\right)=1`$, $`\psi \left(c\right)=1`$. Degree $`2`$ irreducible representations of $`G_7`$ are defined by $`\pi _1(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _1(b)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _1(c)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _2(b)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _2(c)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ with the property $`\pi _1^2=\chi +\chi \phi +\chi \psi +\chi \phi \psi =\pi _2^2.`$ 8. $`G_8=a,b,c:a^4=b^2=c^2=1,ba=a^1b=D_8\times C_2`$ $`𝐆\left(\left(kG_8\right)^{}\right)=\chi \times \phi \times \psi C_2\times C_2\times C_2`$, where $`\chi (a)=\chi (b)=1`$, $`\chi (c)=1`$, $`\phi \left(a\right)=1`$, $`\phi \left(b\right)=\phi \left(c\right)=1`$, $`\psi \left(a\right)=\psi \left(c\right)=1`$, $`\psi \left(b\right)=1`$. Degree $`2`$ irreducible representations of $`G_8`$ are defined by $`\pi _1(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _1(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _1(c)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _2(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _2(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2(c)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ with the property $`\pi _1^2=1+\phi +\psi +\phi \psi =\pi _2^2`$. 9. $`G_9=a,b,c:a^4=c^2=1,b^2=a^2,ba=a^1b=Q_8\times C_2`$ $`𝐆\left(\left(kG_9\right)^{}\right)=\chi \times \phi \times \psi C_2\times C_2\times C_2`$, where $`\chi (a)=\chi (b)=1`$, $`\chi (c)=1`$, $`\phi \left(a\right)=1`$, $`\phi \left(b\right)=\phi \left(c\right)=1`$, $`\psi \left(a\right)=\psi \left(c\right)=1`$, $`\psi \left(b\right)=1`$. Degree $`2`$ irreducible representations of $`G_9`$ are defined by $`\pi _1(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _1(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _1(c)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _2(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _2(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2(c)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ with the property $`\pi _1^2=1+\phi +\psi +\phi \psi =\pi _2^2`$. ## 5. Computations in $`K_0(H)`$ in the case of $`\left|𝐆\left(H^{}\right)\right|=8`$. In this case we have $`8`$ one-dimensional irreducible representations $`\chi _1=1_{K_0(H)},\mathrm{},`$ $`\chi _8𝐆\left(H^{}\right)`$ and two 2-dimensional ones $`\pi _1`$ and $`\pi _2`$. Then, since $`\chi _i\chi _i^1=1_{K_0(H)}`$, $`\chi _i^{}=\chi _i^1`$. $`\mathrm{deg}\left(\chi _i\pi _k\right)=2`$ thus there are two possibilities: 1. $`\chi _i\pi _k=\chi _j+\chi _l`$ 2. $`\chi _i\pi _k=\pi _l`$ Case i) cannot happen, since otherwise $`\pi _k=\chi _i^1\chi _j+\chi _i^1\chi _l`$ is not irreducible. Thus $`\chi _i\pi _k=\pi _l`$. Then it is impossible to have $`\chi _i\pi _k=\pi _k`$ and $`\chi _i\pi _l=\pi _k`$ for $`kl`$, since otherwise $`\pi _l=\chi _i^8\pi _l=\chi _i^7\pi _k=\pi _k`$. Thus $`\chi _i`$ either fixes both $`\pi _1`$ and $`\pi _2`$ or interchanges them. It is easy to check that either all $`\chi _i`$ fix $`\pi _k`$, $`k=1,2`$, or half of $`\chi _i`$ fixes $`\pi _k`$ and half of $`\chi _i`$ interchanges them. Suppose $`\chi _i\pi _k=\pi _k`$, for $`i=1,\mathrm{},8`$ and $`k=1,2`$. Then $$1=m(\pi _k,\chi _i\pi _k)=m(\chi _i^{},\pi _k\pi _k^{})=m(\chi _i^1,\pi _k\pi _k^{})$$ Thus $$\pi _k\pi _k^{}=\underset{i=1}{\overset{8}{}}\chi _i^1=\underset{i=1}{\overset{8}{}}\chi _i$$ but $`\mathrm{deg}\left(\pi _k\pi _k^{}\right)=4`$ and $`\mathrm{deg}\left(_{i=1}^8\chi _i\right)=8`$. Thus all $`\chi _i`$ cannot fix $`\pi _k`$. Now let half of $`\chi _i`$ fix $`\pi _k`$ and half of $`\chi _i`$ interchange them, say $`\chi _i\pi _k`$ $`=`$ $`\pi _k\text{ for }i\text{ odd}`$ $`\chi _i\pi _k`$ $`=`$ $`\pi _l\text{ for }i\text{ even}`$ if $`kl`$. It is clear that $`\chi _i`$ and $`\chi _i^{}=\chi _i^1`$ fix or interchange $`\pi _k`$ simultaneously. Then for $`kl`$ $`1`$ $`=`$ $`m(\pi _k,\chi _i\pi _k)=m(\chi _i^{},\pi _k\pi _k^{})=m(\chi _i^1,\pi _k\pi _k^{})\text{ for }i\text{ odd}`$ $`1`$ $`=`$ $`m(\pi _l,\chi _i\pi _k)=m(\chi _i^{},\pi _k\pi _l^{})=m(\chi _i^1,\pi _k\pi _l^{})\text{ for }i\text{ even}`$ and therefore $`\pi _k\pi _k^{}`$ $`=`$ $`\chi _1+\chi _3+\chi _5+\chi _7`$ $`\pi _k\pi _l^{}`$ $`=`$ $`\chi _2+\chi _4+\chi _6+\chi _8`$ There are two possibilities for the involution: either $`\pi _1^{}=\pi _1`$ and $`\pi _2^{}=\pi _2`$, or $`\pi _1^{}=\pi _2`$ and $`\pi _2^{}=\pi _1`$. It is easy to check that when $`𝐆\left(H^{}\right)`$ is isomorphic to $`C_2\times C_2\times C_2`$ or $`Q_8`$ it does not matter which generators we choose to fix $`\pi _k`$ and which to interchange them. In the case of $`𝐆\left(H^{}\right)D_8`$ or $`C_4\times C_2`$ it matters and should give us two more nonisomorphic structures for $`K_0(H)`$ for each of them, but due to the results of the Section 3 we can see that in the case of $`𝐆\left(H^{}\right)C_4\times C_2`$, $`\pi _k`$ can be fixed only by elements of order $`1`$ or $`2`$ (since $`\left(\pi _k\right)^2`$ is either the sum of all elements of order $`1`$ or $`2,`$ or the sum of all elements of order $`4`$). Now assume that $`𝐆\left(H^{}\right)=\chi ,\phi :\chi ^4=1,\phi ^2=1,\phi \chi =\chi ^1\phi D_8`$ or $`𝐆\left(H^{}\right)=\chi ,\phi :\chi ^4=1,\phi ^2=\chi ^2,\phi \chi =\chi ^1\phi Q_8`$. Then $`\chi ^2`$ is the only nontrivial central element of $`𝐆\left(H^{}\right)`$. Since by \[13, Theorem1\] $`H^{}`$ has a nontrivial central grouplike, this grouplike should be equal to $`\chi ^2`$. Since $`\chi ^2`$ is a central grouplike of order $`2`$, which fixes all $`\pi _k`$, by Proposition 2.2 $`H`$ has a commutative subHopfalgebra of dimension $`8`$. Therefore, it should have the same $`K_0`$-ring as one of the Hopf algebras described in Section 3. Thus the only possible $`K_0`$-ring structure corresponds to $`𝐆\left(H^{}\right)D_8`$ with $`\left(\pi _k\right)^2=1+\chi ^2+\phi +\chi ^2\phi `$ and $`\pi _i^{}=\pi _i`$. Now let’s list all the possible ring structures of $`K_0(H)`$. 5.1. $`𝐆\left(H^{}\right)=\chi \times \phi \times \psi C_2\times C_2\times C_2`$ where $`\pi _1^{}=\pi _1`$ and $`\pi _2^{}=\pi _2`$ and $`\chi \pi _1=\pi _2=\pi _1\chi `$ $`\psi \pi _1=\pi _1=\pi _1\psi `$ $`\chi \pi _2=\pi _1=\pi _2\chi `$ $`\psi \pi _2=\pi _2=\pi _2\psi `$ $`\phi \pi _1=\pi _1=\pi _1\phi `$ $`\pi _1^2=1+\phi +\psi +\phi \psi =\pi _2^2`$ $`\phi \pi _2=\pi _2=\pi _2\phi `$ $`\pi _1\pi _2=\chi +\chi \phi +\chi \psi +\chi \phi \psi =\pi _2\pi _1`$ Examples: $`H_{b:1}`$, $`H_{d:1,1}`$, $`H_{d:1,1}`$, $`k\left(D_8\times C_2\right)`$, $`k\left(Q_8\times C_2\right)`$, $`H_8kC_2`$. 5.2. $`𝐆\left(H^{}\right)=\chi \times \phi \times \psi C_2\times C_2\times C_2`$ where $`\pi _1^{}=\pi _2`$ and $`\pi _2^{}=\pi _1`$ and $`\chi \pi _1=\pi _2=\pi _1\chi `$ $`\psi \pi _1=\pi _1=\pi _1\psi `$ $`\chi \pi _2=\pi _1=\pi _2\chi `$ $`\psi \pi _2=\pi _2=\pi _2\psi `$ $`\phi \pi _1=\pi _1=\pi _1\phi `$ $`\pi _1^2=\chi +\chi \phi +\chi \psi +\chi \phi \psi =\pi _2^2`$ $`\phi \pi _2=\pi _2=\pi _2\phi `$ $`\pi _1\pi _2=1+\phi +\psi +\phi \psi =\pi _2\pi _1`$ Examples: $`H_{c:\sigma _1}`$ and $`kG_7`$, where $`G_7=a,b,c:a^4=b^2=c^2=1,cbc=a^2b`$. 5.3. $`𝐆\left(H^{}\right)=\chi \times \phi C_4\times C_2`$ where $`\pi _1^{}=\pi _1`$ and $`\pi _2^{}=\pi _2`$ and $`\chi \pi _1=\pi _2=\pi _1\chi `$ $`\phi \pi _1=\pi _1=\pi _1\phi `$ $`\pi _1^2=1+\chi ^2+\phi +\chi ^2\phi =\pi _2^2`$ $`\chi \pi _2=\pi _1=\pi _2\chi `$ $`\phi \pi _2=\pi _2=\pi _2\phi `$ $`\pi _1\pi _2=\chi +\chi ^3+\chi \phi +\chi ^3\phi =\pi _2\pi _1`$ Examples: $`H_{a:1}`$, $`H_{a:y}`$, $`H_{b:y}`$, $`H_{b:x^2y}`$, $`H_{d:1,1}`$, $`H_{d:1,1}`$, $`kG_5`$ and $`kG_6`$, where $`G_5=a,b:a^4=b^4=1,b^1ab=a^1`$ and $`G_6=a,b,c:a^4=b^2=c^2=1,bab=ac`$. 5.4. $`𝐆\left(H^{}\right)=\chi \times \phi C_4\times C_2`$ where $`\pi _1^{}=\pi _2`$ and $`\pi _2^{}=\pi _1`$ and $`\chi \pi _1=\pi _2=\pi _1\chi `$ $`\phi \pi _1=\pi _1=\pi _1\phi `$ $`\pi _1^2=1+\chi ^2+\phi +\chi ^2\phi =\pi _2^2`$ $`\chi \pi _2=\pi _1=\pi _2\chi `$ $`\phi \pi _2=\pi _2=\pi _2\phi `$ $`\pi _1\pi _2=\chi +\chi ^3+\chi \phi +\chi ^3\phi =\pi _2\pi _1`$ Examples: $`H_{c:\sigma _0}`$, $`kG_1`$, where $`G_1=a,b:a^8=b^2=1,bab=a^5`$. 5.5. $`𝐆\left(H^{}\right)=\chi ,\phi :\chi ^4=1,\phi ^2=1,\phi \chi =\chi ^1\phi D_8`$, where $`\pi _1^{}=\pi _1`$ and $`\pi _2^{}=\pi _2`$ and $`\chi \pi _1=\pi _2=\pi _1\chi `$ $`\phi \pi _1=\pi _1=\pi _1\phi `$ $`\pi _1^2=1+\chi ^2+\phi +\chi ^2\phi =\pi _2^2`$ $`\chi \pi _2=\pi _1=\pi _2\chi `$ $`\phi \pi _2=\pi _2=\pi _2\phi `$ $`\pi _1\pi _2=\chi +\chi ^3+\chi \phi +\chi ^3\phi =\pi _2\pi _1`$ Examples: $`H_{B:1}`$, $`H_{B:X}`$ and $`kQ_8\mathrm{\#}^\alpha kC_2`$. ###### Remark 5.1. Noncommutative $`K_0(H)`$ should have the structure 5.5. ## 6. Computations in $`K_0(H)`$ in the case of $`\left|𝐆\left(H^{}\right)\right|=4`$. In this case by Theorem 1.1 $`𝐆\left(H^{}\right)C_2\times C_2`$ and we have $`4`$ one-dimensional irreducible representations $`\chi _1=1_{K_0(H)},\mathrm{},`$ $`\chi _4𝐆\left(H^{}\right)`$ and three two-dimensional ones $`\pi _1`$, $`\pi _2`$ and $`\pi _3`$. Then, since $`\chi _i\chi _i=1_{K_0(H)}`$, $`\chi _i^{}=\chi _i`$. The involution is an antihomomorphism of $`K_0(H)`$ of order $`2`$, thus it either fixes all $`\pi _k`$ or interchanges two of them and fixes the third one. Assume that we always have $`\pi _2^{}=\pi _2`$. $`\chi _i\pi _k\chi _j+\chi _l`$ as in the case of $`\left|𝐆\left(H^{}\right)\right|=8`$. Thus multiplication by $`\chi _i`$ permutes $`\pi _k`$. Since $`o\left(\chi _i\right)=1`$ or $`2`$ then each $`\chi _i`$ either fixes all $`\pi _k`$ or interchanges two of them and fixes the third one. There are two possible cases: 1. $`\chi _i\pi _k=\pi _k`$ for $`i=1,\mathrm{},4`$ and $`k=1,2,3`$. Then $`m(\chi _i,\pi _k\pi _k^{})`$ $`=`$ $`m(\pi _k,\chi _i\pi _k)=1`$ (6.1) $`\pi _k\pi _k^{}`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{4}{}}}\chi _i\text{ for }k=1,2,3`$ By \[13, Theorem 1\], one of the $`\chi _i`$ is central of order $`2`$ and therefore by Proposition 2.2, $`H`$ has a commutative subHopfalgebra of order $`8`$. Therefore, it should have the same $`K_0`$-ring as one of the Hopf algebras described in Section 3. But none of these $`K_0`$-rings satisfies (6.1). Therefore this case is not possible. 2. $`\chi _i\pi _k\pi _k`$ for some $`i\{1,\mathrm{},4\}`$ and $`k\{1,2,3\}`$. Then, say, $$\chi _1\pi _k=\chi _3\pi _k=\pi _k\text{ for }k=1,2,3$$ but $`\chi _2\pi _k\pi _k`$, $`\chi _4\pi _k\pi _k`$ for some $`k\{1,2,3\}`$. Assume that $`\pi _1^{}=\pi _3`$, $`\pi _3^{}=\pi _1`$ and $`\chi _2\pi _2\pi _2`$. Then $`1`$ $`=`$ $`m(\pi _2,\chi _i\pi _2)=m(\chi _i,\pi _2\pi _2)\text{ for }i=1,3`$ $`0`$ $`=`$ $`m(\pi _2,\chi _i\pi _2)=m(\chi _i,\pi _2\pi _2)\text{ for }i=2,4`$ Therefore $$\pi _2\pi _2^{}=\chi _1+\chi _3+\pi _r=\pi _2\pi _2$$ Since $`\left(\pi _k\pi _k^{}\right)^{}=\pi _k\pi _k^{}`$ , we get $`\left(\pi _r\right)^{}=\pi _r`$ and and thus $`\pi _r=\pi _2`$, that is $$\pi _2\pi _2^{}=\chi _1+\chi _3+\pi _2$$ Therefore $`R=\{a\chi _1+b\chi _3+c\pi _2K_0(H):a,b,c\}`$ is a hereditary subring of $`K_0(H)`$ (see \[24, Definition 3.10\]). There is a one-to-one correspondence between hereditary subrings of $`K_0(H)`$ and Hopf quotients of $`H`$, that is between hereditary subrings of $`K_0(H)`$ and subHopfalgebras of $`H^{}`$ (see \[22, Theorem 6\] or \[24, Proposition 3.11\]). Thus $`H^{}`$ has a subHopfalgebra of dimension $`1+1+4=6`$, which contradicts Nichols-Zoeller Theorem . Thus without loss of generality $`\chi _2\pi _2=\pi _2`$. Then $`\chi _i\pi _2`$ $`=`$ $`\pi _2\text{ for }i=1,\mathrm{},4`$ $`\chi _i\pi _1`$ $`=`$ $`\pi _1\text{ for }i=1,3`$ $`\chi _i\pi _1`$ $`=`$ $`\pi _3\text{ for }i=2,4`$ $`\chi _i\pi _3`$ $`=`$ $`\pi _3\text{ for }i=1,3`$ $`\chi _i\pi _3`$ $`=`$ $`\pi _1\text{ for }i=2,4`$ and therefore $`1`$ $`=`$ $`m(\pi _2,\chi _i\pi _2)=m(\chi _i,\pi _2\pi _2)\text{ for }i=1,\mathrm{},4`$ $`0`$ $`=`$ $`m(\pi _2,\chi _i\pi _k)=m(\chi _i,\pi _k\pi _2^{})\text{ for }i=1,\mathrm{},4\text{}k2`$ $`1`$ $`=`$ $`m(\pi _k,\chi _i\pi _k)=m(\chi _i,\pi _k\pi _k^{})\text{ for }i=1,3`$ $`0`$ $`=`$ $`m(\pi _k,\chi _i\pi _k)=m(\chi _i,\pi _k\pi _k^{})\text{ for }i=2,4`$ $`0`$ $`=`$ $`m(\pi _3,\chi _i\pi _1)=m(\chi _i,\pi _1\pi _3^{})\text{ for }i=1,3`$ $`1`$ $`=`$ $`m(\pi _3,\chi _i\pi _1)=m(\chi _i,\pi _1\pi _3^{})\text{ for }i=2,4`$ Thus we get $`\pi _2\pi _2^{}`$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{4}{}}}\chi _i`$ $`\pi _1\pi _1^{}`$ $`=`$ $`\chi _1+\chi _3+\pi _2=\pi _3\pi _3^{}`$ $`\pi _1\pi _3^{}`$ $`=`$ $`\chi _2+\chi _4+\pi _t`$ $`\pi _k\pi _2^{}`$ $`=`$ $`\alpha _1\pi _1+\alpha _2\pi _2+\alpha _3\pi _3\text{ for }k2`$ and then $`1`$ $`=`$ $`m(\pi _2,\pi _k^{}\pi _k)=m(\pi _k,\pi _k\pi _2^{})\text{ for }k2`$ $`0`$ $`=`$ $`m(\pi _k,\pi _l^{}\pi _l)=m(\pi _l,\pi _l\pi _k^{})\text{ for }k2`$ Therefore $`\pi _1\pi _2^{}`$ $`=`$ $`\pi _1+\pi _3`$ $`\pi _3\pi _2^{}`$ $`=`$ $`\pi _1+\pi _3`$ and $$1=m(\pi _3,\pi _1\pi _2^{})=m(\pi _2,\pi _3^{}\pi _1)=m(\pi _2,\pi _1\pi _3^{})$$ So, finally, $$\pi _1\pi _3^{}=\chi _2+\chi _4+\pi _2$$ Now let’s list all the possible ring structures of $`K_0(H)`$. 6.1. $`𝐆\left(H^{}\right)=\chi \times \phi C_2\times C_2`$ where $`\pi _1^{}=\pi _1`$, $`\pi _2^{}=\pi _2`$ and $`\pi _3^{}=\pi _3`$ and $`\chi \pi _1=\pi _3=\pi _1\chi `$ $`\phi \pi _1=\pi _1=\pi _1\phi `$ $`\pi _1\pi _2=\pi _1+\pi _3=\pi _2\pi _1`$ $`\chi \pi _2=\pi _2=\pi _2\chi `$ $`\phi \pi _2=\pi _2=\pi _2\phi `$ $`\pi _1\pi _3=\chi +\chi \phi +\pi _2=\pi _3\pi _1`$ $`\chi \pi _3=\pi _1=\pi _3\chi `$ $`\phi \pi _3=\pi _3=\pi _3\phi `$ $`\pi _2\pi _3=\pi _1+\pi _3=\pi _3\pi _2`$ $`\pi _1^2=1+\phi +\pi _2=\pi _3^2`$ $`\pi _2^2=1+\chi +\phi +\chi \phi `$ Examples: $`H_{C:1}`$, $`H_{C:\sigma _1},`$ $`kD_{16}`$ and $`kQ_{16}`$. 6.2. $`𝐆\left(H^{}\right)=\chi \times \phi C_2\times C_2`$ where $`\pi _1^{}=\pi _3`$, $`\pi _2^{}=\pi _2`$ and $`\pi _3^{}=\pi _1`$ and $`\chi \pi _1=\pi _3=\pi _1\chi `$ $`\phi \pi _1=\pi _1=\pi _1\phi `$ $`\pi _1\pi _2=\pi _1+\pi _3=\pi _2\pi _1`$ $`\chi \pi _2=\pi _2=\pi _2\chi `$ $`\phi \pi _2=\pi _2=\pi _2\phi `$ $`\pi _1\pi _3=1+\phi +\pi _2=\pi _3\pi _1`$ $`\chi \pi _3=\pi _1=\pi _3\chi `$ $`\phi \pi _3=\pi _3=\pi _3\phi `$ $`\pi _2\pi _3=\pi _1+\pi _3=\pi _3\pi _2`$ $`\pi _1^2=\chi +\chi \phi +\pi _2=\pi _3^2`$ $`\pi _2^2=1+\chi +\phi +\chi \phi `$ Examples: $`H_E`$ and $`kG_2`$, where $`G_2=a,b:a^8=b^2=1,bab=a^3`$. We can now prove Theorem 1.3: ###### Proof. In Sections 5 and 6 we have described all possible Grothendieck ring structures of non-commutative semisimple Hopf algebras of dimension $`16`$ and there are exactly $`7`$ of them. Only one of these $`K_0`$-rings is not commutative, namely $`K_{\text{5}.\text{5}}`$, which corresponds to nonabelian $`𝐆\left(H^{}\right)D_8`$. Therefore, by \[24, Theorem 4.1\] all Hopf algebras with non-commutative $`K_0`$-ring are twistings of each other with a $`2`$-pseudo-cocycle. Moreover, by \[24, 4.5\], Hopf algebras with non-commutative $`K_0`$-rings are not twistings of group algebras. If $`𝐆\left(H^{}\right)`$ is abelian then there are $`6`$ possibilities for the $`K_0`$-ring structure, all of which are commutative, namely $`K_{\text{5}.\text{5}}=K_0\left(k\left(D_8\times C_2\right)\right)`$, $`K_{\text{5}.\text{5}}=K_0\left(kG_7\right)`$, $`K_{\text{5}.\text{5}}=K_0\left(kG_5\right)`$, $`K_{\text{5}.\text{5}}=K_0\left(kG_1\right)`$, $`K_{\text{6}.\text{6}}=K_0\left(kD_{16}\right)`$ and $`K_{\text{6}.\text{6}}=K_0\left(kG_2\right)`$. Thus by \[24, Theorem 4.1\] $`H`$ is a twisting of one of these group algebras with a $`2`$-pseudo-cocycle. Since $`H`$ is semisimple, $`K_0\left(H\right)_{}k`$ is also semisimple by \[32, Lemma 2\]. Therefore, if $`K_0\left(H\right)`$ is commutative, as algebras $`K_0\left(H\right)_{}kk^{\left(10\right)}`$ when $`\left|𝐆\left(H^{}\right)\right|=8`$ and $`K_0\left(H\right)_{}kk^{\left(7\right)}`$ when $`\left|𝐆\left(H^{}\right)\right|=4`$. If $`K_0\left(H\right)`$ is not commutative, that is $`K_0\left(H\right)=K_{\text{5}.\text{5}}`$, it is easy to see that $`dimZ\left(K_0\left(H\right)\right)=7`$ and thus $`K_0\left(H\right)_{}kk^{\left(6\right)}M_2\left(k\right)`$. ∎ ## 7. Twistings of group algebras with a $`2`$-cocycle. All nonabelian groups $`G`$, considered in this section, have an abelian subgroup $`F=\{1,c,b,cb\}C_2\times C_2`$. $`kF\left(kF\right)^{}`$ thus we can identify $`\delta _xkF`$ with the elements of the dual basis. Now define $`JkFkF`$ as follows: (7.1) $`J`$ $`=`$ $`\delta _1\delta _1+\delta _1\delta _c+\delta _1\delta _b+\delta _1\delta _{cb}`$ $`+\delta _c\delta _1+\delta _c\delta _c+i\delta _c\delta _bi\delta _c\delta _{cb}`$ $`+\delta _b\delta _1i\delta _b\delta _c+\delta _b\delta _b+i\delta _b\delta _{cb}`$ $`+\delta _{cb}\delta _1+i\delta _{cb}\delta _ci\delta _{cb}\delta _b+\delta _{cb}\delta _{cb}`$ where $`\delta _1`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(1+c+b+cb\right)`$ $`\delta _c`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(1+cbcb\right)`$ $`\delta _b`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(1c+bcb\right)`$ $`\delta _{cb}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(1cb+cb\right)`$ We can rewrite $`J`$ as $`J`$ $`=`$ $`{\displaystyle \frac{1}{8}}(511+c1+b1+cb1`$ $`+1c+cc+\left(12i\right)bc+\left(1+2i\right)cbc`$ $`+1b+\left(1+2i\right)cb+bb+\left(12i\right)cbb`$ $`+1cb+(12i)ccb+(1+2i)bcb+cbcb)`$ Such a $`J`$ is a 2-cocycle for $`kF`$ and since $`JkGkG`$, it is also a 2-cocycle for $`kG`$. Thus we can form $`\left(kG\right)_J`$ which is a Hopf algebra by , \[24, 2.8\]. By \[31, 6.4\], $`\left(kG\right)_J`$ is non-cocommutative if and only if $`J^1\left(\tau J\right)`$ does not lie in the centralizer of $`\mathrm{\Delta }\left(kG\right)`$ in $`kGkG`$. Moreover, by \[24, Theorem 4.1\] $`K_0\left(\left(kG\right)_J\right)K_0\left(kG\right)`$. Since $`J`$ is a 2-cocycle, then by $`\left(kG\right)_J`$ is triangular. We now discuss Examples $`2`$, $`12`$ and $`13`$ in the tables. We used GAP to compute $`𝐆\left(H\right)`$ in Examples $`12`$ and $`13`$. Example $`\mathrm{𝟐}`$. $`H=\left(k\left(D_8\times C_2\right)\right)_J`$, where $`F=\{1,c,b,cb\}C_2\times C_2`$ is a subgroup of $`D_8\times C_2=a,b,c:a^4=b^2=c^2=1,ba=a^1b`$ and $`J`$ is given by the formula $`\left(\text{7.1}\right)`$. Then 1. $`𝐆\left(H\right)=a^2,b,cC_2\times C_2\times C_2`$. 2. $`𝐆\left(H^{}\right)𝐆\left(k\left(D_8\times C_2\right)\right)C_2\times C_2\times C_2`$ and $`K_0\left(H\right)K_0\left(k\left(D_8\times C_2\right)\right)K_{\text{5}.\text{5}}`$. Example $`\mathrm{𝟏𝟐}`$. $`H=\left(kD_{16}\right)_J`$, where $`F=\{1,a^4,b,a^4b\}C_2\times C_2`$ is a subgroup of $`D_{16}=a,b:a^8=b^2=1,ba=a^1b`$ and $`J`$ is given by the formula $`\left(\text{7.1}\right)`$. Then 1. $`𝐆\left(H\right)=b,g=\frac{1}{2}(a^2+a^2b+a^6+a^6b):g^4=b^2=1,bgb=g^1D_8`$. 2. $`𝐆\left(H^{}\right)𝐆\left(\left(kD_{16}\right)^{}\right)C_2\times C_2`$ and $`K_0\left(H\right)K_0\left(kD_{16}\right)K_{\text{6}.\text{6}}`$. Example $`\mathrm{𝟏𝟑}`$. $`H=\left(kG_2\right)_J`$, where $`F=\{1,a^4,b,a^4b\}C_2\times C_2`$ is a subgroup of $`G_2=a,b:a^8=b^2=1,ba=a^3b`$ and $`J`$ is given by the formula $`\left(\text{7.1}\right)`$. Then 1. $`𝐆\left(H\right)D_8`$. 2. $`𝐆\left(H^{}\right)𝐆\left(\left(kG_2\right)^{}\right)C_2\times C_2`$ and $`K_0\left(H\right)K_0\left(kG_2\right)K_{\text{6}.\text{6}}`$. ## 8. A construction using smash coproducts. Let $`H=kQ_8\mathrm{\#}^\alpha kC_2`$, a smash coproduct of $`kQ_8`$ and $`kC_2`$ (see \[19, 10.6.1\] or \[24, Proposition 3.8\]), where $`Q_8=a,b:a^4=1,b^2=a^2,ba=a^1b`$, $`C_2=\{1,g\}`$. $`H`$ has the algebra structure of $`kQ_8kC_2`$ and the following comultiplication, antipode and counit: $`\mathrm{\Delta }\left(x\mathrm{\#}\delta _{g^k}\right)`$ $`=`$ $`\underset{r+t=k}{{\displaystyle }}\left(x_1\mathrm{\#}\delta _{g^r}\right)\left(\alpha _{g^r}\left(x_2\right)\mathrm{\#}\delta _{g^t}\right)`$ $`S\left(x\mathrm{\#}\delta _{g^k}\right)`$ $`=`$ $`\alpha _{g^k}\left(S\left(x\right)\right)\mathrm{\#}\delta _{g^k}`$ $`\epsilon \left(x\mathrm{\#}\delta _{g^k}\right)`$ $`=`$ $`\epsilon \left(x\right)\delta _{g^k,1}`$ where $`\delta _1=\frac{1}{2}\left(1+g\right)`$, $`\delta _g=\frac{1}{2}\left(1g\right)`$, $`xkQ_8`$ and $`\alpha :GAut\left(kQ_8\right)`$ is defined by $`\alpha _1\left(x\right)`$ $`=`$ $`x`$ $`\alpha _g\left(a\right)`$ $`=`$ $`b`$ $`\alpha _g\left(b\right)`$ $`=`$ $`a,`$ see \[24, Erratum\]. It follows from the above that (8.1) $`\mathrm{\Delta }\left(a\mathrm{\#}1\right)`$ $`=`$ $`\mathrm{\Delta }\left(a\mathrm{\#}\delta _1\right)+\mathrm{\Delta }\left(a\mathrm{\#}\delta _g\right)`$ $`=`$ $`a\mathrm{\#}\delta _1a\mathrm{\#}\delta _1+a\mathrm{\#}\delta _gb\mathrm{\#}\delta _g+a\mathrm{\#}\delta _1a\mathrm{\#}\delta _g+a\mathrm{\#}\delta _gb\mathrm{\#}\delta _1`$ $`=`$ $`a\mathrm{\#}\delta _1a\mathrm{\#}1+a\mathrm{\#}\delta _gb\mathrm{\#}1`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(a\mathrm{\#}1a\mathrm{\#}1+a\mathrm{\#}ga\mathrm{\#}1+a\mathrm{\#}1b\mathrm{\#}1a\mathrm{\#}gb\mathrm{\#}1\right)`$ and (8.2) $$\mathrm{\Delta }\left(b\mathrm{\#}1\right)=\frac{1}{2}\left(b\mathrm{\#}1b\mathrm{\#}1+b\mathrm{\#}gb\mathrm{\#}1+b\mathrm{\#}1a\mathrm{\#}1b\mathrm{\#}ga\mathrm{\#}1\right)$$ Let’s describe $`𝐆\left(H\right)`$, $`𝐆\left(H^{}\right)`$ and $`K_0\left(H\right)`$. By straightforward computations, using $`\left(\text{8.1}\right)`$ and $`\left(\text{8.2}\right)`$, $`𝐆\left(H\right)=a^2\mathrm{\#}1\times 1\mathrm{\#}g`$. $`𝐆\left(H^{}\right)`$is generated by the multiplicative characters $`\chi `$ and $`\phi `$, defined by $`\chi (a)=\chi (g)=1`$, $`\chi (b)=1`$ and $`\phi \left(a\right)=1`$, $`\phi \left(g\right)=\phi \left(b\right)=1`$. Then $`\chi ^1(a)=1`$, $`\chi ^1(b)=\chi ^1(g)=1`$ and $`\phi \chi \phi =\chi ^1`$. Therefore $`𝐆\left(H^{}\right)D_8`$. Degree $`2`$ irreducible representations of $`H`$ are defined by $`\pi _1(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _1(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _1(g)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ $`\pi _2(a)=\left(\begin{array}{cc}\hfill i& \hfill 0\\ \hfill 0& \hfill i\end{array}\right)`$ $`\pi _2(b)=\left(\begin{array}{cc}\hfill 0& \hfill 1\\ \hfill 1& \hfill 0\end{array}\right)`$ $`\pi _2(g)=\left(\begin{array}{cc}\hfill 1& \hfill 0\\ \hfill 0& \hfill 1\end{array}\right)`$ with the property $`\pi _1^2=\pi _2^2=1+\chi ^2+\phi +\chi ^2\phi `$ and therefore $`K_0\left(H\right)K_{\text{5}.\text{5}}`$. ## 9. Main results. ###### Theorem 9.1. Every nontrivial semisimple Hopf algebra $`H`$ of dimension $`16`$ has a commutative subHopfalgebra of dimension $`8`$. ###### Proof. By \[13, Theorem 1\] $`H^{}`$ has a central grouplike $`g`$ of order $`2`$. Thus we get a short exact sequence of Hopf algebras (9.1) $$kg\stackrel{𝑖}{}H^{}\stackrel{𝜋}{}K$$ If $`K`$ is cocommutative, $`K^{}H`$ is commutative and we are done. If $`K`$ is commutative, but not cocommutative, then $`K^{}=k𝐆\left(H\right)H`$ and $`𝐆\left(H\right)`$ is nonabelian of order $`8`$. Applying Proposition 2.2 and results of Section 5 we get that $`H^{}`$ has a commutative subHopfalgebra of dimension $`8`$. Therefore $`H^{}`$ was described in Section 3.3, case B), and it has a group algebra of dimension $`8`$ as a quotient. Therefore $`H`$ has a commutative subHopfalgebra of dimension $`8`$. Now assume that $`K`$ is neither commutative nor cocommutative. Then $`KK^{}H_8`$ and as algebras $`H^{}kg\mathrm{\#}_\sigma K`$, a crossed product of Hopf algebras with an action $`:Kkgkg`$ and a cocycle $`\sigma :KKkg`$. Since $`g`$ is central in $`H^{}`$, the action $``$ is trivial. Let’s prove that $`\sigma `$ is also trivial. By \[16, Theorem 4.8\] $`KH_8`$ doesn’t have nontrivial right Galois objects, thus for any $`2`$-cocycle $`\alpha :KKk`$ the crossed product $`{}_{\alpha }{}^{}K=k\mathrm{\#}_\alpha K`$ is trivial, that is $`ab=\alpha (a_1,b_1)a_2b_2`$. Write $`kg=k\frac{1+g}{2}k\frac{1g}{2}`$ and $`\sigma \left(ab\right)=\alpha _1\left(ab\right)\frac{1+g}{2}+\alpha _2\left(ab\right)\frac{1g}{2}`$. Then for $`j=1,2`$, $`\alpha _j:KKk`$ are $`2`$-cocycles and $`\left(1\mathrm{\#}a\right)\left(1\mathrm{\#}b\right)`$ $`=`$ $`{\displaystyle \sigma (a_1,b_1)\mathrm{\#}a_2b_2}={\displaystyle \frac{1+g}{2}}\mathrm{\#}{\displaystyle \alpha _1(a_1,b_1)a_2b_2}`$ $`+{\displaystyle \frac{1g}{2}}\mathrm{\#}{\displaystyle \alpha _2(a_1,b_1)a_2b_2}={\displaystyle \frac{1+g}{2}}\mathrm{\#}ab+{\displaystyle \frac{1g}{2}}\mathrm{\#}ab=1\mathrm{\#}ab`$ Therefore as algebras $`H^{}kgK`$. As coalgebras $`HkgK^{}kgH_8`$, $`H`$ has $`8`$ grouplikes and we are done by the previous argument. ∎ We now prove Theorem 1.2: ###### Proof. 1. Assume $`𝐆\left(H\right)`$ is abelian of order $`8`$. By Theorem 1.1, $`𝐆\left(H\right)C_2\times C_2\times C_2`$ or $`𝐆\left(H\right)C_4\times C_2`$ and by Proposition 3.1 in this case $`𝐆\left(H^{}\right)`$ is also abelian of order $`8`$. In Propositions 3.2 and 3.3 we have shown that there are exactly $`7`$ nonisomorphic Hopf algebras with $`𝐆\left(H\right)C_4\times C_2`$ and at most $`4`$ nonisomorphic Hopf algebras with $`𝐆\left(H\right)C_2\times C_2\times C_2`$. Now we show that there are $`4`$ distinct Hopf algebras with $`𝐆\left(H\right)C_2\times C_2\times C_2`$. There are $`2`$ nonisomorphic examples of Hopf algebras with $`𝐆\left(H\right)𝐆\left(H^{}\right)C_2\times C_2\times C_2`$, namely $`H_8kC_2`$, which is not triangular (if it were triangular, so would be $`H_8`$), and $`\left(k\left(D_8\times C_2\right)\right)_J`$, which is triangular (see Section 7), and $`2`$ more nonisomorphic Hopf algebras with $`𝐆\left(H\right)C_2\times C_2\times C_2`$ and $`𝐆\left(H^{}\right)C_4\times C_2`$, namely $`\left(H_{b:1}\right)^{}`$ and $`\left(H_{c:\sigma _1}\right)^{}`$. Comparing the structures of $`H_8`$ (see \[11, 2.3, 2.4, 2.8\]) and $`H_{d:1,1}`$ we see that $`H_8kC_2H_{d:1,1}`$, and therefore $`H_{d:1,1}\left(k\left(D_8\times C_2\right)\right)_J`$. 2. Assume that $`𝐆\left(H\right)`$ is nonabelian. Then, by Theorem 9.1, $`H`$ has a commutative subHopfalgebra of dimension $`8`$. By Proposition 3.4 $`𝐆\left(H\right)D_8`$, $`𝐆\left(H^{}\right)=C_2\times C_2`$ and there are exactly $`2`$ such Hopf algebras, $`H_{C:1}H_{B:1}^{}`$ and $`H_EH_{B:X}^{}`$. Comparing their $`K_0`$-rings with $`K_0`$-rings of examples described in Section 7 we see that $`H_{C:1}\left(kD_{16}\right)_J`$ and $`H_E\left(kG_2\right)_J`$. 3. Assume that $`𝐆\left(H\right)`$ is abelian of order $`4`$. By Theorem 1.1, $`𝐆\left(H\right)C_2\times C_2`$. By Theorem 9.1, $`H`$ has a commutative subHopfalgebra of dimension $`8`$ and therefore it was described in Section 3. There are exactly $`3`$ Hopf algebras with this group of grouplikes: two of them, $`H_{B:1}H_{C:1}^{}`$ and $`H_{B:X}H_E^{}`$, have $`𝐆\left(H\right)^{}D_8`$ and one of them, $`H_{C:\sigma _1}`$ has $`𝐆\left(H\right)^{}C_2\times C_2`$ and therefore should be selfdual. Comparing the quotients of $`H_B`$ and $`kQ_8\mathrm{\#}^\alpha kC_2`$ we see that $`H_{B:X}kQ_8\mathrm{\#}^\alpha kC_2`$. ACKNOWLEDGMENTS I would like to thank my Ph.D. advisor Professor Susan Montgomery for numerous discussions, suggestions and comments about this paper. Part of this paper is contained in my Ph.D. thesis in the University of Southern California. The rest of the work was done while I was a postdoctoral fellow at MSRI. I am grateful to MSRI and the organizers of the Noncommutative Algebra program for the support. I would also like to thank the referee for Proposition 2.2 and for the suggestion to use in the proof of Theorem 9.1.
warning/0004/astro-ph0004160.html
ar5iv
text
# Gas properties of HII and Starburst galaxies: relation with the stellar population ## 1 Introduction The investigation of nearby star-forming galaxies plays an important role in the interpretation of the ever increasing data on distant galaxies, as the so-called Lyman-break galaxies seem to be well described by local Starburst galaxies (Meurer et al., 1999). Melnick et al. (1999) propose that HII galaxies can be used as distance estimators over a wide range of redshifts. HII galaxies are among the less luminous star-forming galaxies. Their emission-line spectra and relatively low gaseous metallicities ($``$ 0.1 to 0.25 solar, e.g. Peña et al., 1991) would be consistent with the idea that they are very young galaxies undergoing their first episodes of star formation. On the other hand, in a recent study, Schulte-Ladbeck and Crone (1998) have concluded that in the blue compact dwarf galaxy VII Zw 403 there are older stellar population components. In a previous study (Raimann et al., 1999, hereafter Paper I) we have investigated the stellar population properties of a sample dominated by nearby HII galaxies, but including also Starburst and Seyfert 2 galaxies for comparison purposes. We have considered their continuum and emission/absorption line properties, grouped them into high signal-to-noise templates, and performed stellar population syntheses using a base of stellar cluster spectra (Bica, 1988; Bica and Alloin, 1986). We concluded that most HII galaxies present important flux contributions from populations as old as 500 Myr. In the present paper, we investigate the gaseous properties of the templates obtained in Paper I, after subtraction of the underlying stellar population. We use the emission-line fluxes to classify the spectra according to their excitation characteristics in diagnostic diagrams, to calculate the gaseous abundance and to obtain the age of the last generation of (ionizing) stars. The role of the underlying stellar population, including its effect on the interpretation of data from distant star-forming galaxies and its relation to the emission line properties, is also explored. We present the data set in section 2. The diagnostic diagrams are discussed in section 3. The gas metallicity, the age of the ionizing stellar population and relation between the metallicity and the non-ionizing stellar population are studied in section 4. Possible uses of emission-line galaxies for cosmological studies are discussed in section 5. The concluding remarks are given in section 6. ## 2 The sample The original sample consists of 185 emission-line galaxy spectra obtained by Terlevich and collaborators, most of them (156) discussed individually in Terlevich et al. (1991). The average aperture used in the observations corresponds to $`1.1\times 1.9`$ kpc. More details can be found in Paper I, where we have grouped the spectra according to their continuum, absorption and emission line properties, in order to obtain improved signal-to-noise (S/N) ratio spectra. This strategy was necessary to constrain the stellar population syntheses, mainly for the HII groups. The resulting spectral groups are listed in Table 1 ordered from bluer spectra at the top to redder at the bottom of the table. The groups are named after the member galaxy with the best S/N ratio. The number of the galaxies in each group is shown in column 2. In two cases, G\_UM448 and G\_NGC1510, there were no other similar spectra, and we have thus considered them as “groups” of only one galaxy. The spectral type, according to distribution of emission line spectra of the groups in the Baldwin, Phillips and Terlevich (BPT, 1981) diagnostic diagrams, is given in column 3. In column 4 we list the average absolute magnitudes $`<`$M<sub>B</sub>$`>`$, obtained using the apparent magnitudes and radial velocities from the NASA/IPAC Extragalactic Database (NED)<sup>1</sup><sup>1</sup>1The NASA/IPAC Extragalactic Database (NED) is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. (see Paper I), for $`H_0=75`$ km s<sup>-1</sup> Mpc<sup>-1</sup>, and in columns 5, 6 and 7 we provide the average H$`\beta `$ emission line luminosities, average dimensions at the galaxy corresponding to the angular aperture and average H$`\beta `$ equivalent width, respectively. In columns 8 and 9 we list the S/N ratios in the continuum of each template for spectral regions around $`\lambda `$4200Å and $`\lambda `$5400Å. The spectra of the groups were dereddened according to $`E(BV)_i`$, the internal reddening affecting the stellar population which was obtained in the synthesis (Paper I). The synthesized stellar population from Paper I was then subtracted from the spectrum of each group. The importance of the subtraction of the underlying stellar population for emission line measurements has been discussed in detail by Bonatto et al. (1989). The main effect is on H$`\beta `$, with consequences on gas internal reddening determination and calculated emission line ratios involving this line. Three representative spectra are shown in Fig. 1. In the top of each panel we show the population-free spectrum and in the bottom the dereddened spectrum previously to stellar population subtraction together with the corresponding synthesized spectrum used in the subtraction. The emission spectra for the remaining groups were presented in Raimann (1998). The emission-line fluxes of the population-free spectra were then measured and the corresponding values, relative to H$`\beta `$, are shown in Table 2. In the cases of line blending (e.g. H$`\alpha +`$\[NII\]$`\lambda \lambda 6548,84`$ and \[SII\]$`\lambda \lambda 6717,31`$) the profiles were constrained by adjusting Gaussians of the same width. We have then calculated the residual gas reddening assuming a Galactic reddening law (Seaton, 1979), case B recombination, and an intrinsic ratio H$`\alpha `$/H$`\beta =2.9`$ (Osterbrock, 1989). Only two groups, G\_NGC1510 and G\_NGC3281, presented significant gas reddenings with $`E(BV)_{gas}=`$ 0.24 and 0.32, respectively. The other groups have $`E(BV)_{gas}0.05`$, which are listed in column 2 of Table 3. ## 3 BPT Diagnostic diagrams The reddening-corrected emission-line fluxes were used to locate the groups in the BPT diagnostic diagrams $`log(`$\[OIII\]$`\lambda 5007`$/H$`\beta )`$ $`\times `$ $`log(`$\[NII\]$`\lambda `$6584/H$`\alpha )`$, $`log(`$\[OIII\]$`\lambda 5007`$/H$`\beta )`$ $`\times `$ $`log(`$\[SII\]$`\lambda \lambda `$6717,6731/H$`\alpha )`$ and $`log(`$\[OIII\]$`\lambda 5007`$/H$`\beta )`$ $`\times `$ $`log(`$\[OI\]$`\lambda `$6300/H$`\alpha )`$ (Baldwin et al., 1981; Veilleux and Osterbrock, 1987), shown in Figs. 2 and 3. Fig. 2 contains the present galaxy groups data, together with those of Starburst Nuclear (therein referred to as MGB), Seyfert, LINER and HII galaxies from Coziol (1996). In this diagram the groups G\_UM103, G\_NGC4507 and G\_NGC3281 are located in the region occupied by Seyfert 2 galaxies while G\_UM477 and G\_Mrk710 are in the Nuclear Starburst region. The groups G\_Mrk711, G\_Cam0949-2126, G\_UM140 and G\_NGC3089 are located in a region intermediate between those of HII and Nuclear Starbursts. The remaining groups – the majority – are located in the HII galaxies’ region. The classification is shown in column 3 of Table 1. It can be observed that the HII galaxy groups are located closer to the HII region loci as defined by the dashed lines in this diagram than the individual HII galaxies of Coziol’s (1996) sample. Fig. 3 contains only our groups data, together with the HII models of McCall et al. (1985). In Fig. 3a the HII galaxy groups are also located close to the HII region loci as defined by the dashed line. In order to check if the groups are indeed closer to the theoretical loci than the individual HII galaxies, we have plotted in Fig. 4 the groups data together with the original data for each individual galaxy (Terlevich et al., 1991). It can be observed that the effect is still present. Two factors contribute to this. The subtraction of the stellar population which increases the H$`\beta `$ emission (reducing $`log(`$\[OIII\]/H$`\beta )`$) and the grouping of the individual spectra which produces higher S/N spectra. In HII galaxies with significant contribution of $`t50`$ Myr stellar populations the first effect is particularly important, because these components have strong Balmer absorption lines. We conclude that with spectra of high S/N ratio and corrected for the stellar population absorptions the HII galaxies in the BPT diagnostic diagrams $`log(`$\[OIII\]$`\lambda 5007`$/H$`\beta )`$ $`\times `$ $`log(`$\[NII\]$`\lambda `$6584/H$`\alpha )`$ and $`log(`$\[OIII\]$`\lambda 5007`$/H$`\beta )`$ $`\times `$ $`log(`$\[SII\]$`\lambda \lambda `$6717,6731/H$`\alpha )`$ get closer to the theoretical HII region loci presented by Evans and Dopita (1985) and McCall et al (1985). Previous works have proposed models with very high temperatures to cover the region occupied by HII galaxies (Tresse et al., 1996 and Rola et al., 1997). Our results suggest that this is not necessary and that HII galaxies behave similarly to HII regions. We therefore suggest this observational locus as reference for theoretical models of HII galaxies in the BPT diagrams. In the diagnostic diagram $`log(`$\[OIII\]$`\lambda `$5007/H$`\beta )`$ $`\times `$ $`log(`$\[OI\]$`\lambda `$6300/H$`\alpha )`$ (Fig. 3b) the galaxy groups for which it was possible to measure \[OI\]$`\lambda `$6300 are located close to the HII region loci as defined by the dashed line. However, for many groups – mainly those with the lowest \[OIII\]/H$`\beta `$ – we have obtained only upper limits for \[OI\]$`\lambda `$6300; these limits suggest sistematically lower values for the galaxies when compared to HII region models. On the other hand, there are not enough HII regions data in the literature with low \[OIII\]/H$`\beta `$ and with measured \[OI\]$`\lambda `$6300, so these models are difficult to be compared even with HII region data. The only star forming group which approaches considerably the AGN region is the HII/Starburst group G\_Mrk711. It would be interesting to study in detail galaxies in this group (Paper I), because they might contain intrinsically high temperature HII regions, or alternatively that locus might reflect a composite spectrum, i.e. AGN mixed to HII region emissions due to aperture effects (Pastoriza et al., 1999 and Storchi-Bergmann, 1991). ## 4 Age and metallicity effects We now study the gas properties, using the population-free spectra, in order to investigate the gas metallicity and age of the ionizing stellar population, and their relation to the loci occupied by the groups in the BPT diagrams. ### 4.1 Metallicity Since the metallicity is very sensitive to gas temperature, which is not uniform in the nebulae, it is important to consider the ionization structure (Garnett, 1992). We have adopted the two-zone model of Campbell et al. (1986), where the emission-line ratio \[OIII\]$`\lambda \lambda 4959,5007/\lambda 4363`$ gives the temperature ($`T^{++}`$) of the high ionization species (like O<sup>++</sup> and Ne<sup>++</sup>) and the emission line ratio \[NII\]$`\lambda \lambda 6548,84/\lambda 5755`$ gives the temperature ($`T^+`$) of the low ionization species (like O<sup>+</sup>, N<sup>+</sup> and S<sup>+</sup>). The density is obtained through the emission-line ratio \[SII\]$`\lambda 6717/\lambda 6731`$. We were able to measure \[OIII\]$`\lambda 4363`$ for the HII galaxies and Seyfert 2’s, while for the remaining groups only upper limits could be obtained (column 5 of Table 2). For the latter groups we have used the empirical calibration of Pagel et al. (1979) extrapolated as described in Schmitt et al. (1994) and Storchi-Bergmann et al. (1996) to calculate $`T^{++}`$. We could not detect \[NII\]$`\lambda 5755`$ in any of the spectra. We have thus used the relation of Campbell et al. (1986) to calculate $`T^+`$. The ionic abundance calculations were performed using a three-level atom model (McCall, 1984). The resulting $`T^{++}`$, Ne and ionic abundances are shown in Table 3. The total oxygen abundance (O/H) was calculated by adding the contributions of O<sup>0</sup>, O<sup>+</sup> and O<sup>++</sup>. The nitrogen abundance was calculated by assuming that N/O=N<sup>+</sup>/O<sup>+</sup>, based upon the rough coincidences between the ionization potentials of the two ions (Storchi-Bergmann et al., 1994 and references therein). The ten HII galaxy groups (G\_NGC1510, G\_UM504, G\_UM71, G\_UM461, G\_Tol1924-416, G\_UM448, G\_NGC1487, G\_Tol1004-296, G\_Tol0440-381 and G\_Cam1148-2020) span a range in oxygen abundance of $`7.87<12+log`$(O/H)$`<8.32`$. With small variations, because of different methodologies and sample, these values agree with those of Peña et al. (1991) and Terlevich et al. (1991). The relatively large metallicities attained by the HII galaxies are consistent with the result found in Paper I, that most HII galaxy groups are not single generation, but present previous generations of stars which have enriched the gas. The groups classified as Starbursts – G\_UM477 and G\_Mrk711 – have near solar metallicity, while the ones classified as intermediate between HII and Starburst – G\_Cam0949-2126, G\_Mrk711, G\_UM140 and G\_NGC3089 – have somewhat lower metallicity: 12+log(O/H)$``$8.6. It can be noticed in Table 3 that the Seyfert 2 groups G\_UM103, G\_NGC3281 and G\_NGC4507 have very high temperatures when compared with those predicted by photoionization models (Storchi-Bergmann et al., 1990) which make the calculated chemical abundances lower than expected. This result is typical of Seyfert galaxies and is due to density stratification, to the presence of matter-bounded clouds (Binette et al., 1996), collisional de-excitation and a hard ionizing continuum, effects not present in HII regions (Viegas-Aldrovandi and Gruenwald, 1988), making the above temperature calibration not valid for the Seyferts. Besides, Storchi-Bergmann et al. (1996) have pointed out how critical is the stellar population subtraction to the measurement of the \[OIII\]$`\lambda 4363`$ in Seyfert’s, which usually causes one to overestimate the strength of this line. We have thus revised the metallicities of the Seyfert 2 groups using the calibrations of Storchi-Bergmann et al. (1998) for the narrow-line region (NLR) of active galaxies which do not depend on \[OIII\]$`\lambda 4363`$. These calibrations are a function of the emission-line ratios \[OIII\]$`\lambda \lambda `$4959,5007/H$`\beta `$, \[NII\]$`\lambda \lambda `$6548,84/H$`\alpha `$ and \[OII\]$`\lambda 3727`$/\[OIII\]$`\lambda \lambda 4959,5007`$. We find that the (O/H) abundance is solar for G\_NGC4507 and slightly below solar for G\_UM103 and G\_NGC3281. In Fig. 5 we show $`log`$(N/O) plotted against $`12+log`$(O/H) for the HII galaxies and Starbursts. The distribution of our galaxy groups is consistent with the relation found by Storchi-Bergmann et al. (1994) for star-forming galaxies covering 8.3 $`<`$ 12+log(O/H) $`<`$ 9.4 (solid line in this figure). This relation is predicted by a simple model of galactic chemical evolution with instantaneous recycling, in which nitrogen has a secondary origin and oxygen has primary origin (Storchi-Bergmann et al., 1994 and references therein). Vila-Costas and Edmunds (1993) have shown that the secondary behavior dominates for 12+$`log`$(O/H) $`>`$ 8.2, while for lower abundances nitrogen has mainly a primary origin \[log(O/N) is approximately constant\]. We have too few galaxies in the low abundance end to reach any conclusion there, but we can say that the data are consistent also with the results of Vila-Costas and Edmunds, as all data points in the low metallicity end are located above the line denoting secondary behavior. Finally, we notice that the BPT sequence of HII galaxies is also a sequence of \[NII\]/H$`\alpha `$ values. We have thus plotted in Fig. 6, $`12+log`$(O/H) $`\times log(`$\[NII\]/H$`\alpha )`$ for the HII and Intermediate HII/Starburst groups and conclude that they are very well correlated, with a Spearman rank correlation coefficient of 0.90. $`Log(`$\[NII\]/H$`\alpha )`$ can thus be used as a metallicity index for these galaxies. A linear regression to the data in Fig. 6 gives: $$12+log(O/H)=8.89(\pm 0.07)+0.53(\pm 0.06)\times log([NII]/H\alpha ).$$ This kind of relation has been found by Storchi-Bergmann et al. (1994) for a sample dominated by more metal-rich starbursts. The main difference encountered here is that, for the lower metallicity HII galaxies of Fig. 6, \[NII\]/H$`\alpha `$ shows a wider dynamical range as a function of $`12+log`$(O/H) than do the more metal-rich Starbursts. ### 4.2 Age indicators The equivalent width W(H$`\beta `$), the \[OIII\]$`\lambda \lambda `$4959,5007/H$`\beta `$ ratio and relative volume $`R`$ of the He<sup>+</sup> and H<sup>+</sup> zones are age indicators for HII regions (Dottori, 1981, 1987). Dottori and Bica (1981) have shown that W(H$`\beta `$) can be used to date HII regions of the Magellanic Clouds. Copetti et al. (1986) studied the evolution of the above properties with age through stellar evolution models with a single burst of star formation, which can be used to derive the age of the ionizing star clusters from W(H$`\beta `$) or \[OIII\]/H$`\beta `$. More recently, Stasińska and Leitherer (1996) constructed a new grid of models representing an HII region produced by an evolving starburst embedded in a gas cloud of the same metallicity. They concluded that both W(H$`\beta `$) and W(\[OIII\]) were good age indicators, but not \[OIII\]/H$`\beta `$. According to them, the \[OIII\]/H$`\beta `$ ratio is a poor chronometer because its time dependence is mild for ages smaller than 5 Myr and its behaviour is affected by the ionization parameter. We have thus used the models of Stasińska and Leitherer (1996) to derive the age of the ionizing star cluster for the present HII galaxy groups from the measured W(H$`\beta `$) and W(\[OIII\]). In Figs. 7a and 7b we present a sequence of models (continuous line) in W(\[OIII\]) and W(H$`\beta `$) versus age diagrams together with the data from our groups (horizontal lines). The models with the parameters best suited to our data correspond to $`Z=0.25Z_{}`$. In the above models the continuum is only due to the nebular and ionizing stellar population. As concluded in Paper I, in the galaxy groups there is additional contribution of older stellar populations to the continuum, so that it is necessary to subtract that contribution before calculating W(H$`\beta `$) and W(\[OIII\]). In order to illustrate the effect of this additional component in the derived ages, we present in Fig. 7b both the data previously to the subtraction as dashed lines, and the corrected data as solid lines. The age derived from W(H$`\beta `$) is plotted against the age derived from W(\[OIII\]) in Fig. 7c. Although the ages derived from W(\[OIII\]) are in most cases up to $``$ 1 Myr larger than the ones from W(H$`\beta `$), they are well correlated with Spearman rank correlation cofficient of 0.98. Using both diagrams (Figs. 7a and 7b) we derive an age range of 2.7–5.0 Myr for the ionizing populations in the HII galaxy groups. This age range corresponds to the predicted Wolf-Rayet (WR) phase of a young star cluster (Schaerer & Vacca, 1998). Indeed, WR features have been detected in our HII galaxy spectra and were discussed in Paper I. As the age does not vary much among the HII galaxy groups, the sequence defined in the BPT diagram must be dominated by a varying metallicity (Sect. 4.1). An age effect is suggested when we investigate the relation between the gas metallicity and the contribution of the stellar population with ages $`t>100`$ Myr to the spectra. This can be observed in Fig. 8 where we plot $`12+log`$(O/H) against the percentage of stellar population with ages $`>`$ 100 Myr. This diagram suggests that most of the gas enrichment has been produced by stellar populations older than 100 Myr. The HII galaxies are the ones with smallest contributions of stars older than 100 Myr and most metal-poor while the Seyfert’s have the largest contributions of older populations and are the most metal-rich. There are only two deviant points, corresponding to the Starbursts. These cases can be understood as due to exceptionally high star formation rates in luminous evolved galaxies, spiral or interacting, so that light from stars younger than 100 Myr dominate the observed fluxes of the galaxies. ## 5 Possible cosmological uses Recently, Melnick et al. (1999) have called attention to the possible use of HII galaxies as cosmological probes. In principle (i) emission line velocity dispersions, (ii) HII galaxy luminosities as candles, and (iii) stellar population/emission line properties as probes of galaxy evolution would be important tools to explore. In the following we discuss the latter two possibilities. As can be seen in Table 1 the $`<`$M<sub>B</sub>$`>`$ values vary a lot among HII galaxy groups. Although the sources of B magnitudes are heterogeneous we do not expect that broad band magnitudes be useful as candles, since the contribution from underlying stellar populations of different ages varies considerably, as found in the population syntheses (Paper I). However, H$`\beta `$ emission line luminosities appear to be less dispersed both within and among groups. This is especially so, by considering aperture effects among groups (Table 1). In order to further explore this possibility it would be important to observe whole sets of local calibrators by means of narrow filter CCD imaging in emission lines and neighbouring continuum. For high redshift galaxies, many lines used in the traditional diagnostic diagrams move out of the optical spectral range and typically the spectra have a low S/N ratio in the continuum. Frequently, in these cases, the two strongest emission lines observable in the optical window are \[OII\]$`\lambda 3727`$ Å and H$`\beta `$. Rola et al. (1997, hereafter R97) used the equivalent widths of these lines to construct new diagnostic diagrams in order to classify the spectra of distant emission-line galaxies observed in deep galaxy redshift surveys. Our data in the diagnostic diagram W(\[OII\])/W(H$`\beta )\times log(`$W(H$`\beta ))`$ are shown in Fig. 9. In agreement with R97, the Seyfert galaxies occupy the region with W(\[OII\])/W(H$`\beta )3.5`$ while the Starbursts and HII galaxies occupy the region with W(\[OII\])/W(H$`\beta )3.5`$ and $`log(`$W(H$`\beta ))>1.0`$. Although this diagram segregates AGNs from HII/Starburst galaxies we find that it does not segregate between HII and Starburst galaxies. We have investigated the effect of the stellar population contribution in the above diagram, by subtracting the contribution of the older age components to the continuum of the HII and Starburst galaxies, and leaving only the ionizing continuum for the calculation of the equivalent widths. The result is presented in Fig. 10 which shows a shift in the loci of the galaxies with W(H$`\beta )`$ as low as $``$ 10 Å to values larger than 50 Å, and for the W(\[OII\])/W($`H\beta )`$ from values as large as $`3`$ down to values lower than 1.5. It can be concluded that equivalent widths W(\[OII\]) and W(H$`\beta )`$ of Starburst and HII galaxies show a much smaller range of values once the older stellar population contributions are subtracted, which means that the intrinsic properties of star forming regions do not vary as much as suggested by the diagram before subtraction (Fig. 9). In fact, we conclude that the value of $`log(`$W(H$`\beta ))`$ can be used as an indicator of the percentage contribution of the non-ionizing stellar population to the continuum. In Fig. 11 we plot this percentage $`P`$ against $`log(`$W(H$`\beta ))`$ together with a linear regression to the data: $$P(\%)=153.04(\pm 8.76)56.52(\pm 5.31)\times log(W(H\beta )),$$ for W(H$`\beta )`$ in Å. The Spearman rank correlation coefficient to this correlation is -0.90. W(H$`\beta )`$ is thus a powerful tool for estimating the contribution of older stars to the spectrum of a galaxy. The relation above is very useful, especially for distant galaxies, for which W(H$`\beta )`$ can be obtained once a continuum can be detected underneath H$`\beta `$. On the other hand, back to the R97 diagram, we can conclude that the correction by the stellar population does not move the HII and Starburst galaxies to regions in the diagram outside the limits suggested by R97. But, for very distant galaxies, age ranges will necessarily become narrower and fractions of intrinsically young galaxies should increase; in both cases the lower right part of R97’s diagram should become more and more populated for such high redshift samples, as simulated in our analysis above. ## 6 Concluding remarks We have analysed the emission-line spectra of 19 galaxy templates, obtained from the grouping of 185 emission-line galaxy spectra, as described in Paper I. After correction for internal reddening and subtraction of the synthesized population spectrum, both from Paper I, the emission-line fluxes of each template were measured and corrected for additional gaseous reddening, when present. According to the corresponding locus of each template in the BPT diagrams, the present sample comprises 10 groups of HII galaxies, 3 groups of Seyfert 2, 2 groups of Nuclear Starbursts and 4 groups of intermediate cases between Nuclear Starbursts and HII galaxies. The HII galaxy groups define a much tighter sequence in the BPT diagrams than the individual galaxies, suggesting that the spread obtained in previous studies is due to a combination of lower S/N ratio spectra and of not taking into account the contribution of the underlying stellar population. With population subtraction in spectra of improved S/N ratio, the HII galaxy groups get closer to the theoretical sequences of HII regions presented by Evans and Dopita (1985) and McCall et al. (1985), suggesting that they are more similar to HII regions than concluded in previous works. The resulting sequence of HII templates in the BPT diagram is suggested as a fiducial observational locus for future models of HII galaxies. From the emission-line ratios, we calculate the gas metallicity and age of the ionizing stellar population, and investigate the effect of these two parameters in the BPT diagram above. We conclude that the sequence defined by the HII galaxy templates is primarily due to metallicity, which spans the range 7.87 $`<`$ 12+$`log`$(O/H) $`<`$ 8.32 ($``$ 1/11 to 1/4 solar), while the age of the ionizing stellar population ranges only from 2.7 to 5.0 Myr. A connection with age is suggested when we relate the gas metallicity with the percentage contribution of stellar population components older than 100 Myr, which may indicate that the metal enrichment is mostly due to previous stellar generations, whose signatures are present in the spectral distribution even for the bluest HII galaxies, as discussed in Paper I. The larger the contribution of older stellar components, the more metal rich is the gas. We find a good correlation between $`log(`$\[NII\]/H$`\alpha )`$ and $`12+log`$(O/H) and propose a calibration to obtain the latter from the former. We also explore the effect of the stellar population contribution to the equivalent width diagnostic diagrams of Rola et al. (1997). We conclude that the observed ranges in W(\[OII\])/W(H$`\beta )`$ and W(H$`\beta )`$ are essentially due to the non-ionizing stellar population contribution. By relating $`log(`$W(H$`\beta ))`$ to the percentage contribution of this population, we conclude that there is a tight correlation between these two quantities. We thus propose a calibration which can be used to estimate the non-ionizing stellar population contribution to the spectra from the measured W(H$`\beta )`$, which is particularly useful for probing properties of distant galaxies. ## Acknowledgments T.S.B., E.B. and H.S. (during part of this work) acknowledge support from the Brazilian Institution CNPq, and D.R. from CAPES. We thank Iranderly F. de Fernandes (as CNPq undergraduate fellow) for work related to this project. We thank an anonymous referee for valuable suggestions which helped to improve the paper.
warning/0004/astro-ph0004227.html
ar5iv
text
# Can our Universe be inhomogeneous on large sub-horizon scales? ## I Introduction The last year or so has seen the first serious attempts to provide some direct connections between “fundamental” high-energy physics and “low-energy” standard cosmology . Although this “top-down” approach is still at a very early stage, a number of crucial general trends already became apparent. For example, since high-energy theories are formulated in higher dimensions, any low energy limit will necessarily involve dimensional reduction, and possibly also compactification . This turns out to be crucial because as a result of this process the low energy, four-dimensional coupling constants become functions of the radii of extra dimensions, which are often variables. One can therefore end up with low-energy effective models in which some of the “fundamental” constants of nature are time and/or space-varying quantities. There are a number of known examples of such models . On the other hand, there are recent tentative suggestions of a time-variation of the fine structure constant , but these require further confirmation. It is therefore interesting to study the possible observational signatures of such variations, and in particular to find out how such observational signatures constrain the possible models. It turns out to be easier to study this issue by constructing simple “toy models”. This provides a “bottom-up” approach, in which one gives up the possibility of testing particular assumptions from first principles, but instead has the possibility of exploring a larger patch of parameter space. This idea goes back at least to Dirac, and had its first detailed realization with the Brans-Dicke model , which has a varying $`G`$. A number of toy models have recently been constructed to analyse possible variations of the fine structure constant , the speed of light and electric charge . A somewhat related approach is that of “quintessence” (see for example ). These are essentially models with a time-varying cosmological constant. Here we consider the possible observational effects of having a universe made up of different domains, each with a different value of the cosmological constant. Such a structure would dramatically influence the future evolution of the universe . Recent observations of Type Ia Supernovae up to redshifts of about $`z1`$ , when combined with CMBR anisotropy data, seem to indicate that our patch of the universe is currently characterised by the parameters $`\mathrm{\Omega }_\mathrm{\Lambda }^00.7`$ and $`\mathrm{\Omega }_m^00.3`$, implying that the cosmological constant become important only very recently. As has been pointed out before, the Supernovae measurements are local, and so they can not be extrapolated all the way to the horizon. For example, we could be living in a small, sub-critical bubble, and our local neighbourhood could have a value of $`\mathrm{\Omega }_\mathrm{\Lambda }^00.7`$ that is uncharacteristic. Here we discuss some basic consequences of such a scenario. We shall assume that different regions of space have different values of the vacuum energy density, separated by domain walls. This can be achieved if there is a scalar field, say $`\varphi `$, which within each region sits in one of a number of possible minima of a time-independent potential. The above simplifying assumptions could be relaxed; for example one could instead consider quintessence-type fields. This would introduce quantitative differences, but would not change the basic qualitative results we are discussing. In the following section we describe our numerical simulations of the evolution of the domain wall network. We then proceed to discuss the basic features of the structure formation mechanism for this scenario in section III. Finally we present our results in section IV and discuss our conclusions in section V. ## II Evolution of the domain walls We consider the evolution of a network of domain walls in a k=0 Friedmann-Robertson-Walker universe with line element: $$ds^2=a^2(\eta )(d\eta ^2dx^2dy^2dz^2),$$ (1) where $`a(\eta )`$ is the cosmological expansion factor and $`\eta `$ is the conformal time. The dynamics of a scalar filed $`\varphi `$ is determined by the Lagrangian density, $$=\frac{1}{4\pi }\left(\frac{1}{2}\varphi _{,\alpha }\varphi ^{,\alpha }+V(\varphi )\right),$$ (2) where we will take $`V(\varphi )`$ to be a generic $`\varphi ^4`$ potential with two degenerate minima $$V(\varphi )=V_0\left(\frac{\varphi ^2}{\varphi _0^2}1\right)^2+C(\varphi ).$$ (3) where $`C(\varphi )`$ smoothly interpolates between $`V_{}`$ at $`\varphi =\varphi _0`$ and $`V_+`$ at $`\varphi =+\varphi _0`$ and is otherwise such that the potential $`V(\varphi )`$ has two minima at $`\pm \varphi _0`$ which have different energies. This obviously admits domain wall solutions . The precise form of the function $`C(\varphi )`$ is not important as it will not affect domain walls dynamics if $`C(\varphi )V_0`$ for all possible values of the scalar field $`\varphi `$. By varying the action $$S=𝑑\eta d^3x\sqrt{g},$$ (4) with respect to $`\varphi `$ we obtain the field equation of motion: $$\frac{^2\varphi }{\eta ^2}+2\frac{\dot{a}}{a}\frac{\varphi }{\eta }^2\varphi =a^2\frac{V}{\varphi }.$$ (5) with $$^2=\frac{^2}{x^2}+\frac{^2}{y^2}+\frac{^2}{z^2}.$$ (6) When making numerical simulations of the evolution of domain wall networks (or indeed other defects) it is also often convenient to modify the equation of motion for the scalar field $`\varphi `$ in such a way that the comoving thickness of the walls is fixed in comoving coordinates. This is known as the PRS algorithm and it will not significantly affect the large-scale dynamics of domain walls. Hence, we will modify the evolution equation for the scalar field $`\varphi `$ according to the PRS prescription: $$\frac{^2\varphi }{\eta ^2}+\beta _1\frac{\dot{a}}{a}\frac{\varphi }{\eta }^2\varphi =a^{\beta _2}\frac{V}{\varphi }.$$ (7) where $`\beta _1`$ and $`\beta _2`$ are constants. We choose $`\beta _2=0`$ in order for the walls to have constant comoving thickness and $`\beta _1=3`$ by requiring that the momentum conservation law for how a wall slows down in an expanding universe is maintained . We perform two-dimensional simulations of domain wall evolution for which $`^2\varphi /z^2=0`$. These have the advantage of allowing a larger dynamic range and better resolution than tree-dimensional simulations. We assume the initial value of $`\varphi `$ to be a random variable between $`\varphi _0`$ and $`\varphi _0`$ and the initial value of $`\dot{\varphi }`$ to be equal to zero everywhere. We normalise the numerical simulations so that $`\varphi _0=1`$. We have checked that the initial conditions are unimportant (as expected), because the domain wall network rapidly approaches a scaling solution with its statistical properties being independent of the initial configuration. ## III Evolution of cosmological perturbations As described previously, we shall assume that the universe is made up of several regions (domains), with different vacuum energy densities. We will discuss the case where there are two such possible values, but it would be easy to generalise this to a distribution with a continuous range of values. Moreover, one assumes that the thin region separating any two of the domains considered (domain wall) is not relevant for structure formation. This happens if the potential of the field is small enough at the origin. Vacuum energy becomes dominant only for recent epochs and so we shall be concerned with the evolution of perturbations only in the matter-dominated era, neglecting the contribution of the radiation component. The average vacuum density is $`\rho _\mathrm{\Lambda }=(V_++V_{})/2`$, so that $$\overline{\mathrm{\Omega }}_\mathrm{\Lambda }\rho _\mathrm{\Lambda }/\rho _c=\frac{V_++V_{}}{2\rho _c},$$ (8) where $`\rho _c`$ is the critical density. We define $`\mathrm{\Delta }\mathrm{\Omega }_\mathrm{\Lambda }`$ in a particular domain as $$\mathrm{\Delta }\mathrm{\Omega }_\mathrm{\Lambda }=\mathrm{\Delta }\rho _\mathrm{\Lambda }/\rho _c=\frac{2VV_+V_{}}{2\rho _c}$$ (9) where $`V`$ is the value of the vacuum energy density inside the domain. In our case $`\mathrm{\Delta }\mathrm{\Omega }_\mathrm{\Lambda }`$ can have one of two possible values $`\mathrm{\Delta }\mathrm{\Omega }_\mathrm{\Lambda }=\pm |V_+V_{}|/2\rho _c`$. In the synchronous gauge, the linear evolution equation for cold dark matter density perturbations, $`\delta _m`$, in a flat universe with a non-zero cosmological constant can be written as $$\ddot{\delta }_m+\dot{\delta }_m\frac{3}{2}^2\left(\overline{\mathrm{\Omega }}_m\delta _m2\mathrm{\Delta }\mathrm{\Omega }_\mathrm{\Lambda }\right)=0,$$ (10) where the evolution of the scale factor $`a`$, is governed by the Friedmann equation $$^2=_0^2\left(\overline{\mathrm{\Omega }}_m^0a^1+(1\overline{\mathrm{\Omega }}_m^0)a^2\right).$$ (11) Here a dot represents a derivative with respect to conformal time, the superscript ‘0’ means that the quantities are to be evaluated at the present time, and $`=\dot{a}/a`$. Note that the average matter and vacuum energy densities at an arbitrary epoch can be written as $$\overline{\mathrm{\Omega }}_m=\frac{\overline{\mathrm{\Omega }}_m^0}{\overline{\mathrm{\Omega }}_m^0+(1\overline{\mathrm{\Omega }}_m^0)a^3}$$ (12) and $`\overline{\mathrm{\Omega }}_\mathrm{\Lambda }=1\overline{\mathrm{\Omega }}_m`$, where we have also chosen $`a_0=1`$. We start the simulation sufficiently early (say at a red-shift $`z_i=10^3`$) so that we do not have to worry about the initial compensation. Because our aim is to investigate the average equation of state of each domain we shall assume the following initial conditions for eq. (10): $$\delta _m(\eta _i)=0,\dot{\delta }_m(\eta _i)=0.$$ (13) We will parametrise the density perturbations $`\delta _m`$ in each domain as a fluctuation $`\mathrm{\Delta }\mathrm{\Omega }_m=\rho _m\delta _m/\rho _c=\delta _m\overline{\mathrm{\Omega }}_m`$ in the local value of the matter density. It is easy to show that the time component $$\tau _{00}=\frac{3}{8\pi G}^2\left(\mathrm{\Delta }\mathrm{\Omega }_m+\mathrm{\Delta }\mathrm{\Omega }_\mathrm{\Lambda }\right)+\frac{2}{8\pi G}\dot{\delta },$$ (14) of the pseudo-stress-energy tensor, $`\tau _{\mu \nu }`$, must be compensated on super-horizon scales (with $`\tau _{00}k^4`$) . Here, the quantity $`\dot{\delta }/3`$ can be interpreted as a fractional variation in the local expansion rate parametrised by $`h`$. We verified that: $$\mathrm{\Delta }\mathrm{\Omega }_\mathrm{\Lambda }+\mathrm{\Delta }\mathrm{\Omega }_m2\frac{\mathrm{\Delta }h}{h}0,$$ (15) where $`\mathrm{\Delta }h/h=\dot{\delta }/3`$ except near the boundary between different domains as expected from the previous discussion. Note that the factor of two in the ‘Hubble’ term of the above expression arises because the Friedmann equation relates $`^2`$ to the average matter and vacuum energy densities. This can be confirmed in Fig. 1, which shows the value of the three terms above, as well as of its sum, for a particular simulation. The cancellation is almost perfect everywhere except where there are domain walls. A more detailed study also shows that at the domain walls the area where the sum of the three terms is positive is equal to that where the sum is negative, so that these average out to zero over the whole box. ## IV Results and discussion One key assumption of our model is that the energy scale of the phase transition which produced the domain walls is sufficiently low so that the domain walls have a negligible effect on structure formation. The standard bound of 1 MeV obviously applies here. The main consequence of this assumption is that the inhomogeneities are not generated by the domain walls but are due to the different vacuum densities in different domains. The fact that the vacuum densities only become important at late times explains why these inhomogeneities in the local values of the cosmological parameters $`\mathrm{\Omega }_\mathrm{\Lambda }^0`$, $`\mathrm{\Omega }_m^0`$ and $`h`$ are created only at recent times. Different values of $`\mathrm{\Omega }_\mathrm{\Lambda }^0`$ and $`\mathrm{\Omega }_m^0`$ lead to different linear growth factors from early times to the present. For primordial perturbations in a flat $`\mathrm{\Lambda }CDM`$ model, the quantity $$g(\mathrm{\Omega }_m^0)=\frac{5\mathrm{\Omega }_m^0}{2}\left(\frac{1}{70}+\frac{209\mathrm{\Omega }_m^0}{140}\frac{(\mathrm{\Omega }_m^0)^2}{140}+(\mathrm{\Omega }_m^0)^{4/7}\right)^1$$ (16) provides a very good fit to the suppression of growth of density perturbations relative to that of a $`\mathrm{\Omega }_m^0=1`$ universe . This rescaling was also shown to be valid for generic topological defect models for structure formation on all scales of cosmological interest and for any reasonable combination of the cosmological parameters $`h`$, $`\mathrm{\Omega }_m^0`$ and $`\mathrm{\Omega }_\mathrm{\Lambda }^0`$ . Our model will thus lead to a universe made up of several domains, in which the growth factor $`g(\mathrm{\Omega }_m^0)`$ has different values. The result is an inhomogeneous universe today with, for example, the abundance of clusters of galaxies varying from one position to another. Hence, we expect the presence of occasional ‘great walls’ separating domains with different values of the cosmological parameters at high red-shift. However, given that in our model the domain walls do not generate any cosmologically relevant fluctuations these ‘great walls’ will simply provide a smooth transition between domains with different values of the cosmological parameters. This can clearly be seen in Figs. 1 and 2. Given that the length-scale corresponding to these inhomogeneities is expected to be close to the horizon scale (simply due to the dynamics of the domain walls), we may not be able to realize that we live in an inhomogeneous universe. This happens because as we look far away we are also looking backwards in time and the universe will get more and more homogeneous as the red-shift of the cosmological objects we are looking at increases (note that $`\mathrm{\Delta }\mathrm{\Omega }_\mathrm{\Lambda }=\mathrm{\Delta }\mathrm{\Omega }_\mathrm{\Lambda }^0/(\overline{\mathrm{\Omega }}_m^0(1+z)^3+(1\overline{\mathrm{\Omega }}_m^0))`$ decreases very rapidly with red-shift). In fact, the higher the red-shift we are looking at the larger is the possibility of finding a domain with different local values of the cosmological parameters but the harder is to realize that. This is clearly illustrated in Fig. 2, which shows the values of $`\mathrm{\Omega }_\mathrm{\Lambda }`$, $`\mathrm{\Omega }_m`$ and $`\mathrm{\Delta }h/h`$ for a particular simulation. One should notice the different colour schemes being used for $`z=0`$ and $`z=1`$; had we used the same colour scheme for the earlier redshift, no significant fluctuations would be detectable by visual inspection. This also explains why CMB fluctuations created at last scattering will be completely negligible. Significant CMB fluctuations can only be created at late times but even these are expected to be small if we do not live near the edge of our domain. However, we know that to be true (if $`\mathrm{\Delta }\mathrm{\Omega }_\mathrm{\Lambda }^0`$ is not too small) because otherwise the Universe would look very anisotropic. This makes the observational detection of this effect somewhat non-trivial. The best way of doing it should be through the determination of the number density of objects as a function of redshift in different directions, assuming that one has a reliable understanding of other possible evolutionary effects. Specific examples would be the counting of X-ray or Sunyaev-Zel’dovich galaxy clusters as a function of the red-shift $`z`$ such as can be performed by XMM or Planck , large-scale velocity flows or gravitational lensing statistics of extragalactic surveys . Another possibility is to look for cosmological anisotropies out to $`z1`$ with Supernovae Ia. Results from a recent analysis using a combined sample of 79 high and low red-shift supernovae are consistent with an homogeneous and isotropic universe but do not exclude the existence of significant anisotropies on cosmological scales. Unfortunately, some of their assumptions and results do not apply to our model. Such an investigation within the scope of our model is more complex because the allowed variations on the local values of the cosmological parameters are not independent and the spatial distribution of the patches with different values of the cosmological parameters is unknown. A simplified analysis of Supernova and CMB data constraints was recently performed in ## V Conclusions In this paper we have provided a simple example of a cosmological scenario where the universe becomes inhomogeneous at a very recent epoch, in a way which is perfectly consistent with current observations assuming that we are not very close to one of the boundaries. The inhomogeneity arises due to the onset of vacuum energy domination, if the value of the ‘cosmological’ constant is different in different domains. This in turn implies that the subsequent dynamics of each patch of the universe will be different, leading to different values of other cosmological parameters in each patch, such as the matter density and the Hubble constant. The size of each patch is determined by the dynamics of the domain walls but is expected to be of the order of the horizon (although causality prevents it from exceeding it). This fact makes the detection of the cosmological signatures of this kind of model more difficult given that the contribution of the vacuum energy density rapidly becomes negligible with increasing red-shift. ###### Acknowledgements. We are grateful to A. Lasenby, G. Rocha and P. Viana for useful discussion and comments. We thank “Fundação para a Ciência e Tecnologia” (FCT) for financial support, and “Centro de Astrofísica da Universidade do Porto” (CAUP) for the facilities provided. C. M. is funded by FCT under “Programa PRAXIS XXI” (grant no. PRAXIS XXI/BPD/11769/97).
warning/0004/math-ph0004003.html
ar5iv
text
# General Theory of Lee-Yang Zeros in Models with First-Order Phase Transitions ## Potts Model: The Hamiltonian is $$\beta H=J\underset{x,y}{}\delta _{\sigma _x,\sigma _y}h\underset{x}{}\delta _{\sigma _x,1}.$$ (7) with spins $`\sigma \{1,2,\mathrm{},q\}`$, real coupling $`J>0`$, and complex field $`h`$. For $`h=0`$ this is the standard Potts model, with a $`q`$-fold degenerate ordered phase at large $`J`$ and a disordered phase at small $`J`$, coexisting at $`J_c^q\frac{1}{d}\mathrm{log}q`$. The transition is first-order for large $`q`$ KS , while it is presumably second order for $`qq_c(d)`$. For $`h0`$ and $`q`$ large, the phase diagram was determined first by formal expansion G , and recently by rigorous probabilistic methods BBCK . The Lee-Yang zeros of (7) were studied numerically in KC , where it was suggested that the zeros lie on almost circular curves slightly outside the unit circle, for $`J`$ both above and below $`J_c^q`$. While, by three-phase coexistence, this turns out to be incorrect for $`J<J_c^q`$ (see Fig. 2), we prove that this is indeed the case for $`JJ_c^q`$, thus resolving a controversy in KC . The model (7) has three phases: the disordered phase $`(D)`$ with degeneracy $`q_D=1`$, and two ordered phases: a magnetized $`(M)`$ and a non-magnetized $`(O)`$ phase, with degeneracies $`q_M=1`$ and $`q_O=q1`$, characterized by abundances of $`\sigma _z=1`$ and $`\sigma _z=\text{const.}1`$, respectively. Let us abbreviate $`z=e^h`$, $`\kappa _d=d(2d1)`$, $`Q_z^{(k)}=q1+z^k`$, and $`Q_z=Q_z^{(1)}`$. The free energies are given BBCK by $`e^{\beta f_D}`$ $`=Q_z\mathrm{exp}\left\{d(e^J1)Q_z^{(2)}/Q_z^2+\kappa _d(e^J1)^2Q_z^{(3)}/Q_z^3(\kappa _d+1/2)(e^J1)^2[Q_z^{(2)}/Q_z^2]^2+𝒪(1/q^{34/d})\right\}`$ $`e^{\beta f_M}`$ $`=z\mathrm{exp}\left\{dJ+e^{2dJ}(Q_zz^11)+de^{(4d1)J}(Q_z^2+e^JQ_z^{(2)})z^2(d+1/2)e^{4dJ}Q_z^2z^2+𝒪(1/q^{32/d})\right\}`$ $`e^{\beta f_O}`$ $`=\mathrm{exp}\left\{dJ+e^{2dJ}(Q_z1)+de^{(4d1)J}(Q_z^2+e^JQ_z^{(2)})(d+1/2)e^{4dJ}Q_z^2+𝒪(1/q^{32/d})\right\}.`$ The zeros of the periodic Potts partition function are depicted in Fig. 2. In particular, for $`J_{}<J<J_+`$ (where $`J_+=J_c^q`$ and $`J_{}J_c^{q2}`$), the loci do not lie on a single closed curve but rather split the complex plane into three pieces, corresponding to the regions of stability of the three phases above. The number of zeros on the inner arc is roughly $`V/(2\pi q)`$, so one needs to take $`V`$ quite large and tune $`J`$ to fall inside the narrow window $`(J_{},J_+)`$ to find any interior zeros. This explains why these zeros were not detected in previous numerical work KC . Despite their appearance, none of the curves in Fig. 2 is a circle. This is verified by finding the coexistence curves (2) for three distinct pairs $`k,\mathrm{}\{D,M,O\}`$. When $`JJ_c^q`$, only the phases $`M`$ and $`O`$ are relevant, and the asymptotic location of the zeros is given by $`\mathrm{}𝔢f_O=\mathrm{}𝔢f_M`$. For $`z=re^{i\theta }`$, this easily implies $$r=1+qe^{2dJ}(1\mathrm{cos}\theta )+𝒪(1/q^2),$$ so that for $`JJ_c^q`$ and $`V\mathrm{}`$, all zeros with $`|\theta |1/\sqrt{q}`$ are asymptotically outside the unit circle. By invoking arguments similar to remark , this extends to all $`\theta `$ BBCK . There are two finite-volume corrections: an outward shift of order $`1/V`$ due to $`f_O>f_{O,L}^{\text{eff}}`$ (see Eq. (2)) and an error $`𝒪(e^{L/L_0})`$ coming from (1). Since $`1/V𝒪(e^{L/L_0})`$, this proves the initial numerical observation in KC . To make the interesting features clearly visible, Figs. 1 and 2 were drawn for values of $`e^{4J}`$ and $`q`$ for which we have not proved convergence of our expansions. However, as established above, all the depicted behaviors indeed occur once $`e^J`$ (or $`1/q`$) and $`e^{L/L_0}`$ are small enough. In summary, we identify the loci of complex zeros with the complex phase coexistence curves. For particular models, we use this identification to map the precise location of these zeros. We find that, in general, the loci are non-uniform and that the resulting curves are non-circular; if more that two phases are present, the curves also have bifurcation (i.e., splitting) points. The research of R.K. was partly supported by the grant GAČR 201/00/1149.
warning/0004/math0004136.html
ar5iv
text
# The Aspinwall-Morrison calculation and Gromov-Witten theory ## 1. Introduction I. A bit of history. (See for a good reference on the history of the problem.) One of the problems in the old and recent story of mirror symetry has been the issue of multiple covers on a Calabi-Yau 3-fold $`X`$. In the pre Gromov-Witten era, this problem can be explained in terms of topological field theories. Let $`X`$ be a Calabi-Yau threefold and $`H_1,H_2,H_3H^2(X)`$. The corresponding 3-point correlator in the A-model of $`X`$ is a path integral that can be expressed as follows: (1) $$H_1,H_2,H_3=_XH_1H_2H_3+\underset{\beta H_2(X)}{}N_\beta (H_1,H_2,H_3)q^\beta .$$ We explain the notation. The parameter $`q=(q_1,\mathrm{}q_k)`$ is a local coordinate on the K$`\ddot{a}`$hler moduli space of $`X`$. Let $`(d_1,\mathrm{},d_k)`$ be the coordinates of $`\beta `$ with respect to an integral base of the Mori cone of $`X`$. Then $`q^\beta :=q_1^{d_1}\mathrm{}q_k^{d_k}`$. The path integral is not a well defined notion, but beyond that, and probably more importantly, there is no rigorous definition of $`N_\beta (H_1,H_2,H_3)`$ in the framework of topological field theories. Let $`Z_i`$ for $`i=1,2,3`$ be a cycle whose fundamental class is Poincar$`\stackrel{´}{e}`$ dual to $`H_i`$. Heuristically, the “invariant” $`N_\beta (H_1,H_2,H_3)`$ is described as the “number” of holomorphic maps in (2) $$\{f:^1X|f_{}([^1])=\beta ,f(0)Z_1,f(1)Z_2,f(\mathrm{})Z_3\}.$$ This is certainly not precise, for there may be infinitely many such maps. Let $`CX`$ be a smooth rational curve. Fix an isomorphism $`g:^1C`$. For any degree $`k`$ multiple cover $`f:^1^1`$ the composition $`fg:^1C`$ satisfies $`(fg)_{}([^1])=k[C]`$. One would then naturally ask: What is the contribution of the space of degree $`k`$ multiple covers of $`C`$ to the “invariant” $`N_{k[C]}(H_1,H_2,H_3)`$? Since this question is about the numbers $`N_{k[C]}(H_1,H_2,H_3)`$, it is not a well posed one. It can be made precise in the framework of Gromov-Witten theory. The answer was conjectured in by looking at the classical example of a Calabi-Yau. If $`X`$ is a quintic threefold then $`H^2(X)`$ is one dimensional. Let $`H`$ be its generator. The 3-point correlator of the quintic can be calculated explicitly: (3) $$H,H,H=5+\underset{d=1}{\overset{\mathrm{}}{}}n_dd^3\frac{q^d}{1q^d},$$ where $`n_d`$ is the virtual number of degree $`d`$ rational curves (instantons) in the quintic. The instanton number $`n_d`$ agrees with the number of degree $`d`$ rational curves in the quintic if every rational curve of degree $`d`$ is smooth, isolated and with normal bundle $`N=𝒪(1)𝒪(1)`$. This is not the case for there are $`6`$-nodal rational plane quintic curves on a generic quintic threefold (see ), hence a rigorous definition of the instanton numbers $`n_d`$ did not exist. The last equation can be transformed as follows: (4) $$H,H,H=5+\underset{d=1}{\overset{\mathrm{}}{}}(\underset{k|d}{}n_kk^3)q^d.$$ By comparing it to the equation (1) we can see that: (5) $$N_d(H,H,H)=\underset{k|d}{}n_kk^3.$$ It looks that each degree $`k`$ rational curve $`C`$ in the quintic 3-fold $`X`$ contributes by: (6) $$_CH_CH_CH$$ to $`N_d(H,H,H)`$ for any $`d`$ such that $`k|d`$. For a general Calabi-Yau $`X`$, the (pre Gromov-Witten) multiple cover formula can be formulated as follows: Let $`CX`$ be a smooth, rational curve such that $`N_{C/X}=𝒪_C(1)𝒪_C(1)`$. The contribution of degree $`k`$ multiple covers of $`C`$ in $`N_{k[C]}(H_1,H_2,H_3)`$ is: (7) $$_CH_1_CH_2_CH_3.$$ It was in this form that the multiple cover formula was proven by Aspinwall and Morrison in and by Voisin in . It follows from the above equation that: $`N_\beta (H_1,H_2,H_3)={\displaystyle \underset{\beta =d\gamma }{}}n_\gamma {\displaystyle _\gamma }H_1{\displaystyle _\gamma }H_2{\displaystyle _\gamma }H_3`$ (8) $`=({\displaystyle \underset{\beta =d\gamma }{}}n_\gamma d^3){\displaystyle _\beta }H_1{\displaystyle _\beta }H_2{\displaystyle _\beta }H_3`$ where $`n_\gamma `$ is the virtual number (instantons) of rational curves of type $`\gamma `$ in $`X`$. A rigorous definition of $`N_\beta `$ and $`n_\beta `$ requires a new conceptual framework which is now known as Gromov-Witten theory. Let $`X`$ be a smooth, projective manifold and $`\beta H_2(X)`$. There is a moduli stack $`\overline{M}_{0,n}(X,\beta )`$ which parametrizes pointed, stable maps of degree $`\beta `$. Universal properties of these maduli stacks imply the existence of several features: $`e=(e_1,e_2,\mathrm{},e_n):\overline{M}_{0,n}(X,\beta )X^n,\pi _n:\overline{M}_{0,n}(X,\beta )\overline{M}_{0,n1}(X,\beta )`$ (9) $`\pi :\overline{M}_{0,n}(X,\beta )\overline{M}_{0,0}(X,\beta ),\widehat{\pi }:\overline{M}_{0,n}(X,\beta )\overline{M}_{0,n}.`$ The map $`e`$ evaluates the pointed, stable map at the marked points, $`\pi _n`$ forgets the last marked point and collapses the unstable components of the source curve, $`\pi `$ forgets the marked points and $`\widehat{\pi }`$ forgets the map and stabilizes the pointed source curve. The expected dimension of $`\overline{M}_{0,n}(X,\beta )`$ is $`\text{dim}X+_\beta (K_X)+n3`$. The dimension of the moduli stack of stable maps may be greater than the expected dimension. In this case, a Chow class of the expected dimension has been constructed. It plays the role of the fundamental class, hence it is called the virtual fundamental class and denoted by $`[\overline{M}_{0,n}(X,\beta )]^{\text{vir}}`$ (see ,). Let $`X`$ be a Calabi-Yau threefold and $`H_1,H_2,H_3H^2(X)`$. In the Gromov-Witten setting: (10) $$N_\beta (H_1,H_2,H_3):=_{[\overline{M}_{0,3}(X,\beta )]^{\text{virt}}}e_1^{}(H_1)e_2^{}(H_2)e_3^{}(H_3).$$ The expected dimension of $`\overline{M}_{0,0}(X,\beta )`$ is zero. Let: (11) $$N_\beta :=\text{deg}([\overline{M}_{0,0}(X,\beta )]^{\text{virt}})$$ By the divisor axiom: (12) $$N_\beta (H_1,H_2,H_3)=N_\beta _\beta H_1_\beta H_2_\beta H_3.$$ Let $`CX`$ be a smooth rational curve with $`N_{C/X}=𝒪_C(1)𝒪_C(1)`$. The moduli space $`\overline{M}_{0,0}(X,d[C])`$ contains a component of positive dimension, namely $`\overline{M}_{0,0}(C,d)`$. The dimension of this component is $`2d2`$. Consider the following diagram: $$\begin{array}{ccc}\overline{M}_{0,1}(C,d)& \stackrel{e_1}{}& C\\ \pi & & \\ \overline{M}_{0,0}(C,d)\end{array}$$ The sheaf: (13) $$V_d:=R^1\pi _{}(𝒪_C(1)𝒪_C(1))$$ is locally free of rank $`2d2`$. Let $`𝔼_d`$ be its top chern class. An assertion of Kontsevich in , which was proven by Behrend in , states that the part of $`[\overline{M}_{0,0}(X,\beta )]^{\text{virt}}`$ supported in $`\overline{M}_{0,0}(C,d)`$ is Poincar$`\stackrel{´}{e}`$ dual to $`𝔼_d`$. The multiple cover formula in this context says that: (14) $$_{\overline{M}_{0,0}(C,d)}𝔼_d=d^3$$ i.e. the curve $`C`$ contributes by $`d^3`$ to to $`N_{d[C]}`$. The multiple cover formula in this form was proven by Kontsevich , Lian-Liu-Yau , Manin and Pandharipande . By the divisor property, the multiple cover formula in this context follows from: (15) $$_{\overline{M}_{0,3}(C,d)}e_1^{}(h)e_2^{}(h)e_3^{}(h)\pi ^{}(𝔼_d)=1$$ The instanton numbers $`n_\gamma `$ are defined inductively by: (16) $$N_\beta =\underset{\beta =k\gamma }{}n_\gamma k^3$$ The point of this introduction is that the Aspinwall-Morrison calculation deals with concepts and questions that were not well defined at the time. Hence their calculation, although useful and convincing, is incomplete. The purpose of this paper is to relate the two calculations, hence justifying the Aspinwall-Morrison calculation and closing this historic chapter in the subject. We show in passing the connection between the two formulations of the multiple cover formula for the quintic threefold: (17) $$N_d(H,H,H)=d^3N_d=d^3\underset{k|d}{}n_k(\frac{k}{d})^3=\underset{k|d}{}n_kk^3$$ II. A review of the Aspinwall-Morrison calculation. Consider a Calabi-Yau threefold $`X`$ and a rational curve $`CX`$ such that $`N_{C/X}=𝒪_C(1)𝒪_C(1)`$. Let: (18) $$N_d(C):=\{f:^1X|f(^1)=C,\text{deg}f=d\}$$ be the space of parameterized maps from $`^1`$ to $`X`$. Since $`C`$ is isolated, $`N_d(C)`$ is a component of the space of all maps from $`^1`$ to $`X`$. At a moduli point $`[f]`$, the tangent space and the obstruction space are given respectively by $`H^0(f^{}(T_X))`$ and $`H^1(f^{}(T_X))`$, i.e. locally $`N_d(C)`$ is given by dim $`H^1(f^{}(T_X))`$ equations in the tangent space. The virtual dimension is: (19) $$\text{dim}H^0(f^{}(T_X))\text{dim}H^1(f^{}(T_X))=3.$$ The space $`N_d(C)`$ compactifies to $`\overline{N}_d(C)=^{2d+1}`$. Let $`\overline{\mathrm{\Gamma }}`$ be the compactification of the universal graph $`\mathrm{\Gamma }N_d(C)\times ^1\times C`$ and $`H`$ the hyperplane class in $`\overline{N}_d(C)`$.. The dimension of $`H^1(f^{}(T_X))`$ is $`2d2`$ for any $`f`$. These vector spaces fit together to form a bundle $`𝒰_d`$ over $`N_d(C)`$. Let $`p_i`$ be the $`i`$-th projection on $`\overline{N}_d(C)\times ^1\times C`$. The bundle $`𝒰_d`$ extends to: (20) $$U_d:=R^1p_{1}^{}{}_{}{}^{}(p_3^{}(T_X|C)|_{\overline{\mathrm{\Gamma }}})$$ over $`\overline{N}_d(C)`$. A calculation in yields $`U_d=𝒪(1)^{d2}`$. Based primarily on considerations from topological field theories, Aspinwall and Morrison asserted that the cycle corresponding to the degree $`d`$ multiple covers of $`C`$ is Poincar$`\stackrel{´}{e}`$ dual to $`c_{\text{top}}(U_d)=H^{2d2}`$. We will see that this is consistent with the notion of the virtual fundamental class. Let $`H_iH^2(X)`$ for $`i=1,2,3`$ and $`Z_i`$ their Poincar$`\stackrel{´}{e}`$ duals. The space: (21) $$\{fN_d(C)|f(0)=0\}$$ gives rise to a linear subspace of $`\overline{N}_d(C)`$. Therefore: $`\mathrm{\#}\{fN_d(C)|f(0)=0,f(1)=1,f(\mathrm{})=\mathrm{}\}`$ (22) $`={\displaystyle _{\overline{N}_d(C)}}HHHc_{\text{top}}U_d=1.`$ It follows that the contribution of $`N_d(C)`$ to: (23) $$\mathrm{\#}\{f:^1X|f_{}[^1]=d[C],f(0)Z_1,f(1)Z_2,f(\mathrm{})Z_3\}$$ is (24) $$_CH_1_CH_2_CH_3.$$ We emphasize that the multiple cover formula in this approach follows from: (25) $$_{\overline{N}_d(C)}HHHc_{\text{top}}U_d=_{\overline{N}_d(C)}H^{2d+1}=1.$$ III. The connection to the Gromov-Witten theory. The main result in this paper is the following: ###### Proposition 1.0.1. There exists a birational morphism: (26) $$\alpha :\overline{M}_{0,3}(C,d)\overline{N}_d(C)$$ such that: 1. $`\alpha _{}(e_i^{}(h))=H`$ for $`i=1,2,3.`$ 2. $`\alpha _{}(e_1^{}(h)e_2^{}(h)e_3^{}(h))=H^3`$ 3. $`\alpha _{}(e_1^{}(h)e_2^{}(h)e_3^{}(h)\pi ^{}(𝔼_d))=H^{2d+1}`$. This proposition implies that the equations $`(15)`$ and $`(25)`$ are equivalent, hence connecting the Aspinwall-Morrison calculation to the Gromov-Witten theory. Acknowledgements. The problem was suggested to the author by Sheldon Katz (see also the note in ) who was very helpful through this work. We would also like to thank Jun Li for fruitful discussions on the subject. ## 2. Relation of the Aspinwall-Morrison formula with Gromov-Witten invariants The space of nonparameterized degree $`d`$ maps $`f:^1^n`$ has two particular compactifications that have been employed successfully especially in proving mirror theorems for projective spaces: the nonlinear sigma model (or the graph space): (27) $$M_d^n:=\overline{M}_{0,0}(^n\times ^1,(d,1))$$ and the linear sigma model: (28) $$N_d^n:=(H^0(𝒪_^1(d))).$$ Elements of $`N_d^n`$ are $`(n+1)`$-tuples $`[P_0,\mathrm{},P_n]`$ of degree $`d`$ polynomials in two variables $`w_0,w_1`$. The linear sigma model $`N_d`$ is a projective space via the identification $`[P_0,\mathrm{}]=[_ia_iw_0^iw_1^{d_i},\mathrm{}]=[a_0,\mathrm{},a_d,\mathrm{}]`$. Note that $`N_d^1=\overline{N}_d(C)`$ for $`C^1`$. Let $`H`$ be the hyperplane class in $`N_d^n`$. There exists a birational morphism $`\varphi :M_d^nN_d^n`$. We describe this morphism set-theoretically. Let $`(C^{},f)M_d^n`$. There is a unique component $`C_0`$ of $`C^{}`$ that is mapped with degree $`1`$ to $`^1`$. Let $`C_1,\mathrm{},C_r`$ be the irreducible components of the rest of the curve and $`q_i=(c_i,d_i)`$ the nodes of $`C^{}`$ on $`C_0`$. Let $`di`$ be the degree of the map $`p_2f:C^{}^n`$ on $`C_i`$ for $`i=0,1,\mathrm{},r`$. Let $`R(w_0,w_1)=_{i=1}^r(c_iw_1d_iw_0)^{d_i}`$. If the restriction of the map $`p_2f`$ is given by $`[Q_0,\mathrm{},Q_n]`$ then: (29) $$\varphi (C^{},f):=[RQ_0,\mathrm{},RQ_n].$$ A proof of the fact that $`\varphi `$ is a morphism is given by J. Li in . The first step in connecting the Aspinwall-Morrison calculation to Gromov-Witten invariants is showing that $`M_d^n`$ and $`N_d^n`$ are birational models for $`\overline{M}_{0,3}(^n,d)`$. ###### Lemma 2.0.1. There exists a birational map $`\psi :\overline{M}_{0,3}(^n,d)M_d^n`$. Proof. Consider the following diagram: $$\begin{array}{ccc}\overline{M}_{0,4}(^n,d)& \stackrel{(\widehat{\pi },e_4)}{}& \overline{M}_{0,4}\times ^n\\ \pi _4& & \\ \overline{M}_{0,3}(^n,d).\end{array}$$ Since $`\overline{M}_{0,4}^1`$ and $`e_4`$ is stable in the fibers of $`\pi _4`$, the above diagram exhibits a stable family of maps of degree $`(1,d)`$ parametrized by $`\overline{M}_{0,3}(^n,d)`$. Universal properties of $`M_d^n`$ yield a morphism: (30) $$\psi :\overline{M}_{0,3}(^n,d)M_d^n.$$ The map $`\psi `$ is an isomorphism in the smooth locus, hence it is a birational map.$``$ Let $`\pi _4:\overline{M}_{0,4}\overline{M}_{0,3}=\{pt\}`$ be the map that forgets the last marked point and $`\sigma _i`$ be the section of the i-th marked point for $`i=1,2,3`$. Choose coordinates on $`\overline{M}_{0,4}^1`$ such that the images of these three sections are respectively $`0=[1,0],\mathrm{}=[0,1],1=[1,1]`$. Let (31) $$\alpha :=\varphi \psi :\overline{M}_{0,3}(^n,d)N_d^n.$$ ###### Proposition 2.0.1. Let $`h`$ be the hyperplane class of $`^n`$. 1. $`\alpha _{}(e_i^{}(h))=H`$ for $`i=1,2,3`$. 2. $`\alpha _{}(e_1^{}(h)e_2^{}(h)e_3^{}(h))=H^3`$ Proof. Let (32) $$\nu _1:N_d>^n$$ be a rational map defined by (33) $$\nu _1([P_0,P_1,\mathrm{},P_n])=[P_0(1,0),P_1(1,0),\mathrm{},P_n(1,0)].$$ This map is defined in the complement $`U`$ of a codimension $`n+1`$ linear subspace $`P(W_1)`$ of $`N_d^n`$. Clearly $`\nu _1^{}(h)=H`$ on $`U`$. The preimage $`D_{1,23}`$ of $`P(W_1)`$ in $`\overline{M}_{0,3}(^n,d)`$ is a sum of $`d`$ boundary divisors $`D(\{x_1\},\{x_2,x_3\},d_1,d_2)`$ with $`d_1>0`$ and $`d_1+d_2=d`$. The evaluation map $`e_1`$ over $`U`$ factors through the rational map $`\nu _1`$. It follows that (34) $$e_1^{}(h)=\alpha ^{}(H)+D_1,$$ where $`D_1`$<sup>1</sup><sup>1</sup>1It can be shown that $`D_1=_{d_1}d_1D(\{x_1\},\{x_2,x_3\},d_1,dd_1)`$ but this is not important in this paper. is a divisor supported in $`D_{1,23}`$. Using the evaluations at $`1`$ and $`\mathrm{}`$ on $`N_d^n`$, we obtain: (35) $$e_2^{}(h)=\alpha ^{}(H)+D_2$$ and (36) $$e_3^{}(h)=\alpha ^{}(H)+D_3,$$ where $`D_2`$ is a divisor supported in $`D_{2,13}`$ and $`D_3`$ is supported in $`D_{3,12}`$. The $`\psi `$-image of $`D(\{x_1\},\{x_2,x_3\},d_1,d_2)`$ does not detect the movement of the marking $`x_1`$ along its incident component, hence it is a codimension $`2`$ cycle in $`M_d^n`$. It follows that $`\psi _{}(D_1)=0`$. Similarly $`\psi _{}(D_2)=0`$ and $`\psi _{}(D_3)=0`$. Both $`\psi `$ and $`\varphi `$ are birational hence by the projection formula: (37) $$\alpha _{}(e_i^{}(h))=H$$ for $`i=1,2,3`$. Let $`D^{}D_{1,23},D^{\prime \prime }D_{2,13},D^{\prime \prime \prime }D_{3,12}`$ be irreducible boundary divisors. The intersection of any two of them either is $`0`$ or its image is a codimension $`4`$ cycle in $`M_d^n`$. It follows that: (38) $$\psi _{}(D^{}D^{\prime \prime })=\psi _{}(D^{}D^{\prime \prime \prime })=\psi _{}(D^{\prime \prime }D^{\prime \prime \prime })=0.$$ Notice also that: (39) $$D^{}D^{\prime \prime }D^{\prime \prime \prime }=0.$$ The projection formula yields: (40) $$\psi _{}(e_1^{}(h)e_2^{}(h)e_3^{}(h)=\psi _{}(\underset{i}{}(\psi ^{}(\varphi ^{}(H))+D_i))=\underset{i}{}(\varphi ^{}(H))=\varphi ^{}(H^3).$$ The lemma follows from the fact that $`\varphi `$ is a birational map.$``$ Return now to the case $`n=1`$ of our interest. Let $`\rho :M_d^1\overline{M}_{0,0}(C,d)`$ be the natural morphism. The composition: (41) $$\rho \psi :\overline{M}_{0,3}(C,d)\overline{M}_{0,0}(C,d)$$ is the map $`\pi `$ that forgets the $`3`$ marked points and stabilizes the source curve. Recall Kontsevich’s obstruction bundle $`V_d`$ on $`\overline{M}_{0,0}(C,d)`$. Its fiber is $`H^1(C^{},f^{}(𝒪(1)𝒪(1)))`$. Its top chern class is $`𝔼_d`$. We are now ready to exhibit the connection between the Aspinwall-Morrison calculation and Gromov-Witten invariants. ###### Proposition 2.0.2. $`\alpha _{}\left(e_1^{}(h)e_2^{}(h)e_3^{}(h)\pi ^{}(𝔼_d)\right)=H^{2d+1}.`$ Proof. Let $`E_d`$ be the top chern class of the bundle $`\rho ^{}(V_d)`$ on $`M_d^1`$. Recall from part II of the introduction that $`H^{2d2}`$ is the top chern class of the Aspinwall-Morrison obstruction bundle $`U_d`$ on $`N_d^1`$. It is shown in that $`\varphi _{}(E_d)=H^{2d2}`$. On the other hand $`\psi ^{}(E_d)=\pi ^{}(𝔼_d)`$. But $`\psi `$ is birational, hence by the projection formula $`\psi _{}(\pi ^{}(𝔼_d))=E_d`$. We compute: $`\alpha _{}({\displaystyle \underset{i}{}}e_i^{}(h)𝔼_d)=\alpha _{}({\displaystyle \underset{i}{}}e_i^{}(h)\psi ^{}(E_d))=\varphi _{}(\psi _{}({\displaystyle \underset{i}{}}e_i^{}(h))E_d)`$ (42) $`=\varphi _{}(\varphi ^{}(H^3)E_d)=H^3\varphi _{}(E_d)=H^3H^{2d2}=H^{2d+1}.`$ The proposition is proven.$``$ The last proposition yields: (43) $$_{\overline{M}_{0,3}(C,d)}\underset{i=1}{\overset{3}{}}e_i^{}(h)𝔼_d=_{\overline{N}_d(C)}\alpha _{}(\underset{i=1}{\overset{3}{}}e_i^{}(h)\psi ^{}(E_d))=_{\overline{N}_d(C)}H^{2d+1}=1,$$ i.e. the Aspinwall-Morrison calculation is a pushforward of Kontsevich’s calculation from $`\overline{M}_{0,3}(C,d)`$ to the projective space $`\overline{N}_d(C)`$. E-mail: elezi@math.stanford.edu Address: Department of Mathematics, Stanford University, Stanford CA, 94305.
warning/0004/quant-ph0004090.html
ar5iv
text
# Path Integral Methods and ApplicationsLectures given at Rencontres du Vietnam: VIth Vietnam School of Physics, Vung Tau, Vietnam, 27 December 1999 - 8 January 2000. ## 1 Introduction ### 1.1 Historical remarks We are all familiar with the standard formulations of quantum mechanics, developed more or less concurrently by Schroedinger, Heisenberg and others in the 1920s, and shown to be equivalent to one another soon thereafter. In 1933, Dirac made the observation that the action plays a central role in classical mechanics (he considered the Lagrangian formulation of classical mechanics to be more fundamental than the Hamiltonian one), but that it seemed to have no important role in quantum mechanics as it was known at the time. He speculated on how this situation might be rectified, and he arrived at the conclusion that (in more modern language) the propagator in quantum mechanics “corresponds to” $`\mathrm{exp}iS/\mathrm{}`$, where $`S`$ is the classical action evaluated along the classical path. In 1948, Feynman developed Dirac’s suggestion, and succeeded in deriving a third formulation of quantum mechanics, based on the fact that the propagator can be written as a sum over all possible paths (not just the classical one) between the initial and final points. Each path contributes $`\mathrm{exp}iS/\mathrm{}`$ to the propagator. So while Dirac considered only the classical path, Feynman showed that all paths contribute: in a sense, the quantum particle takes all paths, and the amplitudes for each path add according to the usual quantum mechanical rule for combining amplitudes. Feynman’s original paper,<sup>1</sup><sup>1</sup>1References are not cited in the text, but a short list of books and articles which I have found interesting and useful is given at the end of this article. which essentially laid the foundation of the subject (and which was rejected by Physical Review!), is an all-time classic, and is highly recommended. (Dirac’s original article is not bad, either.) ### 1.2 Motivation What do we learn from path integrals? As far as I am aware, path integrals give us no dramatic new results in the quantum mechanics of a single particle. Indeed, most if not all calculations in quantum mechaincs which can be done by path integrals can be done with considerably greater ease using the standard formulations of quantum mechanics. (It is probably for this reason that path integrals are often left out of undergraduate-level quantum mechanics courses.) So why the fuss? As I will mention shortly, path integrals turn out to be considerably more useful in more complicated situations, such as field theory. But even if this were not the case, I believe that path integrals would be a very worthwhile contribution to our understanding of quantum mechanics. Firstly, they provide a physically extremely appealing and intuitive way of viewing quantum mechanics: anyone who can understand Young’s double slit experiment in optics should be able to understand the underlying ideas behind path integrals. Secondly, the classical limit of quantum mechanics can be understood in a particularly clean way via path integrals. It is in quantum field theory, both relativistic and nonrelativistic, that path integrals (functional integrals is a more accurate term) play a much more important role, for several reasons. They provide a relatively easy road to quantization and to expressions for Green’s functions, which are closely related to amplitudes for physical processes such as scattering and decays of particles. The path integral treatment of gauge field theories (non-abelian ones, in particular) is very elegant: gauge fixing and ghosts appear quite effortlessly. Also, there are a whole host of nonperturbative phenomena such as solitons and instantons that are most easily viewed via path integrals. Furthermore, the close relation between statistical mechanics and quantum mechanics, or statistical field theory and quantum field theory, is plainly visible via path integrals. In these lectures, I will not have time to go into great detail into the many useful applications of path integrals in quantum field theory. Rather than attempting to discuss a wide variety of applications in field theory and condensed matter physics, and in so doing having to skimp on the ABCs of the subject, I have chosen to spend perhaps more time and effort than absolutely necessary showing path integrals in action (pardon the pun) in quantum mechanics. The main emphasis will be on quantum mechanical problems which are not necessarily interesting and useful in and of themselves, but whose principal value is that they resemble the calculation of similar objects in the more complex setting of quantum field theory, where explicit calculations would be much harder. Thus I hope to illustrate the main points, and some technical complications and hangups which arise, in relatively familiar situations that should be regarded as toy models analogous to some interesting contexts in field theory. ### 1.3 Outline The outline of the lectures is as follows. In the next section I will begin with an introduction to path integrals in quantum mechanics, including some explicit examples such as the free particle and the harmonic oscillator. In Section 3, I will give a “derivation” of classical mechanics from quantum mechanics. In Section 4, I will discuss some applications of path integrals that are perhaps not so well-known, but nonetheless very amusing, namely, the case where the configuration space is not simply connected. (In spite of the fancy terminology, no prior knowledge of high-powered mathematics such as topology is assumed.) Specifically, I will apply the method to the Aharonov-Bohm effect, quantum statistics and anyons, and monopoles and charge quantization, where path integrals provide a beautifully intuitive approach. In Section 5, I will explain how one can approach statistical mechanics via path integrals. Next, I will discuss perturbation theory in quantum mechanics, where the technique used is (to put it mildly) rather cumbersome, but nonetheless illustrative for applications in the remaining sections. In Section 7, I will discuss Green’s functions (vacuum expectation values of time-ordered products) in quantum mechanics (where, to my knowledge, they are not particularly useful), and will construct the generating functional for these objects. This groundwork will be put to good use in the following section, where the generating functional for Green’s functions in field theory (which are useful!) will be elucidated. In Section 9, I will discuss instantons in quantum mechanics, and will at least pay lip service to important applications in field theory. I will finish with a summary and a list of embarrassing omissions. I will conclude with a few apologies. First, an educated reader might get the impression that the outline given above contains for the most part standard material. S/he is likely correct: the only original content to these lectures is the errors.<sup>2</sup><sup>2</sup>2Even this joke is borrowed from somewhere, though I can’t think of where. Second, I have made no great effort to give complete references (I know my limitations); at the end of this article I have listed some papers and books from which I have learned the subject. Some are books or articles wholly devoted to path integrals; the majority are books for which path integrals form only a small (but interesting!) part. The list is hopelessly incomplete; in particular, virtually any quantum field theory book from the last decade or so has a discussion of path integrals in it. Third, the subject of path integrals can be a rather delicate one for the mathematical purist. I am not one, and I have neither the interest nor the expertise to go into detail about whether or not the path integral exists, in a strict sense. My approach is rather pragmatic: it works, so let’s use it! ## 2 Path Integrals in Quantum Mechanics ### 2.1 General discussion Consider a particle moving in one dimension, the Hamiltonian being of the usual form: $$H=\frac{p^2}{2m}+V(q).$$ The fundamental question in the path integral (PI) formulation of quantum mechanics is: If the particle is at a position $`q`$ at time $`t=0`$, what is the probability amplitude that it will be at some other position $`q^{}`$ at a later time $`t=T`$? It is easy to get a formal expression for this amplitude in the usual Schroedinger formulation of quantum mechanics. Let us introduce the eigenstates of the position operator $`\widehat{q}`$, which form a complete, orthonormal set: $$\widehat{q}|q=q|q,q^{}|q=\delta (q^{}q),𝑑q|qq|=1.$$ (When there is the possibility of an ambiguity, operators will be written with a “hat”; otherwise the hat will be dropped.) Then the initial state is $`|\psi (0)=|q`$. Letting the state evolve in time and projecting on the state $`|q^{}`$, we get for the amplitude $`A`$, $$A=q^{}|\psi (T)K(q^{},T;q,0)=q^{}\left|e^{iHT}\right|q.$$ (1) (Except where noted otherwise, $`\mathrm{}`$ will be set to 1.) This object, for obvious reasons, is known as the propagator from the initial spacetime point $`(q,0)`$ to the final point $`(q^{},T)`$. Clearly, the propagator is independent of the origin of time: $`K(q^{},T+t;q,t)=K(q^{},T;q,0)`$. We will derive an expression for this amplitude in the form of a summation (integral, really) over all possible paths between the initial and final points. In so doing, we derive the PI from quantum mechanics. Historically, Feynman came up with the PI differently, and showed its equivalence to the usual formulations of quantum mechanics. Let us separate the time evolution in the above amplitude into two smaller time evolutions, writing $`e^{iHT}=e^{iH(Tt_1)}e^{iHt_1}`$. The amplitude becomes $$A=q^{}\left|e^{iH(Tt_1)}e^{iHt_1}\right|q.$$ Inserting a factor 1 in the form of a sum over the position eigenstates gives $`A`$ $`=`$ $`q^{}\left|e^{iH(Tt_1)}\underset{=1}{\underset{}{{\displaystyle 𝑑q_1|q_1q_1|}}}e^{iHt_1}\right|q`$ (2) $`=`$ $`{\displaystyle 𝑑q_1K(q^{},T;q_1,t_1)K(q_1,t_1;q,0)}.`$ This formula is none other than an expression of the quantum mechanical rule for combining amplitudes: if a process can occur a number of ways, the amplitudes for each of these ways add. A particle, in propagating from $`q`$ to $`q^{}`$, must be somewhere at an intermediate time $`t_1`$; labelling that intermediate position $`q_1`$, we compute the amplitude for propagation via the point $`q_1`$ \[this is the product of the two propagators in (2)\] and integrate over all possible intermediate positions. This result is reminiscent of Young’s double slit experiment, where the amplitudes for passing through each of the two slits combine and interfere. We will look at the double-slit experiment in more detail when we discuss the Aharonov-Bohm effect in Section 4. We can repeat the division of the time interval $`T`$; let us divide it up into a large number $`N`$ of time intervals of duration $`\delta =T/N`$. Then we can write for the propagator $$A=q^{}\left|\left(e^{iH\delta }\right)^N\right|q=q^{}\left|\underset{N\mathrm{times}}{\underset{}{e^{iH\delta }e^{iH\delta }\mathrm{}e^{iH\delta }}}\right|q.$$ We can again insert a complete set of states between each exponential, yielding $`A`$ $`=`$ $`q^{}\left|e^{iH\delta }{\displaystyle 𝑑q_{N1}}\right|q_{N1}q_{N1}\left|e^{iH\delta }{\displaystyle 𝑑q_{N2}}\right|q_{N2}q_{N2}|\mathrm{}`$ (3) $`\mathrm{}{\displaystyle 𝑑q_2|q_2q_2\left|e^{iH\delta }𝑑q_1\right|q_1q_1\left|e^{iH\delta }\right|q}`$ $`=`$ $`{\displaystyle 𝑑q_1\mathrm{}𝑑q_{N1}q^{}\left|e^{iH\delta }\right|q_{N1}q_{N1}\left|e^{iH\delta }\right|q_{N2}\mathrm{}}`$ $`\mathrm{}q_1\left|e^{iH\delta }\right|q`$ $``$ $`{\displaystyle 𝑑q_1\mathrm{}𝑑q_{N1}K_{q_N,q_{N1}}K_{q_{N1},q_{N2}}\mathrm{}K_{q_2,q_1}K_{q_1,q_0}},`$ where we have defined $`q_0=q`$, $`q_N=q^{}`$. (Note that these initial and final positions are not integrated over.) This expression says that the amplitude is the integral of the amplitude of all $`N`$-legged paths, as illustrated in Figure 1. Apart from mathematical details concerning the limit when $`N\mathrm{}`$, this is clearly going to become a sum over all possible paths of the amplitude for each path: $$A=\underset{\mathrm{paths}}{}A_{\mathrm{path}},$$ where $$\underset{\mathrm{paths}}{}=𝑑q_1\mathrm{}𝑑q_{N1},A_{\mathrm{path}}=K_{q_N,q_{N1}}K_{q_{N1},q_{N2}}\mathrm{}K_{q_2,q_1}K_{q_1,q_0}.$$ Let us look at this last expression in detail. The propagator for one sub-interval is $`K_{q_{j+1},q_j}=q_{j+1}\left|e^{iH\delta }\right|q_j`$. We can expand the exponential, since $`\delta `$ is small: $`K_{q_{j+1},q_j}`$ $`=`$ $`q_{j+1}\left|\left(1iH\delta {\displaystyle \frac{1}{2}}H^2\delta ^2+\mathrm{}\right)\right|q_j`$ (4) $`=`$ $`q_{j+1}|q_ji\delta q_{j+1}\left|H\right|q_j+o(\delta ^2).`$ The first term is a delta function, which we can write<sup>3</sup><sup>3</sup>3Please do not confuse the delta function with the time interval, $`\delta `$. $$q_{j+1}|q_j=\delta (q_{j+1}q_j)=\frac{dp_j}{2\pi }e^{ip_j(q_{j+1}q_j)}.$$ (5) In the second term of (4), we can insert a factor 1 in the form of an integral over momentum eigenstates between $`H`$ and $`|q_j`$; this gives $`i\delta q_{j+1}\left|\left({\displaystyle \frac{\widehat{p}^2}{2m}}+V(\widehat{q})\right){\displaystyle \frac{dp_j}{2\pi }}\right|p_jp_j|q_j`$ $`=i\delta {\displaystyle \frac{dp_j}{2\pi }\left(\frac{p_{j}^{}{}_{}{}^{2}}{2m}+V(q_{j+1})\right)q_{j+1}|p_jp_j|q_j}`$ $`=i\delta {\displaystyle \frac{dp_j}{2\pi }\left(\frac{p_{j}^{}{}_{}{}^{2}}{2m}+V(q_{j+1})\right)e^{ip_j(q_{j+1}q_j)}},`$ (6) using $`q|p=\mathrm{exp}ipq`$. In the first line, we view the operator $`\widehat{p}`$ as operating to the right, while $`V(\widehat{q})`$ operates to the left. The expression (6) is asymmetric between $`q_j`$ and $`q_{j+1}`$; the origin of this is our choice of putting the factor 1 to the right of $`H`$ in the second term of (4). Had we put it to the left instead, we would have obtained $`V(q_j)`$ in (6). To not play favourites, we should choose some sort of average of these two. In what follows I will simply write $`V(\overline{q}_j)`$ where $`\overline{q}_j=\frac{1}{2}(q_j+q_{j+1})`$. (The exact choice does not matter in the continuum limit, which we will take eventually; the above is a common choice.) Combining (5) and (6), the sub-interval propagator is $`K_{q_{j+1},q_j}`$ $`=`$ $`{\displaystyle \frac{dp_j}{2\pi }e^{ip_j(q_{j+1}q_j)}\left(1i\delta \left(\frac{p_{j}^{}{}_{}{}^{2}}{2m}+V(\overline{q}_j)\right)+o(\delta ^2)\right)}`$ (7) $`=`$ $`{\displaystyle \frac{dp_j}{2\pi }e^{ip_j(q_{j+1}q_j)}e^{i\delta H(p_j,\overline{q}_j)}(1+o(\delta ^2))}.`$ There are $`N`$ such factors in the amplitude. Combining them, and writing $`\dot{q}_j=(q_{j+1}q_j)/\delta `$, we get $$A_{\mathrm{path}}=\underset{j=0}{\overset{N1}{}}\frac{dp_j}{2\pi }\mathrm{exp}i\delta \underset{j=0}{\overset{N1}{}}(p_j\dot{q}_jH(p_j,\overline{q}_j)),$$ (8) where we have neglected a multiplicative factor of the form $`(1+o(\delta ^2))^N`$, which will tend toward one in the continuum limit. Then the propagator becomes $`K`$ $`=`$ $`{\displaystyle 𝑑q_1\mathrm{}𝑑q_{N1}A_{\mathrm{path}}}`$ (9) $`=`$ $`{\displaystyle \underset{j=1}{\overset{N1}{}}dq_j\underset{j=0}{\overset{N1}{}}\frac{dp_j}{2\pi }\mathrm{exp}i\delta \underset{j=0}{\overset{N1}{}}(p_j\dot{q}_jH(p_j,\overline{q}_j))}.`$ Note that there is one momentum integral for each interval ($`N`$ total), while there is one position integral for each intermediate position ($`N1`$ total). If $`N\mathrm{}`$, this approximates an integral over all functions $`p(t)`$, $`q(t)`$. We adopt the following notation: $$\overline{)K𝒟p(t)𝒟q(t)\mathrm{exp}i_0^Tdt(p\dot{q}H(p,q)).}$$ (10) This result is known as the phase-space path integral. The integral is viewed as over all functions $`p(t)`$ and over all functions $`q(t)`$ where $`q(0)=q`$, $`q(T)=q^{}`$. But to actually perform an explicit calculation, (10) should be viewed as a shorthand notation for the more ponderous expression (9), in the limit $`N\mathrm{}`$. If, as is often the case (and as we have assumed in deriving the above expression), the Hamiltonian is of the standard form, namely $`H=p^2/2m+V(q)`$, we can actually carry out the momentum integrals in (9). We can rewrite this expression as $$K=\underset{j=1}{\overset{N1}{}}dq_j\mathrm{exp}i\delta \underset{j=0}{\overset{N1}{}}V(\overline{q}_j)\underset{j=0}{\overset{N1}{}}\frac{dp_j}{2\pi }\mathrm{exp}i\delta \underset{j=0}{\overset{N1}{}}\left(p_j\dot{q}_jp_{j}^{}{}_{}{}^{2}/2m\right).$$ The $`p`$ integrals are all Gaussian, and they are uncoupled. One such integral is $$\frac{dp}{2\pi }e^{i\delta (p\dot{q}p^2/2m)}=\sqrt{\frac{m}{2\pi i\delta }}e^{i\delta m\dot{q}^2/2}.$$ (The careful reader may be worried about the convergence of this integral; if so, a factor $`\mathrm{exp}ϵp^2`$ can be introduced and the limit $`ϵ0`$ taken at the end.) The propagator becomes $`K`$ $`=`$ $`{\displaystyle \underset{j=1}{\overset{N1}{}}dq_j\mathrm{exp}}i\delta {\displaystyle \underset{j=0}{\overset{N1}{}}}V(\overline{q}_j){\displaystyle \underset{j=0}{\overset{N1}{}}}\left(\sqrt{{\displaystyle \frac{m}{2\pi i\delta }}}\mathrm{exp}i\delta {\displaystyle \frac{m\dot{q}_j^2}{2}}\right)`$ (11) $`=`$ $`\left({\displaystyle \frac{m}{2\pi i\delta }}\right)^{N/2}{\displaystyle \underset{j=1}{\overset{N1}{}}dq_j\mathrm{exp}i\delta \underset{j=0}{\overset{N1}{}}\left(\frac{m\dot{q}_j^2}{2}V(\overline{q}_j)\right)}.`$ The argument of the exponential is a discrete approximation of the action of a path passing through the points $`q_0=q,q_1,\mathrm{},q_{N1},q_N=q^{}`$. As above, we can write this in the more compact form $$\overline{)K=𝒟q(t)e^{iS[q(t)]}.}$$ (12) This is our final result, and is known as the configuration space path integral. Again, (12) should be viewed as a notation for the more precise expression (11), as $`N\mathrm{}`$. ### 2.2 Examples To solidify the notions above, let us consider a few explicit examples. As a first example, we will compute the free particle propagator first using ordinary quantum mechanics and then via the PI. We will then mention some generalizations which can be done in a similar manner. #### 2.2.1 Free particle Let us compute the propagator $`K(q^{},T;q,0)`$ for a free particle, described by the Hamiltonian $`H=p^2/2m`$. The propagator can be computed straightforwardly using ordinary quantum mechanics. To this end, we write $`K`$ $`=`$ $`q^{}\left|e^{iHT}\right|q`$ (13) $`=`$ $`q^{}\left|e^{iT\widehat{p}^2/2m}{\displaystyle \frac{dp}{2\pi }}\right|pp|q`$ $`=`$ $`{\displaystyle \frac{dp}{2\pi }e^{iTp^2/2m}q^{}|pp|q}`$ $`=`$ $`{\displaystyle \frac{dp}{2\pi }e^{iT(p^2/2m)+i(q^{}q)p}}.`$ The integral is Gaussian; we obtain $$K=\left(\frac{m}{2\pi iT}\right)^{1/2}e^{im(q^{}q)^2/2T}.$$ (14) Let us now see how the same result can be attained using PIs. The configuration space PI (12) is $`K`$ $`=`$ $`\underset{N\mathrm{}}{lim}\left({\displaystyle \frac{m}{2\pi i\delta }}\right)^{N/2}{\displaystyle \underset{j=1}{\overset{N1}{}}dq_j\mathrm{exp}i\frac{m\delta }{2}\underset{j=0}{\overset{N1}{}}\left(\frac{q_{j+1}q_j}{\delta }\right)^2}`$ $`=`$ $`\underset{N\mathrm{}}{lim}\left({\displaystyle \frac{m}{2\pi i\delta }}\right)^{N/2}{\displaystyle }{\displaystyle \underset{j=1}{\overset{N1}{}}}dq_j\mathrm{exp}i{\displaystyle \frac{m}{2\delta }}[(q_Nq_{N1})^2+(q_{N1}q_{N2})^2+\mathrm{}`$ $`+(q_2q_1)^2+(q_1q_0)^2],`$ where $`q_0=q`$ and $`q_N=q^{}`$ are the initial and final points. The integrals are Gaussian, and can be evaluated exactly, although the fact that they are coupled complicates matters significantly. The result is $`K`$ $`=`$ $`\underset{N\mathrm{}}{lim}\left({\displaystyle \frac{m}{2\pi i\delta }}\right)^{N/2}{\displaystyle \frac{1}{\sqrt{N}}}\left({\displaystyle \frac{2\pi i\delta }{m}}\right)^{(N1)/2}e^{im(q^{}q)^2/2N\delta }`$ $`=`$ $`\underset{N\mathrm{}}{lim}\left({\displaystyle \frac{m}{2\pi iN\delta }}\right)^{1/2}e^{im(q^{}q)^2/2N\delta }.`$ But $`N\delta `$ is the total time interval $`T`$, resulting in $$K=\left(\frac{m}{2\pi iT}\right)^{1/2}e^{im(q^{}q)^2/2T},$$ in agreement with (14). A couple of remarks are in order. First, we can write the argument of the exponential as $`T\frac{1}{2}m((q^{}q)/T)^2`$, which is just the action $`S[q_c]`$ for a particle moving along the classical path (a straight line in this case) between the initial and final points. Secondly, we can restore the factors of $`\mathrm{}`$ if we want, by ensuring correct dimensions. The argument of the exponential is the action, so in order to make it a pure number we must divide by $`\mathrm{}`$; furthermore, the propagator has the dimension of the inner product of two position eigenstates, which is inverse length; in order that the coefficient have this dimension we must multiply by $`\mathrm{}^{1/2}`$. The final result is $$\overline{)K=\left(\frac{m}{2\pi i\mathrm{}T}\right)^{1/2}e^{iS[q_c]/\mathrm{}}.}$$ (15) This result typifies a couple of important features of calculations in this subject, which we will see repeatedly in these lectures. First, the propagator separates into two factors, one of which is the phase $`\mathrm{exp}iS[q_c]/\mathrm{}`$. Second, calculations in the PI formalism are typically quite a bit more lengthy than using standard techniques of quantum mechanics. #### 2.2.2 Harmonic oscillator As a second example of the computation of a PI, let us compute the propagator for the harmonic oscillator using this method. (In fact, we will not do the entire computation, but we will do enough to illustrate a trick or two which will be useful later on.) Let us start with the somewhat-formal version of the configuration-space PI, (12): $$K(q^{},T;q,0)=𝒟q(t)e^{iS[q(t)]}.$$ For the harmonic oscillator, $$S[q(t)]=_0^T𝑑t\left(\frac{1}{2}m\dot{q}^2\frac{1}{2}m\omega ^2q^2\right).$$ The paths over which the integral is to be performed go from $`q(0)=q`$ to $`q(T)=q^{}`$. To do this PI, suppose we know the solution of the classical problem, $`q_c(t)`$: $$\ddot{q}_c+\omega ^2q_c=0,q_c(0)=q,q_c(T)=q^{}.$$ We can write $`q(t)=q_c(t)+y(t)`$, and perform a change of variables in the PI to $`y(t)`$, since integrating over all deviations from the classical path is equivalent to integrating over all possible paths. Since at each time $`q`$ and $`y`$ differ by a constant, the Jacobian of the transformation is 1. Furthermore, since $`q_c`$ obeys the correct boundary conditions, the paths $`y(t)`$ over which we integrate go from $`y(0)=0`$ to $`y(T)=0`$. The action for the path $`q_c(t)+y(t)`$ can be written as a power series in $`y`$: $`S[q_c(t)+y(t)]`$ $`=`$ $`{\displaystyle _0^T}𝑑t\left({\displaystyle \frac{1}{2}}m\dot{q}_{c}^{}{}_{}{}^{2}{\displaystyle \frac{1}{2}}m\omega ^2q_{c}^{}{}_{}{}^{2}\right)+\underset{=0}{\underset{}{(\text{linear in }y)}}+{\displaystyle _0^T}𝑑t\left({\displaystyle \frac{1}{2}}m\dot{y}^2{\displaystyle \frac{1}{2}}m\omega ^2y^2\right).`$ The term linear in $`y`$ vanishes by construction: $`q_c`$, being the classical path, is that path for which the action is stationary! So we may write $`S[q_c(t)+y(t)]=S[q_c(t)]+S[y(t)]`$. We substitute this into (12), yielding $$K(q^{},T;q,0)=e^{iS[q_c(t)]}𝒟y(t)e^{iS[y(t)]}.$$ (16) As mentioned above, the paths $`y(t)`$ over which we integrate go from $`y(0)=0`$ to $`y(T)=0`$: the only appearance of the initial and final positions is in the classical path, i.e., in the classical action. Once again, the PI separates into two factors. The first is written in terms of the action of the classical path, and the second is a PI over deviations from this classical path. The second factor is independent of the initial and final points. This separation into a factor depending on the action of the classical path and a second one, a PI which is independent of the details of the classical path, is a recurring theme, and an important one. Indeed, it is often the first factor which contains most of the useful information contained in the propagator, and it can be deduced without even performing a PI. It can be said that much of the work in the game of path integrals consists in avoiding having to actually compute one! As for the evaluation of (16), a number of fairly standard techniques are available. One can calculate the PI directly in position space, as was done above for the harmonic oscillator (see Schulman, chap. 6). Alternatively, one can compute it in Fourier space (writing $`y(t)=_ka_k\mathrm{sin}(k\pi t/T)`$ and integrating over the coefficients $`\{a_k\}`$). This latter approach is outlined in Feynman and Hibbs, Section 3.11. The result is $$K(q^{},T;q,0)=\left(\frac{m\omega }{2\pi i\mathrm{sin}\omega T}\right)^{1/2}e^{iS[q_c(t)]}.$$ (17) The classical action can be evaluated straightforwardly (note that this is not a PI problem, nor even a quantum mechanics problem!); the result is $$S[q_c(t)]=\frac{m\omega }{2\mathrm{sin}\omega T}\left((q_{}^{}{}_{}{}^{2}+q^2)\mathrm{cos}\omega T2q^{}q\right).$$ We close this section with two remarks. First, the PI for any quadratic action can be evaluated exactly, essentially since such a PI consists of Gaussian integrals; the general result is given in Schulman, Chapter 6. In Section 6, we will evaluate (to the same degree of completeness as the harmonic oscillator above) the PI for a forced harmonic oscillator, which will prove to be a very useful tool for computing a variety of quantities of physical interest. Second, the following fact is not difficult to prove, and will be used below (Section 4.2.). $`K(q^{},T;q,0)`$ (whether computed via PIs or not) is the amplitude to propagate from one point to another in a given time interval. But this is the response to the following question: If a particle is initially at position $`q`$, what is its wave function after the elapse of a time $`T`$? Thus, if we consider $`K`$ as a function of the final position and time, it is none other than the wave function for a particle with a specific initial condition. As such, the propagator satisfies the Schroedinger equation at its final point. ## 3 The Classical Limit: “Derivation” of the Principle of Least Action Since the example calculations performed above are somewhat dry and mathematical, it is worth backing up a bit and staring at the expression for the configuration space PI, (12): $$K=𝒟q(t)e^{iS[q(t)]/\mathrm{}}.$$ This innocent-looking expression tells us something which is at first glance unbelievable, and at second glance really unbelievable. The first-glance observation is that a particle, in going from one position to another, takes all possible paths between these two positions. This is, if not actually unbelievable, at the very least least counter-intuitive, but we could argue away much of what makes us feel uneasy if we could convince ourselves that while all paths contribute, the classical path is the dominant one. However, the second-glance observation is not reassuring: if we compare the contribution of the classical path (whose action is $`S[q_c]`$) with that of some other, arbitrarily wild, path (whose action is $`S[q_w]`$), we find that the first is $`\mathrm{exp}iS[q_c]`$ while the second is $`\mathrm{exp}iS[q_w]`$. They are both complex numbers of unit magnitude: each path taken in isolation is equally important. The classical path is no more important than any arbitrarily complicated path! How are we to reconcile this really unbelievable conclusion with the fact that a ball thrown in the air has a more-or-less parabolic motion? The key, not surprisingly, is in how different paths interfere with one another, and by considering the case where the rough scale of classical action of the problem is much bigger than the quantum of action, $`\mathrm{}`$, we will see the emergence of the Principle of Least Action. Consider two neighbouring paths $`q(t)`$ and $`q^{}(t)`$ which contribute to the PI (Figure 2). Let $`q^{}(t)=q(t)+\eta (t)`$, with $`\eta (t)`$ small. Then we can write the action as a functional Taylor expansion about the classical path:<sup>4</sup><sup>4</sup>4The reader unfamiliar with manipulation of functionals need not despair; the only rule needed beyond standard calculus is the functional derivative: $`\delta q(t)/\delta q(t^{})=\delta (tt^{})`$, where the last $`\delta `$ is the Dirac delta function. $$S[q^{}]=S[q+\eta ]=S[q]+𝑑t\eta (t)\frac{\delta S[q]}{\delta q(t)}+o(\eta ^2).$$ The two paths contribute $`\mathrm{exp}iS[q]/\mathrm{}`$ and $`\mathrm{exp}iS[q^{}]/\mathrm{}`$ to the PI; the combined contribution is $$Ae^{iS[q]/\mathrm{}}\left(1+\mathrm{exp}\frac{i}{\mathrm{}}𝑑t\eta (t)\frac{\delta S[q]}{\delta q(t)}\right),$$ where we have neglected corrections of order $`\eta ^2`$. We see that the difference in phase between the two paths, which determines the interference between the two contributions, is $`\mathrm{}^1𝑑t\eta (t)\delta S[q]/\delta q(t)`$. We see that the smaller the value of $`\mathrm{}`$, the larger the phase difference between two given paths. So even if the paths are very close together, so that the difference in actions is extremely small, for sufficiently small $`\mathrm{}`$ the phase difference will still be large, and on average destructive interference occurs. However, this argument must be rethought for one exceptional path: that which extremizes the action, i.e., the classical path, $`q_c(t)`$. For this path, $`S[q_c+\eta ]=S[q_c]+o(\eta ^2)`$. Thus the classical path and a very close neighbour will have actions which differ by much less than two randomly-chosen but equally close paths (Figure 3). This means that for fixed closeness of two paths (I leave it as an exercise to make this precise!) and for fixed $`\mathrm{}`$, paths near the classical path will on average interfere constructively (small phase difference) whereas for random paths the interference will be on average destructive. Thus heuristically, we conclude that if the problem is classical (action $`\mathrm{}`$), the most important contribution to the PI comes from the region around the path which extremizes the PI. In other words, the particle’s motion is governed by the principle that the action is stationary. This, of course, is none other than the Principle of Least Action from which the Euler-Lagrange equations of classical mechanics are derived. ## 4 Topology and Path Integrals in Quantum Mechanics: Three Applications In path integrals, if the configuration space has holes in it such that two paths between the same initial and final point are not necessarily deformable into one another, interesting effects can arise. This property of the configuration space goes by the following catchy name: non-simply-connectedness. We will study three such situations: the Aharonov-Bohm effect, particle statistics, and magnetic monopoles and the quantization of electric charge. ### 4.1 Aharonov-Bohm effect The Aharonov-Bohm effect is one of the most dramatic illustrations of a purely quantum effect: the influence of the electromagnetic potential on particle motion even if the particle is perfectly shielded from any electric or magnetic fields. While classically the effect of electric and magnetic fields can be understood purely in terms of the forces these fields create on particles, Aharonov and Bohm devised an ingenious thought-experiment (which has since been realized in the laboratory) showing that this is no longer true in quantum mechanics. Their effect is best illustrated by a refinement of Young’s double-slit experiment, where particles passing through a barrier with two slits in it produce an interference pattern on a screen further downstream. Aharonov and Bohm proposed such an experiment performed with charged particles, with an added twist provided by a magnetic flux from which the particles are perfectly shielded passing between the two slits. If we perform the experiment first with no magnetic flux and then with a nonzero and arbitrary flux passing through the shielded region, the interference pattern will change, in spite of the fact that the particles are perfectly shielded from the magnetic field and feel no electric or magnetic force whatsoever. Classically we can say: no force, no effect. Not so in quantum mechanics. PIs provide a very attractive way of understanding this effect. Consider first two representative paths $`𝐪_1(t)`$ and $`𝐪_2(t)`$ (in two dimensions) passing through slits 1 and 2, respectively, and which arrive at the same spot on the screen (Figure 5). Before turning on the magnetic field, let us suppose that the actions for these paths are $`S[𝐪_1]`$ and $`S[𝐪_2]`$. Then the interference of the amplitudes is determined by $$e^{iS[𝐪_1]/\mathrm{}}+e^{iS[𝐪_2]/\mathrm{}}=e^{iS[𝐪_1]/\mathrm{}}\left(1+e^{i(S[𝐪_2]S[𝐪_1])/\mathrm{}}\right).$$ The relative phase is $`\varphi _{12}(S[𝐪_2]S[𝐪_1])/\mathrm{}`$. Thus these two paths interfere constructively if $`\varphi _{12}=2n\pi `$, destructively if $`\varphi _{12}=(2n+1)\pi `$, and in general there is partial cancellation between the two contributions. How is this result affected if we add a magnetic field, $`𝐁`$? We can describe this field by a vector potential, writing $`𝐁=\times 𝐀`$. This affects the particle’s motion by the following change in the Lagrangian: $$L(\dot{𝐪},𝐪)L^{}(\dot{𝐪},𝐪)=L(\dot{𝐪},𝐪)\frac{e}{c}𝐯𝐀(𝐪).$$ Thus the action changes by $$\frac{e}{c}𝑑t𝐯𝐀(𝐪)=\frac{e}{c}𝑑t\frac{d𝐪(t)}{dt}𝐀(𝐪(t)).$$ This integral is $`𝑑𝐪𝐀(𝐪)`$, the line integral of $`𝐀`$ along the path taken by the particle. So including the effect of the magnetic field, the action of the first path is $$S^{}[𝐪_1]=S[𝐪_1]\frac{e}{c}_{𝐪_1(t)}𝑑𝐪𝐀(𝐪),$$ and similarly for the second path. Let us now look at the interference between the two paths, including the magnetic field. $`e^{iS^{}[𝐪_1]/\mathrm{}}+e^{iS^{}[𝐪_2]/\mathrm{}}`$ $`=`$ $`e^{iS^{}[𝐪_1]/\mathrm{}}\left(1+e^{i(S^{}[𝐪_2]S^{}[𝐪_1])/\mathrm{}}\right)`$ (18) $`=`$ $`e^{iS^{}[𝐪_1]/\mathrm{}}\left(1+e^{i\varphi _{12}^{}}\right),`$ where the new relative phase is $$\varphi _{12}^{}=\varphi _{12}\frac{e}{\mathrm{}c}\left(_{𝐪_2(t)}𝑑𝐪𝐀(𝐪)_{𝐪_1(t)}𝑑𝐪𝐀(𝐪)\right).$$ (19) But the difference in line integrals in (19) is a contour integral: $$_{𝐪_2(t)}𝑑𝐪𝐀(𝐪)_{𝐪_1(t)}𝑑𝐪𝐀(𝐪)=𝑑𝐪𝐀(𝐪)=\mathrm{\Phi },$$ $`\mathrm{\Phi }`$ being the flux inside the closed loop bounded by the two paths. So we can write $$\varphi _{12}^{}=\varphi _{12}\frac{e\mathrm{\Phi }}{\mathrm{}c}.$$ It is important to note that the change of relative phase due to the magnetic field is independent of the details of the two paths, as long as each passes through the corresponding slit. This means that the PI expression for the amplitude for the particle to reach a given point on the screen is affected by the magnetic field in a particularly clean way. Before the magnetic field is turned on, we may write $`A=A_1+A_2`$, where $$A_1=_{\mathrm{slit}1}𝒟𝐪e^{iS[𝐪]/\mathrm{}},$$ and similarly for $`A_2`$. Including the magnetic field, $$A_1^{}=_{\mathrm{slit}1}𝒟𝐪e^{i(S[𝐪](e/c){\scriptscriptstyle 𝑑𝐪𝐀})/\mathrm{}}=e^{ie_1𝑑𝐪𝐀/\mathrm{}c}A_1,$$ where we have pulled the line integral out of the PI since it is the same for all paths passing through slit 1 arriving at the point on the screen under consideration. So the amplitude is $`A`$ $`=`$ $`e^{ie_1𝑑𝐪𝐀/\mathrm{}c}A_1+e^{ie_2𝑑𝐪𝐀/\mathrm{}c}A_2`$ $`=`$ $`e^{ie_1𝑑𝐪𝐀/\mathrm{}c}\left(A_1+e^{ie{\scriptscriptstyle 𝑑𝐪𝐀/\mathrm{}c}}A_2\right)`$ $`=`$ $`e^{ie_1𝑑𝐪𝐀/\mathrm{}c}\left(A_1+e^{ie\mathrm{\Phi }/\mathrm{}c}A_2\right).`$ The overall phase is irrelevant, and the interference pattern is influenced directly by the phase $`e\mathrm{\Phi }/\mathrm{}c`$. If we vary this phase continuously (by varying the magnetic flux), we can detect a shift in the interference pattern. For example, if $`e\mathrm{\Phi }/\mathrm{}c=\pi `$, then a spot on the screen which formerly corresponded to constructive interference will now be destructive, and vice-versa. Since the interference is dependent only on the phase difference mod $`2\pi `$, as we vary the flux we get a shift of the interference pattern which is periodic, repeating itself when $`e\mathrm{\Phi }/\mathrm{}c`$ changes by an integer times $`2\pi `$. ### 4.2 Particle Statistics The path integral can be used to see that particles in three dimensions must obey either Fermi or Bose statistics, whereas particles in two dimensions can have intermediate (or fractional) statistics. Consider a system of two identical particles; suppose that there is a short range, infinitely strong repulsive force between the two. We might ask the following question: if at $`t=0`$ the particles are at $`𝐪_1`$ and $`𝐪_2`$, what is the amplitude that the particles will be at $`𝐪_1^{}`$ and $`𝐪_2^{}`$ at some later time $`T`$? We will first examine this question in three dimensions, and then in two dimensions. #### 4.2.1 Three dimensions According to the PI description of the problem, this amplitude is $$A=\underset{\mathrm{paths}}{}e^{iS[𝐪_1(t),𝐪_2(t)]},$$ where we sum over all two-particle paths going from $`𝐪_1,𝐪_2`$ to $`𝐪_1^{},𝐪_2^{}`$. However there is an important subtlety at play: if the particles are identical, then there are (in three dimensions!) two classes of paths (Figure 6). Even though the second path involves an exchange of particles, the final configuration is the same due to the indistinguishability of the particles. It is more economical to describe this situation in terms of the centre-of-mass position $`𝐐=(𝐪_1+𝐪_2)/2`$ and the relative position $`𝐪=𝐪_2𝐪_1`$. The movement of the centre of mass is irrelevant, and we can concentrate on the relative coordinate $`𝐪`$. We can also assume for simplicity that the final positions are the same as the initial ones. Then the two paths above correspond to the paths in relative position space depicted in Figure 7. The point is, of course, that the relative positions $`𝐪`$ and $`𝐪`$ represent the same configuration: interchanging $`𝐪_1`$ and $`𝐪_2`$ changes $`𝐪𝐪`$. We can elevate somewhat the tone of the discussion by introducing some amount of formalism. The configuration space for the relative position of two identical particles is not $`𝐑^3\{0\}`$,<sup>5</sup><sup>5</sup>5Recall that we have supposed that the particles have an infinite, short-range repulsion; hence the subtraction of the origin (which represents coincident points). as one would have naively thought, but $`(𝐑^3\{0\})/𝐙_2`$. The division by the factor $`𝐙_2`$ indicates that opposite points in the space $`𝐑^3\{0\}`$, namely any point $`𝐪`$ and the point diametrically opposite to it $`𝐪`$, are to be identified: they represent the same configuration. We must keep this in mind when we attempt to draw paths: the second path of Figure 7 is a closed one. A topological space such as our configuration space can be characterized as simply connected or as non-simply connected according to whether all paths starting and finishing at the same point can or cannot be contracted into the trivial path (representing no relative motion of the particles). It is clear from Figure 7 that the first path can be deformed to the trivial path, while the second one cannot, so the configuration space is not simply connected. Clearly any path which does not correspond to an exchange of the particles (a “direct” path) is topologically trivial, while any “exchange” path is not; we can divide the space of paths on the configuration space into two topological classes (direct and exchange). Our configuration space is more precisely described as doubly-connected, since any two topologically nontrivial (exchange) paths taken one after the other result in a direct path, which is trivial. Thus the classes of paths form the elements of the group $`𝐙_2`$ if we define the product of two paths to mean first one path followed by the other (a definition which extends readily to the product of classes). One final bit of mathematical nomenclature: our configuration space, as noted above, is $`(𝐑^3\{0\})/𝐙_2`$, which is not simply connected. We define the (simply connected) covering space as the simply connected space which looks locally like the original space. In our case, the covering space is just $`𝐑^3\{0\}`$. At this point you might well be wondering: what does this have to do with PIs? We can rewrite the PI expression for the amplitude as the following PI in the covering space $`𝐑^3\{0\}`$: $`A(𝐪,T;𝐪,0)`$ $`=`$ $`{\displaystyle \underset{\mathrm{direct}}{}}e^{iS[𝐪]}+{\displaystyle \underset{\mathrm{exchange}}{}}e^{iS[𝐪]}`$ (20) $`=`$ $`\overline{A}(𝐪,T;𝐪,0)+\overline{A}(𝐪,T;𝐪,0).`$ The notation $`\overline{A}`$ is used to indicate that these PIs are in the covering space, while $`A`$ is a PI in the configuration space. The first term is the sum over all paths from $`𝐪`$ to $`𝐪`$; the second is that for paths from $`𝐪`$ to $`𝐪`$. Notice that each sub-path integral is a perfectly respectable PI in its own right: each would be a complete PI for the same dynamical problem but involving distinguishable particles. Since the PI can be thought of as a technique for obtaining the propagator in quantum mechanics, and since (as was mentioned at the end of Section 2) the propagator is a solution of the Schroedinger equation, either of these sub-path integrals also satisfies it. It follows that we can generalize the amplitude $`A`$ to the following expression, which still satisfies the Schroedinger equation: $`A(𝐪,T;𝐪,0)A^\varphi (𝐪,T;𝐪,0)`$ $`=`$ $`{\displaystyle \underset{\mathrm{direct}}{}}e^{iS[𝐪]}+e^{i\varphi }{\displaystyle \underset{\mathrm{exchange}}{}}e^{iS[𝐪]}`$ (21) $`=`$ $`\overline{A}(𝐪,T;𝐪,0)+e^{i\varphi }\overline{A}(𝐪,T;𝐪,0).`$ This generalization might appear to be ad hoc and ill-motivated, but we will see shortly that it is intimately related to particle statistics. There is a restriction on the added phase, $`\varphi `$. To see this, suppose that we no longer insist that the path be a closed one from $`𝐪`$ to $`𝐪`$. Then (21) generalizes to $$A^\varphi (𝐪^{},T;𝐪,0)=\overline{A}(𝐪^{},T;𝐪,0)+e^{i\varphi }\overline{A}(𝐪^{},T;𝐪,0).$$ (22) If we vary $`𝐪^{}`$ continuously to the point $`𝐪^{}`$, we have $$A^\varphi (𝐪^{},T;𝐪,0)=\overline{A}(𝐪^{},T;𝐪,0)+e^{i\varphi }\overline{A}(𝐪^{},T;𝐪,0).$$ (23) But since the particles are identical, the new final configuration $`𝐪^{}`$ is identical to old one $`𝐪^{}`$. (22) and (23) are expressions for the amplitude for the same physical process, and can differ at most by a phase: $$A^\varphi (𝐪^{},T;𝐪,0)=e^{i\alpha }A^\varphi (𝐪^{},T;𝐪,0).$$ Combining these three equations, we see that $$\overline{A}(𝐪^{},T;𝐪,0)+e^{i\varphi }\overline{A}(𝐪^{},T;𝐪,0)=e^{i\alpha }\left(\overline{A}(𝐪^{},T;𝐪,0)+e^{i\varphi }\overline{A}(𝐪^{},T;𝐪,0)\right).$$ Equating coefficients of the two terms, we have $`\alpha =\varphi `$ (up to a $`2\pi `$ ambiguity), and $$e^{i2\varphi }=1.$$ This equation has two physically distinct solutions: $`\varphi =0`$ and $`\varphi =\pi `$. (Adding $`2n\pi `$ results in physically equivalent solutions.) If $`\varphi =0`$, we obtain $$A(𝐪,T;𝐪,0)=\overline{A}(𝐪,T;𝐪,0)+\overline{A}(𝐪,T;𝐪,0),$$ (24) the naive sum of the direct and exchange amplitudes, as is appropriate for Bose statistics. If, on the other hand, $`\varphi =\pi `$, we obtain $$A(𝐪,T;𝐪,0)=\overline{A}(𝐪,T;𝐪,0)\overline{A}(𝐪,T;𝐪,0).$$ (25) The direct and exchange amplitudes contribute with a relative minus sign. This case describes Fermi statistics. In three dimensions, we see that the PI gives us an elegant way of seeing how these two types of quantum statistics arise. #### 4.2.2 Two dimensions We will now repeat the above analysis in two dimensions, and will see that the difference is significant. Consider a system of two identical particles in two dimensions, again adding a short-range, infinitely strong repulsion. Once again, we restrict ourselves to the centre of mass frame, since centre-of-mass motion is irrelevant to the present discussion. The amplitude that two particles starting at relative position $`𝐪=(q_x,q_y)`$ will propagate to a final relative position $`𝐪^{}=(q_x^{},q_y^{})`$ in time $`T`$ is $$A(𝐪^{},T;𝐪,0)=\underset{\mathrm{paths}}{}e^{iS[𝐪_1(t),𝐪_2(t)]},$$ the sum being over all paths from $`𝐪`$ to $`𝐪^{}`$ in the configuration space. Once again, the PI separates into distinct topological classes, but there are now an infinity of possible classes. To see this, consider the three paths depicted in Figure 8, where for simplicity we restrict to the case where the initial and final configurations are the same. It is important to remember that drawing paths in the plane is somewhat misleading: as in three dimensions, opposite points are identified, so that a path from any point to the diametrically opposite point is closed. The first and second paths are similar to the direct and exchange paths of the three-dimensional problem. The third path, however, represents a distinct class of path in two dimensions. The particles circle around each other, returning to their starting points. It is perhaps easier to visualize these paths in a three-dimensional space-time plot, where the vertical axis represents time and the horizontal axes represent space (Figure 9). It is clear that in the third path the particle initially at $`𝐪_1`$ returns to $`𝐪_1`$, and similarly for the other particle: this path does not involve a permutation of the particles. It is also clear that this path cannot be continuously deformed into the first path, so it is in a distinct topological class. (It is critical here that we have excised the origin in relative coordinates – i.e., that we have disallowed configurations where the two particles are at the same point in space.) The existence of this third class of paths generalizes in an obvious way, and we are led to the following conclusion: the paths starting and finishing at relative position $`𝐪`$ can be divided into an infinite number of classes of paths in the plane (minus the origin); a class is specified by the number of interchanges of the particles (keeping track of the sense of each interchange). This is profoundly different from the three-dimensional case, where there were only two classes of paths: direct and exchange. If we characterize a path by the polar angle of the relative coordinate, this angle is $`n\pi `$ in the $`n^{th}`$ class, where $`n`$ is an integer. (For the three paths shown above, $`n=`$ 0, 1, and 2, respectively.) We can write $$A(𝐪,T;𝐪,0)=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}\overline{A}_n(𝐪,T;𝐪,0),$$ where $`\overline{A}_n`$ is the covering-space PI considering only paths of change of polar angle $`n\pi `$. This path integral can again be generalized to $$A(𝐪,T;𝐪,0)=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}C_n\overline{A}_n(𝐪,T;𝐪,0),$$ $`C_n`$ being phases. Since each $`\overline{A}_n`$ satisfies the Schroedinger equation, so does this generalization. Again, a restriction on the phases arises, as can be seen by the following argument. Let us relax the condition that the initial and final points are the same; let us denote the final point $`𝐪^{}`$ by its polar coordinates $`(q^{},\theta ^{})`$. Writing $`A(q^{},\theta ^{})A(𝐪^{},T;𝐪,0)`$, we have $$A(q^{},\theta ^{})=\underset{\mathrm{}}{\overset{\mathrm{}}{}}C_n\overline{A}_n(q^{},\theta ^{}).$$ Now, we can change continuously $`\theta ^{}\theta ^{}+\pi `$, yielding $$A(q^{},\theta ^{}+\pi )=\underset{n=\mathrm{}}{\overset{\mathrm{}}{}}C_n\overline{A}_n(q^{},\theta ^{}+\pi ).$$ (26) Two critical observations can now be made. First, the final configuration is unchanged, so $`A(q^{},\theta ^{}+\pi )`$ can differ from $`A(q^{},\theta ^{})`$ by at most a phase: $$A(q^{},\theta ^{}+\pi )=e^{i\varphi }A(q^{},\theta ^{}).$$ Second, $`\overline{A}_n(q^{},\theta ^{}+\pi )=\overline{A}_{n+1}(q^{},\theta ^{})`$, since this is just two different ways of expressing exactly the same quantity. Applying these two observations to (26), $`e^{i\varphi }{\displaystyle \underset{n=\mathrm{}}{\overset{\mathrm{}}{}}}C_n\overline{A}_n(q^{},\theta ^{})`$ $`=`$ $`{\displaystyle \underset{\mathrm{}}{\overset{\mathrm{}}{}}}C_n\overline{A}_{n+1}(q^{},\theta ^{})`$ (27) $`=`$ $`{\displaystyle \underset{\mathrm{}}{\overset{\mathrm{}}{}}}C_{n1}\overline{A}_n(q^{},\theta ^{}).`$ Equating coefficients of $`\overline{A}_n(q^{},\theta ^{})`$, we get $$C_n=e^{i\varphi }C_{n1}.$$ Choosing $`C_0=1`$, we obtain for the amplitude $$A=\underset{\mathrm{}}{\overset{\mathrm{}}{}}e^{in\varphi }\overline{A}_n,$$ (28) which is the two-dimensional analog of the three-dimensional result (21). The most important observation to be made is that there is no longer a restriction on the angle $`\varphi `$, as was the case in three dimensions. We see that, relative to the “naive” PI (that with $`\varphi =0`$), the class corresponding to a net number $`n`$ of counter-clockwise rotations of one particle around the other contributes with an extra phase $`\mathrm{exp}in\varphi `$. If $`\varphi =0`$ or $`\pi `$, this collapses to the usual cases of Bose and Fermi statistics, respectively. However in the general case the phase relation between different paths is more complicated (not determined by whether the path is “direct” or “exchange”); this new possibility is known as fractional statistics, and particles obeying these statistics are known as anyons. Anyons figure prominently in the accepted theory of the fractional quantum Hall effect, and were proposed as being relevant to high-temperature superconductivity, although that possibility seems not to be borne out by experimental results. Perhaps Nature has other applications of fractional statistics which await discovery. ### 4.3 Magnetic Monopoles and Charge Quantization All experimental evidence so far tells us that all particles have electric charges which are integer multiples of a fundamental unit of electric charge, $`e`$.<sup>6</sup><sup>6</sup>6The unit of electric charge is more properly $`e/3`$, that of the quarks; for simplicity, I will ignore this fact. There is absolutely nothing wrong with a theory of electrodynamics of particles of arbitrary charges: we could have particles of charge $`e`$ and $`\sqrt{17}e`$, for example. It was a great mystery why charge was quantized in the early days of quantum mechanics. In 1931, Dirac showed that the quantum mechanics of charged particles in the presence of magnetic monopoles is problematic, unless the product of the electric and magnetic charges is an integer multiple of a given fundamental value. Thus, the existence of monopoles implies quantization of electric charge, a fact which has fueled experimental searches for and theoretical speculations about magnetic monopoles ever since. We will now recast Dirac’s argument into a modern form in terms of PIs. A monopole of charge $`g`$ positioned at the origin has magnetic field $$𝐁=g\frac{\widehat{𝐞}_r}{r^2}.$$ As is well known, this field cannot be described by a normal (smooth, single-valued) vector potential: writing $`𝐁=\times 𝐀`$ implies that the magnetic flux emerging from any closed surface (and thus the magnetic charge contained in any such surface) must be zero. This fact makes life difficult for monopole physics, for several reasons. Although classically the Maxwell equations and the Lorentz force equation form a complete set of equations for particles (both electrically and magnetically charged, with the simple addition of magnetic source terms) and electromagnetic fields, their derivation from an action principle requires that the electromagnetic field be described in terms of the electromagnetic potential, $`A_\mu `$. Quantum mechanically, things are even more severe: one cannot avoid $`A_\mu `$, because the coupling of a particle to the electromagnetic field is written in terms of $`A_\mu `$, not the electric and magnetic fields. Dirac suggested that if a monopole exists, it could be described by an infinitely-thin and tightly-wound solenoid carrying a magnetic flux equal to that of the monopole. The solenoid is semi-infinite in length, running from the position of the monopole to infinity along an arbitrary path. The magnetic field produced by such a solenoid can be shown to be that of a monopole plus the usual field produced by a solenoid, in this case, an infinitely intense, infinitely narrow tube of flux running from the monopole to infinity along the position of the solenoid (Figure 10). Thus except inside the solenoid, the field produced is that of the monopole. The field inside the solenoid is known as the “Dirac string”. The flux brought into any closed surface including the monopole is now zero, because the solenoid brings in a flux equal to that flowing out due to the monopole. Thus, the combined monopole-solenoid can be described by a vector potential. However, in order for this to be a valid description of the monopole, we must somehow convince ourselves that the solenoid can be made invisible to any electrically charged particle passing by it. We can describe the motion of such a particle by a PI, and two paths passing on either side of the Dirac string can contribute to the PI (Figure 11). But the vector potential of the Dirac string will affect the action of each of these paths differently, as we have seen in the Aharonov-Bohm effect. In order that the interference of these paths be unaffected by the presence of the Dirac string, the relative phase must be an integral multiple of $`2\pi `$. This phase is $$\frac{e}{\mathrm{}c}𝑑𝐪𝐀=\frac{e\mathrm{\Phi }}{\mathrm{}c}=\frac{4\pi eg}{\mathrm{}c}.$$ Setting this to $`2\pi n`$, in order for the motion of a particle of charge $`e`$ to be unaffected by the presence of the Dirac string, the electric charge must be $$e=\frac{2\pi \mathrm{}c}{4\pi g}n=\frac{\mathrm{}c}{2g}n.$$ (29) Thus, the existence of magnetic monopoles requires the quantization of electric charge; the fundamental unit of electric charge is $`2\pi \mathrm{}c/g`$. In modern theories of fundamental physics, Grand Unified Theories also imply quantization of electric charge, apparently avoiding the necessity for magnetic monopoles. But any Grand Unified Theory actually has magnetic monopoles as well (though they are of a nature quite different to the “Dirac monopole”), so the intimate relation between magnetic monopoles and the quantization of electric charge is preserved, albeit in a form quite different from that suggested by Dirac. ## 5 Statistical Mechanics via Path Integrals The path integral turns out to provide an elegant way of doing statistical mechanics. The reason for this is that, as we will see, the central object in statistical mechanics, the partition function, can be written as a PI. Many books have been written on statistical mechanics with emphasis on path integrals, and the objective in this lecture is simply to see the relation between the partition function and the PI. The definition of the partition function is $$Z=\underset{j}{}e^{\beta E_j},$$ (30) where $`\beta =1/k_BT`$ and $`E_j`$ is the energy of the state $`|j`$. We can write $$Z=\underset{j}{}j\left|e^{\beta H}\right|j=\mathrm{Tr}e^{\beta H}.$$ But recall the definition of the propagator: $$K(q^{},T;q,0)=q^{}\left|e^{iHT}\right|q.$$ Suppose we consider $`T`$ to be a complex parameter, and consider it to be pure imaginary, so that we can write $`T=i\beta `$, where $`\beta `$ is real. Then $`K(q^{},i\beta ;q,0)`$ $`=`$ $`q^{}\left|e^{iH(i\beta )}\right|q`$ $`=`$ $`q^{}\left|e^{\beta H}\underset{=1}{\underset{}{{\displaystyle \underset{j}{}}|jj|}}\right|q`$ $`=`$ $`{\displaystyle \underset{j}{}}e^{\beta E_j}q^{}|jj|q`$ $`=`$ $`{\displaystyle \underset{j}{}}e^{\beta E_j}j|qq^{}|j.`$ Putting $`q^{}=q`$ and integrating over $`q`$, we get $$𝑑qK(q,i\beta ;q,0)=\underset{j}{}e^{\beta E_j}j\left|\underset{=1}{\underset{}{𝑑q|qq|}}\right|j=Z.$$ (31) This is the central observation of this section: that the propagator evaluated at negative imaginary time is related to the partition function. We can easily work out an elementary example such as the harmonic oscillator. Recall the path integral for it, (17): $$K(q^{},T;q,0)=\left(\frac{m\omega }{2\pi i\mathrm{sin}\omega T}\right)^{1/2}\mathrm{exp}\left\{i\frac{m\omega }{2\mathrm{sin}\omega T}\left((q_{}^{}{}_{}{}^{2}+q^2)\mathrm{cos}\omega T2q^{}q\right)\right\}.$$ We can put $`q^{}=q`$ and $`T=i\beta `$: $$K(q,i\beta ;q,0)=\left(\frac{m\omega }{2\pi \mathrm{sinh}(\beta \omega )}\right)^{1/2}\mathrm{exp}\left\{\frac{m\omega q^2}{\mathrm{sinh}(\beta \omega )}\left(\mathrm{cosh}(\beta \omega )1\right)\right\}.$$ The partition function is thus $`Z`$ $`=`$ $`{\displaystyle 𝑑qK(q,i\beta ;q,0)}=\left({\displaystyle \frac{m\omega }{2\pi \mathrm{sinh}(\beta \omega )}}\right)^{1/2}\sqrt{{\displaystyle \frac{\pi }{\frac{m\omega }{\mathrm{sinh}(\beta \omega )}\left(\mathrm{cosh}(\beta \omega )1\right)}}}`$ $`=`$ $`\left[2(\mathrm{cosh}(\beta \omega )1)\right]^{1/2}=\left[e^{\beta \omega /2}(1e^{\beta \omega })\right]^1`$ $`=`$ $`{\displaystyle \frac{e^{\beta \omega /2}}{1e^{\beta \omega }}}={\displaystyle \underset{j=0}{\overset{\mathrm{}}{}}}e^{\beta (j+1/2)\omega }.`$ Putting $`\mathrm{}`$ back in, we get the familiar result $$Z=\underset{j=0}{\overset{\mathrm{}}{}}e^{\beta (j+1/2)\mathrm{}\omega }.$$ The previous calculation actually had nothing to do with PIs. The result for $`K`$ was derived via PIs earlier, but it can be derived (more easily, in fact) in ordinary quantum mechaincs. However we can rewrite the partition function in terms of a PI. In ordinary (real) time, $$K(q^{},T;q,0)=𝒟q(t)\mathrm{exp}i_0^T𝑑t\left(\frac{m\dot{q}^2}{2}V(q)\right),$$ where the integral is over all paths from $`(q,0)`$ to $`(q^{},T)`$. With $`q^{}=q`$, $`Ti\beta `$, $$K(q,i\beta ;q,0)=𝒟q(t)\mathrm{exp}i_0^{i\beta }𝑑t\left(\frac{m\dot{q}^2}{2}V(q)\right).$$ where we now integrate along the negative imaginary time axis (Figure 12). Let us define a real variable for this integration, $`\tau =it`$. $`\tau `$ is called the imaginary time, since when the time $`t`$ is imaginary, $`\tau `$ is real. (Kind of confusing, admittedly, but true.) Then the integral over $`\tau `$ is along its real axis: when $`t:0i\beta `$, then $`\tau :0\beta `$. We can write $`q`$ as a function of the variable $`\tau `$: $`q(t)q(\tau )`$; then $`\dot{q}=idq/d\tau `$. The propagator becomes $$K(q,i\beta ;q,0)=𝒟q(\tau )\mathrm{exp}_0^\beta 𝑑\tau \left(\frac{m}{2}\left(\frac{dq}{d\tau }\right)^2+V(q)\right).$$ (32) The integral is over all functions $`q(\tau )`$ such that $`q(0)=q(\beta )=q`$. The result (32) is an “imaginary-time” or “Euclidean” path integral, defined by associating to each path an amplitude (statistical weight) $`\mathrm{exp}S_E`$, where $`S_E`$ is the so-called Euclidean action, obtained from the usual (“Minkowski”) action by changing the sign of the potential energy term. The Euclidean PI might seem like a strange, unphysical beast, but it actually has many uses. One will be discussed in the next section, where use will be made of the fact that at low temperatures the ground state gives the dominant contribution to the partition function. It can therefore be used to find the ground state energy. We will also see the Euclidean PI in Section 9, when discussing the subject of instantons, which are used to describe phenomena such as quantum mechanical tunneling. ## 6 Perturbation Theory in Quantum Mechanics We can use the Euclidean PI to compute a perturbation expansion for the ground state energy (among other things). This is not terribly useful in and of itself (once again, conventional techniques are a good deal easier), but the techniques used are very similar to those used in perturbation theory and Feynman diagrams in field theory. For this reason, we will discuss corrections to the ground state energy of an elementary quantum mechanical system in some detail. From $`Z`$ it is quite easy to extract the ground state energy. (This is a well-known fact of statistical mechanics, quite independent of PIs.) From the definition of $`Z`$, $$Z(\beta )=\underset{j}{}e^{\beta E_j},$$ we can see that the contribution of each state decreases exponentially with $`\beta `$. However, that of the ground state decreases less slowly than any other state. So in the limit of large $`\beta `$ (i.e., low temperature), the ground state contribution will dominate. (This is mathematically straightforward, and also physically reasonable.) One finds $$E_0=\underset{\beta \mathrm{}}{lim}\frac{1}{\beta }\mathrm{log}Z.$$ (33) In fact, we can extract $`E_0`$ from something slightly easier to calculate than $`Z`$. Rather than integrating over the initial (= final) position, as with $`Z`$, let us look at the Euclidean propagator from $`q=0`$ to $`q^{}=0`$ (the choice of zero is arbitrary). $$K_E(0,\beta ;0,0)=q^{}=0\left|e^{\beta H}\right|q=0.$$ We can insert a complete set of eigenstates of $`H`$: $`K_E(0,\beta ;0,0)`$ $`=`$ $`q^{}=0\left|e^{\beta H}{\displaystyle \underset{j}{}}\right|jj|q=0`$ $`=`$ $`{\displaystyle \underset{j}{}}e^{\beta E_j}\varphi _j(0)\varphi _j^{}(0),`$ where $`\varphi _j`$ are the wave functions of $`H`$. Again the ground state dominates as $`\beta \mathrm{}`$, and $$E_0=\underset{\beta \mathrm{}}{lim}\frac{1}{\beta }\mathrm{log}K_E(0,\beta ;0,0).$$ (34) (As $`\beta \mathrm{}`$, the difference between $`\beta ^1\mathrm{log}Z`$ and $`\beta ^1\mathrm{log}K_E`$ goes to zero.) So let us see how we can calculate $`K_E(0,\beta ;0,0)`$ perturbatively via the PI. The starting point is $$K_E(0,\beta ;0,0)=𝒟qe^{S_E(\dot{q},q)},$$ where the paths over which we integrate start and finish at $`q=0`$, and where the Euclidean action is $$S_E=_0^\beta 𝑑\tau \left(\frac{m\dot{q}^2}{2}+V(q)\right),$$ and with $`\dot{q}=dq/d\tau `$. As an example, consider the anharmonic oscillator, with quadratic and quartic terms in the potential: $$K_E(0,\beta ;0,0)=𝒟q\mathrm{exp}𝑑\tau \left(\frac{1}{2}m\dot{q}^2+\frac{1}{2}m\omega ^2q^2+\frac{\lambda }{4!}q^4\right).$$ (35) Clearly it is the quartic term which complicates life considerably; we cannot do the PI exactly.<sup>7</sup><sup>7</sup>7In fact, the situation is exactly like the evaluation of the ordinary integral $$I=_{\mathrm{}}^{\mathrm{}}𝑑x\mathrm{exp}(\frac{1}{2}x^2+\frac{\lambda }{4!}x^4),$$ which looks innocent enough but which cannot be evaluated exactly. The technique which we will develop to evaluate (35) can also be used for this ordinary integral – an amusing and recommended exercise. But we can use the following trick to evaluate it perturbatively in $`\lambda `$. (This trick is far more complicated than necessary for this problem, but is a standard – and necessary! – trick in quantum field theory.) Define $`K_E^0[J]`$, the PI for a harmonic oscillator with a source term (which describes the action of an external force) added to the Lagrangian: $$K_E^0[J]=𝒟q\mathrm{exp}𝑑\tau \left(\frac{1}{2}m\dot{q}^2+\frac{1}{2}m\omega ^2q^2J(\tau )q(\tau )\right).$$ (36) Unlike (35), this PI can be evaluated exactly; we will do this (as much as is necessary, at least) shortly. Once we have evaluated it, how does it help us to compute (35)? To see the use of $`K_E^0[J]`$, acting on it with a derivative has the effect of putting a factor $`q`$ in the PI: for any time $`\tau _1`$, $$\frac{\delta K_E^0[J]}{\delta J(\tau _1)}=𝒟qq(\tau _1)\mathrm{exp}𝑑\tau \left(\frac{1}{2}m\dot{q}^2+\frac{1}{2}m\omega ^2q^2J(\tau )q(\tau )\right).$$ A second derivative puts a second $`q`$ in the PI: $$\frac{\delta ^2K_E^0[J]}{\delta J(\tau _1)\delta J(\tau _2)}=𝒟qq(\tau _1)q(\tau _2)\mathrm{exp}𝑑\tau \left(\frac{1}{2}m\dot{q}^2+\frac{1}{2}m\omega ^2q^2J(\tau )q(\tau )\right).$$ In fact, we can generalize this to an arbitrary functional $`F`$: $$F\left[\frac{\delta }{\delta J}\right]K_E^0[J]=𝒟qF[q]e^{S_E^0[J]},$$ (37) where $`S_E^0[J]`$ is the Euclidean action for the harmonic oscillator with source. (To prove (37), bring $`F[\delta /\delta J]`$ inside the PI; each $`\delta /\delta J`$ in $`F`$ operating on $`\mathrm{exp}S_E^0[J]`$ gives rise to a $`q`$ in front of the exponential.) Now, if we choose $`F[q]=\mathrm{exp}𝑑\tau \frac{\lambda }{4!}q^4`$, we get: $`e^{{\scriptscriptstyle 𝑑\tau {\scriptscriptstyle \frac{\lambda }{4!}}\left({\scriptscriptstyle \frac{\delta }{\delta J}}\right)^4}}K_E^0[J]`$ $`=`$ $`{\displaystyle 𝒟q\mathrm{exp}\left\{𝑑\tau \frac{\lambda }{4!}q^4\right\}e^{S_E^0[J]}}`$ $`=`$ $`{\displaystyle 𝒟q\mathrm{exp}}{\displaystyle 𝑑\tau \left(\frac{1}{2}m\dot{q}^2+\frac{1}{2}m\omega ^2q^2+\frac{\lambda }{4!}q^4J(\tau )q(\tau )\right)}.`$ If we now put $`J=0`$, we have the PI we started with. So the final result is: $$\overline{)K_E(0,\beta ;0,0)=\left(\mathrm{exp}\{d\tau \frac{\lambda }{4!}\left(\frac{\delta }{\delta J}\right)^4\}K_E^0[J]\right)|_{J=0}.}$$ (38) We can, and will, calculate $`K_E^0[J]`$ as an explicit functional of $`J`$. If we then expand the exponential which operates on it in (38), we get a power series in $`\lambda `$: $`K_E(0,\beta ;0,0)`$ $`=`$ $`\{(1{\displaystyle }d\tau {\displaystyle \frac{\lambda }{4!}}\left({\displaystyle \frac{\delta }{\delta J(\tau )}}\right)^4`$ $`+{\displaystyle \frac{1}{2!}}{\displaystyle }d\tau {\displaystyle \frac{\lambda }{4!}}\left({\displaystyle \frac{\delta }{\delta J(\tau )}}\right)^4{\displaystyle }d\tau ^{}{\displaystyle \frac{\lambda }{4!}}\left({\displaystyle \frac{\delta }{\delta J(\tau ^{})}}\right)^4+\mathrm{})K_E^0[J]\left\}\right|_{J=0}`$ $`=`$ $`K_E^0[J]{\displaystyle \frac{\lambda }{4!}}\left({\displaystyle 𝑑\tau \left(\frac{\delta }{\delta J(\tau )}\right)^4K_E^0[J]}\right)|_{J=0}+o(\lambda ^2).`$ Let us now evaluate $`K_E^0[J]`$, $$K_E^0[J]=𝒟q\mathrm{exp}𝑑\tau \left(\frac{1}{2}m\dot{q}^2+\frac{1}{2}m\omega ^2q^2J(\tau )q(\tau )\right).$$ To do this, suppose that we can find the classical path $`q_{cJ}(\tau )`$, the solution of $$m\ddot{q}=m\omega ^2qJ(\tau ),q(0)=q(\beta )=0.$$ (39) Once we have done this, we can perform a change of variables in the PI: we define $`q(\tau )=q_{cJ}(\tau )+y(\tau )`$, and integrate over paths $`y(\tau )`$. This is useful because $`{\displaystyle 𝑑\tau \left(\frac{1}{2}m\dot{q}^2+\frac{1}{2}m\omega ^2q^2J(\tau )q(\tau )\right)}`$ $`=`$ $`{\displaystyle 𝑑\tau \left(\frac{1}{2}m\dot{q}_{cJ}^2+\frac{1}{2}m\omega ^2q_{cJ}^2J(\tau )q_{cJ}(\tau )\right)}`$ $`+\underset{=0}{\underset{}{(\text{linear in }y)}}+{\displaystyle 𝑑\tau \left(\frac{1}{2}m\dot{y}^2+\frac{1}{2}m\omega ^2y^2\right)}.`$ The linear term vanishes because $`q_{cJ}`$ satisfies the equation of motion. So the PI becomes $$K_E^0[J]=e^{S_{Ec}[J]}𝒟y\mathrm{exp}𝑑\tau \left(\frac{1}{2}m\dot{y}^2+\frac{1}{2}m\omega ^2y^2\right).$$ The crucial observation is that the resulting PI is independent of $`J`$: it is an irrelevant constant; call it $`C`$. (In fact, $`C`$ is neither constant \[it depends on $`\beta `$\], nor entirely irrelevant \[it is related to the unperturbed ground state energy, as we will see\]. Crucial for the present purposes is that $`C`$ is independent of $`J`$.) $$K_E^0[J]=Ce^{S_{Ec}[J]},$$ where $$S_{Ec}[J]=𝑑\tau \left(\frac{1}{2}m\dot{q}_{cJ}^2+\frac{1}{2}m\omega ^2q_{cJ}^2J(\tau )q_{cJ}(\tau )\right).$$ This can be simplified by integrating the first term by parts, yielding $$S_{Ec}[J]=𝑑\tau q_{cJ}\left(\frac{1}{2}m\ddot{q}_{cJ}+\frac{1}{2}m\omega ^2q_{cJ}J(\tau )\right).$$ Using the classical equation of motion (39), we get $$S_{Ec}[J]=\frac{1}{2}𝑑\tau J(\tau )q_{cJ}(\tau ).$$ We must still solve the classical problem, (39). The solution can be written in terms of the Green’s function for the problem. Let $`G(\tau ,\tau ^{})`$ be the solution of $`m\left({\displaystyle \frac{d^2}{d\tau ^2}}\omega ^2\right)G(\tau ,\tau ^{})`$ $`=`$ $`\delta (\tau \tau ^{}),`$ $`G(0,\tau ^{})=G(\beta ,\tau ^{})`$ $`=`$ $`0.`$ Then we can immediately write $$q_{cJ}(\tau )=_0^\beta 𝑑\tau ^{}G(\tau ,\tau ^{})J(\tau ^{}),$$ which can be proven by substution into (39). We can now write $$\overline{)K_E^0[J]=C\mathrm{exp}\frac{1}{2}d\tau d\tau ^{}J(\tau )G(\tau ,\tau ^{})J(\tau ^{}).}$$ (40) We can find the Green’s function easily in the limit $`\beta \mathrm{}`$. It is slightly more convenient to treat the initial and final times more symmetrically, so let us choose the time interval to be $`(\beta /2,+\beta /2)`$; in the limit $`\beta \mathrm{}`$ we go from $`\mathrm{}`$ to $`\mathrm{}`$. Then we have $$m\left(\frac{d^2}{d\tau ^2}\omega ^2\right)G(\tau ,\tau ^{})=\delta (\tau \tau ^{}).$$ By taking the Fourier transform, we see that $$G(\tau ,\tau ^{})=\frac{1}{m}_{\mathrm{}}^{\mathrm{}}\frac{dk}{2\pi }\frac{1}{(k^2+\omega ^2)}e^{ik(\tau \tau ^{})}.$$ (41) We can now compute the first-order correction to $`K_E`$ (from which we get the first-order correction to the ground state energy). We have $$K_E=K_E^0[0]\frac{\lambda }{4!}𝑑\tau \left(\frac{\delta }{\delta J(\tau )}\right)^4K_E^0[J]|_{J=0}.$$ (42) Since in the second term we take four derivatives of $`K_E^0[J]`$ and then set $`J=0`$, only the piece of $`K_E^0[J]`$ which is quartic in $`J`$ is relevant: fewer than four $`J`$’s will be killed by the derivatives; more than four will be killed when setting $`J=0`$. $`K_E^0[J]`$ $`=`$ $`C\mathrm{exp}{\displaystyle \frac{1}{2}}{\displaystyle 𝑑\tau 𝑑\tau ^{}J(\tau )G(\tau ,\tau ^{})J(\tau ^{})}`$ (43) $`=`$ $`\mathrm{irrelevant}+C{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{1}{2}}{\displaystyle 𝑑\tau 𝑑\tau ^{}J(\tau )G(\tau ,\tau ^{})J(\tau ^{})}\right)^2`$ $`=`$ $`{\displaystyle \frac{C}{8}}J_1G_{12}J_2J_3G_{34}J_4,`$ where we have used the compact notation $`J_1G_{12}J_2=𝑑\tau _1𝑑\tau _2J(\tau _1)G(\tau _1,\tau _2)J(\tau _2)`$. Substituting (43) into (42), $$K_E=C\left(1\frac{\lambda }{4!}\frac{1}{8}𝑑\tau \left(\frac{\delta }{\delta J(\tau )}\right)^4J_1G_{12}J_2J_3G_{34}J_4+o(\lambda ^2)\right).$$ (44) To ensure that we understand the notation and how functional differentiation works, let us work out a slightly simpler example than the above. Consider $$X\left(\frac{\delta }{\delta J(\tau )}\right)^2J_1G_{12}J_2=\left(\frac{\delta }{\delta J(\tau )}\right)^2𝑑\tau _1𝑑\tau _2J(\tau _1)G(\tau _1,\tau _2)J(\tau _2).$$ The first derivative can act either on $`J_1`$ or $`J_2`$. In either case, it gives a delta function, which will make one of the integrals collapse: $`X`$ $`=`$ $`{\displaystyle \frac{\delta }{\delta J(\tau )}}{\displaystyle 𝑑\tau _1𝑑\tau _2\left(\delta (\tau \tau _1)G(\tau _1,\tau _2)J(\tau _2)+J(\tau _1)G(\tau _1,\tau )\delta (\tau \tau _2)\right)}`$ $`=`$ $`{\displaystyle \frac{\delta }{\delta J(\tau )}}\left({\displaystyle 𝑑\tau _2G(\tau ,\tau _2)J(\tau _2)}+{\displaystyle 𝑑\tau _1J(\tau _1)G(\tau _1,\tau )}\right)`$ In each term the remaining derivative acts similarly and kills the remaining integral; the result is $$X=2G(\tau ,\tau ).$$ The functional derivatives in (44) are a straightforward generalization of this; we find $$K_E=C\left(1\frac{1}{8}\frac{\lambda }{4!}𝑑\tau \mathrm{\hspace{0.17em}4}!G(\tau ,\tau )^2\right).$$ From (41) $`G(\tau ,\tau )=1/2m\omega `$; the $`\tau `$ integral is just the time interval $`\beta `$, and finally we get $$K_E(0,\beta ;0,0)=C\left(1\frac{\beta \lambda }{32m^2\omega ^2}+o(\lambda ^2)\right)=Ce^{\beta \lambda /32m^2\omega ^2}$$ (45) to order $`\lambda `$. Now we can put this expression to good use, extracting the ground state energy from (34): $$E_0=\underset{\beta \mathrm{}}{lim}\frac{1}{\beta }\mathrm{log}K_E(0,\beta ;0,0)=\underset{\beta \mathrm{}}{lim}\frac{1}{\beta }\left(\mathrm{log}C\frac{\beta \lambda }{32m^2\omega ^2}\right).$$ Recall that the constant $`C`$ in (45) depends on $`\beta `$; this dependence must account for the ground state energy; the term linear in $`\lambda `$ gives the first correction to the energy. Thus $$E_0=\frac{1}{2}\mathrm{}\omega +\frac{\mathrm{}^2\lambda }{32m^2\omega ^2}+o(\lambda ^2)$$ where we have reintroduced $`\mathrm{}`$. We can check this result against standard perturbation theory (which is considerably easier!); the first-order correction to the ground state energy is $$_{\mathrm{}}^{\mathrm{}}𝑑q\varphi _0^{}(q)\left(\frac{\lambda }{4!}q^4\right)\varphi _0(q)=\mathrm{}=\frac{\mathrm{}^2\lambda }{32m^2\omega ^2},$$ as above. One’s sanity would be called into question were it suggested that the PI calculation is a serious competitor for standard perturbation theory, although the latter itself gets rapidly more and more messy at higher orders. The technique above also gets messier, but it may well be that its messiness increases less quickly than that of standard perturbation theory. If so, the PI calculation could become competitive with standard perturbation theory at higher orders. But really the main motivation for discussing the above method is that it mimics in a more familiar setting standard perturbation techniques in quantum field theory. To summarize this long and somewhat technical section, let us recall the main features of the above method. We express the ground state energy as an expression involving the propagator, (34). We separate the Lagrangian into a “free” (i.e., quadratic) part and an “interacting” (beyond quadratic) part. Via the PI, we write the interacting propagator in terms of a free propagator with source term added, (38); this expression is amenable to a perturbation expansion. The free propagator can be evaluated explicitly, (40); then (38) can be computed to any desired order. From this we obtain directly the ground state energy to the same order. ## 7 Green’s Functions in Quantum Mechanics In quantum field theory we are interested in objects such as $$0\left|T\widehat{\varphi }(x_1)\widehat{\varphi }(x_2)\mathrm{}\widehat{\varphi }(x_n)\right|0,$$ the vacuum expectation value of a time-ordered product of Heisenberg field operators. This object is known as a Green’s function or as a correlation function. The order of the operators is such that the earliest field is written last (right-most), the second earliest second last, etc. For example, $$T\widehat{\varphi }(x_1)\widehat{\varphi }(x_2)=\{\begin{array}{cc}\hfill \widehat{\varphi }(x_1)\widehat{\varphi }(x_2)& x_1^0>x_2^0\hfill \\ \hfill \widehat{\varphi }(x_2)\widehat{\varphi }(x_1)& x_2^0>x_1^0\hfill \end{array}$$ Green’s functions are related to amplitudes for physical processes such as scattering and decay processes. (This point is explained in most quantum field theory books.) Let us look at the analogous object in quantum mechanics: $$G^{(n)}(t_1,t_2,\mathrm{},t_n)=0\left|T\widehat{q}(t_1)\widehat{q}(t_2)\mathrm{}\widehat{q}(t_n)\right|0.$$ We will develop a PI expression for this. First, we must recast the PI in terms of Heisenberg representation objects. The operator $`\widehat{q}(t)`$ is the usual Heisenberg operator, defined in terms of the Schroedinger operator $`\widehat{q}`$ by $`\widehat{q}(t)=e^{iHt}\widehat{q}e^{iHt}`$. The eigenstates of the Heisenberg operator are $`|q,t`$: $`\widehat{q}(t)|q,t=q|q,t`$. The relation with the time-independent eigenstates is $`|q,t=e^{iHt}|q`$.<sup>8</sup><sup>8</sup>8There is a possible point of confusion here. We all know that Heisenberg states are independent of time, yet the eigenstates of $`\widehat{q}(t)`$ depend on time. Perhaps the best way to view these states $`|q,t`$ is that they form, for any fixed time, a complete set of states. Just like the usual (time-independent) Heisenberg state $`|q`$ describes a particle which is localized at the point $`q`$ at time $`t=0`$, the state $`|q,t`$ describes a particle which is localized at the point $`q`$ at time $`t`$. Then we can write the PI: $$K=q^{}\left|e^{iHT}\right|q=q^{},T|q,0=𝒟qe^{iS}.$$ We can now calculate the “2-point function” $`G(t_1,t_2)`$, via the PI. We will proceed in two steps. First, we will calculate the following expression: $$q^{},T\left|T\widehat{q}(t_1)\widehat{q}(t_2)\right|q,0.$$ We will then devise a method for extracting the vacuum contribution to the initial and final states. Suppose first that $`t_1>t_2`$. Then $`q^{},T\left|T\widehat{q}(t_1)\widehat{q}(t_2)\right|q,0`$ $`=`$ $`q^{},T\left|\widehat{q}(t_1)\widehat{q}(t_2)\right|q,0`$ $`=`$ $`{\displaystyle 𝑑q_1𝑑q_2q^{},T|q_1,t_1\underset{q_1,t_1|q_1}{\underset{}{q_1,t_1|\widehat{q}(t_1)}}\underset{q_2|q_2,t_2}{\underset{}{\widehat{q}(t_2)|q_2,t_2}}q_2,t_2|q,0}`$ $`=`$ $`{\displaystyle 𝑑q_1𝑑q_2q_1q_2q^{},T|q_1,t_1q_1,t_1|q_2,t_2q_2,t_2|q,0}.`$ Each of these matrix elements is a PI: $$q^{},T\left|T\widehat{q}(t_1)\widehat{q}(t_2)\right|q,0=𝑑q_1𝑑q_2q_1q_2_{q_1,t_1}^{q^{},T}𝒟qe^{iS}_{q_2,t_2}^{q_1,t_1}𝒟qe^{iS}_{q,0}^{q_2,t_2}𝒟qe^{iS}.$$ This expression consists of a first PI from the initial position $`q`$ to an arbitrary position $`q_2`$, a second one from there to a second arbitrary position $`q_1`$, and a third one from there to the final position $`q^{}`$. So we are integrating over all paths from $`q`$ to $`q^{}`$, subject to the restriction that the paths pass through the intermediate points $`q_1`$ and $`q_2`$. We then integrate over the two arbitrary positions, so that in fact we are integrating over all paths: we can combine these three path integrals plus the integrations over $`q_1`$ and $`q_2`$ into one PI. The factors $`q_1`$ and $`q_2`$ in the above integral can be incorporated into this PI by simply including a factor $`q(t_1)q(t_2)`$ in the PI. So $$q^{},T\left|\widehat{q}(t_1)\widehat{q}(t_2)\right|q,0=_{q,0}^{q^{},T}𝒟qq(t_1)q(t_2)e^{iS}(t_1>t_2).$$ An identical calculation shows that exactly this same final expression is also valid for $`t_2<t_1`$: magically, the PI does the time ordering automatically. Thus for all times $$q^{},T\left|T\widehat{q}(t_1)\widehat{q}(t_2)\right|q,0=_{q,0}^{q^{},T}𝒟qq(t_1)q(t_2)e^{iS}.$$ As for how to obtain vacuum-to-vacuum matrix elements, our work on statistical mechanics provides us with a clue. We can expand the states $`q^{},T|`$ and $`|q,0`$ in terms of eigenstates of the Hamiltonian. If we evolve towards a negative imaginary time, the contribution of all other states will decay away relative to that of the ground state. We have (resetting the initial time to $`T`$ for convenience) $$q^{},T|q,T0,T|0,T,$$ where on the right the “0” denotes the ground state. The proportionality involves the ground state wave function and an exponential factor $`\mathrm{exp}2iE_0T=\mathrm{exp}2E_0|T|`$. We could perform all calculations in a Euclidean theory and analytically continue to real time when computing physical quantities (many books do this), but to be closer to physics we can also consider $`T`$ not to be pure imaginary and negative, but to have a small negative imaginary phase: $`T=|T|e^{iϵ}`$ ($`ϵ>0`$). In what follows, I will simply write $`T`$, but please keep in mind that it has a negative imaginary part! With this, $$0,T|0,Tq^{},T|q,T=𝒟qe^{iS}.$$ To compute the Green’s functions, we must simply add $`T\widehat{q}(t_1)\widehat{q}(t_2)\mathrm{}\widehat{q}(t_n)`$ to the matrix element, and the corresponding factor $`q(t_1)q(t_2)\mathrm{}q(t_n)`$ inside the PI: $$0,T\left|T\widehat{q}(t_1)\widehat{q}(t_2)\mathrm{}\widehat{q}(t_n)\right|0,T𝒟qq(t_1)q(t_2)\mathrm{}q(t_n)e^{iS}.$$ The proportionality sign is a bit awkward; fortunately, we can rid ourselves of it. To do this, we note that the left hand expression is not exactly what we want: the vacua $`|0,\pm T`$ differ by a phase. We wish to eliminate this phase; to this end, the Green’s functions are defined $`G^{(n)}(t_1,t_2,\mathrm{},t_n)`$ $`=`$ $`0\left|T\widehat{q}(t_1)\widehat{q}(t_2)\mathrm{}\widehat{q}(t_n)\right|0`$ $``$ $`{\displaystyle \frac{0,T\left|T\widehat{q}(t_1)\widehat{q}(t_2)\mathrm{}\widehat{q}(t_n)\right|0,T}{0,T|0,T}}`$ $`=`$ $`{\displaystyle \frac{𝒟qq(t_1)q(t_2)\mathrm{}q(t_n)e^{iS}}{𝒟qe^{iS}}},`$ with no proportionality sign. The wave functions and exponential factors in the numerator and denominator cancel. To compute the numerator, we can once again use the trick we used in perturbation theory in quantum mechanics, namely, adding a source to the action. We define $$Z[J]=\frac{𝒟qe^{i(S+𝑑tJ(t)q(t))}}{𝒟qe^{iS}}=\frac{0|0_J}{0|0_{J=0}}.$$ If we operate on $`Z[J]`$ with $`i^1\delta /\delta J(t_1)`$, this gives $`\left({\displaystyle \frac{1}{i}}{\displaystyle \frac{\delta }{\delta J(t_1)}}Z[J]\right)|_{J=0}`$ $`=`$ $`\left({\displaystyle \frac{𝒟qq(t_1)e^{i(S+𝑑tJ(t)q(t))}}{𝒟qe^{iS}}}\right)|_{J=0}`$ $`=`$ $`{\displaystyle \frac{𝒟qq(t_1)e^{iS}}{𝒟qe^{iS}}}`$ $`=`$ $`{\displaystyle \frac{0,T\left|\widehat{q}(t_1)\right|0,T}{0,T|0,T}}=0\left|\widehat{q}(t_1)\right|0`$ (The expectation values are evaluated in the absence of $`J`$.) Repeating this procedure, we obtain a PI with several $`q`$’s in the numerator. This ordinary product of $`q`$’s in the PI corresponds, as discussed earlier in this section, to a time-ordered product in the matrix element. So we make the following conclusion: $$\left(\frac{1}{i}\frac{\delta }{\delta J(t_1)}\mathrm{}\frac{1}{i}\frac{\delta }{\delta J(t_n)}Z[J]\right)|_{J=0}=\frac{𝒟qq(t_1)\mathrm{}q(t_n)e^{iS}}{𝒟qe^{iS}}=0\left|T\widehat{q}(t_1)\mathrm{}\widehat{q}(t_1)\right|0.$$ For obvious reasons, the functional $`Z[J]`$ is called the generating functional for Green’s functions; it is a very handy tool in quantum field theory and in statistical mechanics. How do we calculate $`Z[J]`$? Let us examine the numerator: $$N𝒟qe^{i(S+{\scriptscriptstyle 𝑑tJ(t)q(t)})}.$$ Suppose initially that $`S`$ is the harmonic oscillator action (denoted $`S_0`$): $$S_0=𝑑t\left(\frac{1}{2}m\dot{q}^2\frac{1}{2}m\omega ^2q^2\right),$$ Then the corresponding numerator, $`N_0`$, is the non-Euclidean (i.e., real-time) version of the propagator $`K_E^0[J]`$ we used in Section 6. We can calculate $`N_0[J]`$ in the same way as $`K_E^0[J]`$. Since the calculation repeats much of that of $`K_E^0[J]`$, we will be succinct. By definition, $$N_0=𝒟q(t)\mathrm{exp}i𝑑t\left(\frac{1}{2}m\dot{q}^2\frac{1}{2}m\omega ^2q^2+Jq\right).$$ We do the path integral over a new variable $`y`$, defined by $`q(t)=q_c(t)+y(t)`$, where $`q_c`$ is the classical solution. Then the PI over $`y`$ is a constant (independent of $`J`$) and we can avoid calculating it. (It will cancel against the denominator in $`Z[J]`$.) Calling it $`C`$, we have $$N_0=Ce^{iS_{0J}[q_c]},$$ where $$S_{0J}[q_c]=𝑑t\left(\frac{1}{2}m\dot{q}_{c}^{}{}_{}{}^{2}\frac{1}{2}m\omega ^2q_c^2+Jq_c\right)=\frac{1}{2}𝑑tJ(t)q_c(t),$$ using the fact that $`q_c`$ satisfies the equation of motion. We can write the classical path in terms of the Green’s function (to be determined shortly), defined by $$\left(\frac{d^2}{dt^2}+\omega ^2\right)G(t,t^{})=i\delta (tt^{}).$$ (46) Then $$q_c(t)=i𝑑t^{}G(t,t^{})J(t^{}).$$ We can now write $$N_0=C\mathrm{exp}\frac{1}{2}𝑑t𝑑t^{}J(t)G(t,t^{})J(t^{}).$$ Dividing by the denominator merely cancels the factor $`C`$, giving our final result: $$\overline{)Z[J]=\mathrm{exp}\frac{1}{2}dtdt^{}J(t)G(t,t^{})J(t^{}).}$$ We can solve (46) for the Green’s function by going into momentum space; the result is $$G(t,t^{})=G(tt^{})=\frac{dk}{2\pi }\frac{i}{k^2\omega ^2}e^{ik(tt^{})}.$$ However, there are poles on the axis of integration. (This problem did not arise in Euclidean space; see (41).) The Green’s function is ambiguous until we give it a “pole prescription”, i.e., a boundary condition. But remember that our time $`T`$ has a small, negative imaginary part. We require that $`G`$ go to zero as $`T\mathrm{}`$. The correct pole prescription then turns out to be $$\overline{)G(tt^{})=\frac{dk}{2\pi }\frac{i}{k^2\omega ^2+iϵ}e^{ik(tt^{})}.}$$ (47) We could at this point do a couple of practice calculations to get used to this formalism. Examples would be to compute perturbatively the generating functional for an action which has terms beyond quadratic (for example, a $`q^4`$ term), or to compute some Green’s function in either the quadratic or quartic theory. But since these objects aren’t really useful in quantum mechanics, without further delay we will go directly to the case of interest: quantum field theory. ## 8 Green’s Functions in Quantum Field Theory It is easy to generalize the PI to many degrees of freedom; we have in fact already done so in Section 4, where particles move in two or three dimensions. It is simply a matter of adding a new index to denote the different degrees of freedom (be they the different coordinates of a single particle in more than one dimension or the particle index for a system of many particles). One of the most important examples of a system with many degrees of freedom is a field theory: $`q(t)\varphi (𝐱,t)=\varphi (x)`$. Not only is this a system of many degrees of freedom, but one of a continuum of degrees of freedom. The passage from a discrete to continuous system in path integrals can be done in the same way as in ordinary classical field theory: we can discretize the field (modeling it by a set of masses and springs, for instance), do the usual path integral manipulations on the discrete system, and take the continuum limit at the end of the calculation. The final result is a fairly obvious generalization of the one-particle results, so I will not dwell on the mundane details of discretization and subsequent taking of the continuum limit. The analog of the quantum mechanical propagator is the transition amplitude to go from one field configuration $`\varphi (𝐱)`$ at $`t=0`$ to another $`\varphi ^{}(𝐱^{})`$ at $`t=T`$: $$K(\varphi ^{}(𝐱^{}),T;\varphi (𝐱),0)=𝒟\varphi e^{iS[\varphi ]},$$ (48) where $`S`$ is the field action, for instance $$S[\varphi ]=d^4x\left(\frac{1}{2}(_\mu \varphi )^2\frac{1}{2}m^2\varphi ^2\right)$$ (49) for the free scalar field. In (48) the integral is over all field configurations $`\varphi (x)`$ obeying the stated initial and final conditions. In field theory, we are not really interested in this object. Rather (as mentioned earlier), we are interested in Green’s functions. Most of the work required to translate (48) into an expression for a Green’s function (generating functional of Green’s functions, more precisely) has already been done in the last section, so let us study a couple of cases. ### 8.1 Free scalar field. For the free scalar field, whose action is given by (49), the generating functional is $$Z_0[J]=\frac{0|0_J}{0|0_{J=0}}.$$ Both numerator and denominator can be written in terms of PIs. The numerator is $$N_0=𝒟\varphi \mathrm{exp}id^4x\left(\frac{1}{2}(_\mu \varphi )^2\frac{1}{2}m^2\varphi ^2+J\varphi \right).$$ We write $`\varphi =\varphi _c+\phi `$, where $`\varphi _c`$ is the classical field configuration, and integrate over the deviation from $`\varphi _c`$. The action can be written $$S[\varphi _c+\phi ]=d^4x\left(\frac{1}{2}(_\mu \varphi _c)^2\frac{1}{2}m^2\varphi _c^2+J\varphi _c\right)+d^4x\left(\frac{1}{2}(_\mu \phi )^2\frac{1}{2}m^2\phi ^2\right),$$ where as usual there is no term linear in $`\phi `$ since $`\varphi _c`$ by definition extremizes the classical action. So $$N_0=C\mathrm{exp}id^4x\left(\frac{1}{2}(_\mu \varphi _c)^2\frac{1}{2}m^2\varphi _c^2+J\varphi _c\right),$$ where $$C=𝒟\phi \mathrm{exp}id^4x\left(\frac{1}{2}(_\mu \phi )^2\frac{1}{2}m^2\phi ^2\right).$$ $`C`$ is independent of $`J`$ and will cancel in $`Z`$. (Indeed, the denominator is equal to $`C`$.) Using the fact that $`\varphi _c`$ obeys the classical equation $$(^2+m^2)\varphi _c=J,$$ we can write $$N_0=C\mathrm{exp}\frac{i}{2}d^4xJ(x)\varphi _c(x).$$ Finally, we can write $`\varphi _c`$ in terms of the Klein-Gordon Green’s function, defined by $$(^2+m^2)\mathrm{\Delta }_F(x,x^{})=i\delta ^4(xx^{}).$$ It is $$\varphi _c(x)=id^4x\mathrm{\Delta }_F(x,x^{})J(x^{}),$$ so $$\overline{)Z_0=\frac{N_0}{C}=\mathrm{exp}\frac{1}{2}d^4xd^4x^{}J(x)\mathrm{\Delta }_F(x,x^{})J(x^{}).}$$ The Green’s function is found by solving its equation in 4-momentum space; the result is $$\mathrm{\Delta }_F(x,x^{})=\frac{d^4k}{(2\pi )^4}\frac{i}{k^2m^2+iϵ}e^{ik(xx^{})}=\mathrm{\Delta }_F(xx^{}),$$ adopting the same pole prescription as in (47). Note that $`\mathrm{\Delta }_F`$ is an even function, $`\mathrm{\Delta }_F(xx^{})=\mathrm{\Delta }_F(x^{}x)`$. Let us calculate a couple of Green’s functions. These calculations are reminiscent of those at the end of Section 6. As a first example, consider $$G_0^{(2)}(x_1,x_2)=0\left|T\widehat{\varphi }(x_1)\widehat{\varphi }(x_2)\right|0=\frac{1}{i^2}\left(\frac{\delta ^2}{\delta J(x_1)\delta J(x_2)}Z_0[J]\right)|_{J=0}.$$ Expanding $`Z_0`$ in powers of $`J`$, $$Z_0[J]=1\frac{1}{2}d^4xd^4x^{}J(x)\mathrm{\Delta }_F(xx^{})J(x^{})+o(J^4).$$ The term quadratic in $`J`$ is the only one that survives both differentiation (which kills the “1”) and the setting of $`J`$ to zero (which kills all higher-order terms). So $$G_0^{(2)}(x_1,x_2)=\frac{1}{i^2}\frac{\delta ^2}{\delta J(x_1)\delta J(x_2)}\left(\frac{1}{2}d^4xd^4x^{}J(x)\mathrm{\Delta }_F(xx^{})J(x^{})\right).$$ There arise two identical terms, depending on which derivative acts on which $`J`$. The result is $$G_0^{(2)}(x_1,x_2)=\mathrm{\Delta }_F(x_1x_2).$$ So the Green’s function (or two-point function) in the quantum field theory sense is also the Green’s function in the usual differential-equations sense. As a second example, the four-point Green’s function is $$G_0^{(4)}(x_1,x_2,x_3,x_4)=\frac{1}{i^4}\left(\frac{\delta }{\delta J(x_1)}\mathrm{}\frac{\delta }{\delta J(x_4)}\mathrm{exp}\frac{1}{2}d^4xd^4x^{}J(x)\mathrm{\Delta }_F(x,x^{})J(x^{})\right)|_{J=0}.$$ This time the only part of the exponential that contributes is the term with four $`J`$’s. $$G_0^{(4)}(x_1,x_2,x_3,x_4)=\frac{\delta }{\delta J(x_1)}\mathrm{}\frac{\delta }{\delta J(x_4)}\frac{1}{2}\left(\frac{1}{2}d^4xd^4x^{}J(x)\mathrm{\Delta }_F(x,x^{})J(x^{})\right)^2.$$ There are $`4!=24`$ terms, corresponding to the number of ways of associating the derivatives with the $`J`$’s. In 8 of them, the Green’s functions which arise are $`\mathrm{\Delta }_F(x_1x_2)\mathrm{\Delta }_F(x_3x_4)`$, and so on. The result is $`G_0^{(4)}(x_1,x_2,x_3,x_4)`$ $`=`$ $`\mathrm{\Delta }_F(x_1x_2)\mathrm{\Delta }_F(x_3x_4)+\mathrm{\Delta }_F(x_1x_3)\mathrm{\Delta }_F(x_2x_4)`$ (50) $`+\mathrm{\Delta }_F(x_1x_4)\mathrm{\Delta }_F(x_2x_3),`$ which can be represented diagramatically as in Figure 13. ### 8.2 Interacting scalar field theory. Usually, if the Lagrangian has a term beyond quadratic we can no longer evaluate exactly the functional integral, and we must resort to perturbation theory. The generating functional method is tailor-made to do this in a systematic fashion. To be specific, consider $`\varphi ^4`$ theory, defined by the Lagrangian density $$=\frac{1}{2}(_\mu \varphi )^2\frac{1}{2}m^2\varphi ^2\frac{\lambda }{4!}\varphi ^4.$$ Then the generating functional is (up to an unimportant constant: we will normalize ultimately so that $`Z[J=0]=1`$) $$Z[J]=C𝒟\varphi \mathrm{exp}id^4x\left(\frac{1}{2}(_\mu \varphi )^2\frac{1}{2}m^2\varphi ^2\frac{\lambda }{4!}\varphi ^4+J\varphi \right).$$ Because of the quartic term, we cannot evaluate the functional integral exactly. But we can use a trick we first saw when discussing perturbation theory in quantum mechanics: replacing the higher-order term by a functional derivative with respect to $`J`$: $`Z[J]`$ $`=`$ $`C{\displaystyle 𝒟\varphi \mathrm{exp}\left\{i\frac{\lambda }{4!}d^4x\varphi ^4\right\}\mathrm{exp}id^4x\left(\frac{1}{2}(_\mu \varphi )^2\frac{1}{2}m^2\varphi ^2+J\varphi \right)}`$ $`=`$ $`C{\displaystyle 𝒟\varphi \mathrm{exp}\left\{i\frac{\lambda }{4!}d^4x\left(\frac{1}{i}\frac{\delta }{\delta J(x)}\right)^4\right\}\mathrm{exp}id^4x\left(\frac{1}{2}(_\mu \varphi )^2\frac{1}{2}m^2\varphi ^2+J\varphi \right)}.`$ We can pull the first exponential out of the integral; the functional integral which remains is that for $`Z_0`$. Adjusting the constant $`C`$ so that $`Z[J=0]=0`$, we get $$Z[J]=\frac{\mathrm{exp}\left\{i\frac{\lambda }{4!}d^4x\left(\frac{1}{i}\frac{\delta }{\delta J(x)}\right)^4\right\}\mathrm{exp}\frac{1}{2}d^4xd^4x^{}J(x)\mathrm{\Delta }_F(x,x^{})J(x^{})}{\left(\mathrm{exp}\left\{i\frac{\lambda }{4!}d^4x\left(\frac{1}{i}\frac{\delta }{\delta J(x)}\right)^4\right\}\mathrm{exp}\frac{1}{2}d^4xd^4x^{}J(x)\mathrm{\Delta }_F(x,x^{})J(x^{})\right)|_{J=0}}.$$ This expression now enables us to compute a perturbative expansion for any Green’s function we desire. This is a rather mechanical job, and the only way to learn it is by doing lots of examples. To illustrate the method, let us look at $`G^{(2)}(x_1,x_2)`$ to the first nontrivial order in $`\lambda `$. We have $$G^{(2)}(x_1,x_2)=\frac{\left(\frac{1}{i^2}\frac{\delta ^2}{\delta J(x_1)\delta J(x_2)}\mathrm{exp}\left\{i\frac{\lambda }{4!}d^4x\left(\frac{1}{i}\frac{\delta }{\delta J(x)}\right)^4\right\}\mathrm{exp}\frac{1}{2}J_a\mathrm{\Delta }_{Fab}J_b\right)|_{J=0}}{\left(\mathrm{exp}\left\{i\frac{\lambda }{4!}d^4x\left(\frac{1}{i}\frac{\delta }{\delta J(x)}\right)^4\right\}\mathrm{exp}\frac{1}{2}J_a\mathrm{\Delta }_{Fab}J_b\right)|_{J=0}},$$ where as in Section 6 $`\mathrm{}`$ implies integration over the positions of the $`J`$’s. In both numerator and denominator, we can expand both exponentials. The only terms that survive are those that have the same total number of derivatives and $`J`$’s. Let us look at the term linear in $`\lambda `$ in the numerator. There are six derivatives, so we need the term from the expansion of the second exponential with six $`J`$’s. For this term, we get the following expression for the numerator: $$\frac{\delta ^2}{\delta J(x_1)\delta J(x_2)}\left(i\frac{\lambda }{4!}\right)d^4x\left(\frac{\delta }{\delta J(x)}\right)^4\frac{1}{3!}\left(\frac{1}{2}\right)^3J_a\mathrm{\Delta }_{Fab}J_bJ_c\mathrm{\Delta }_{Fcd}J_dJ_e\mathrm{\Delta }_{Fef}J_f.$$ There are now a total of $`6!=720`$ terms! However, only two distinct analytical expressions result. The first of these arises if the derivatives at $`x_1`$ and $`x_2`$ act on different $`\mathrm{}`$’s. A little combinatorial head-scratching tells us that there are 576 such terms, yielding the following expression: $$\frac{i\lambda }{2}d^4x\mathrm{\Delta }_F(x_1x)\mathrm{\Delta }_F(xx)\mathrm{\Delta }_F(xx_2),$$ (51) which can be represented pictorially as in Figure 14 (a). The only other expression arises when the derivatives at $`x_1`$ and $`x_2`$ act on the same $`\mathrm{}`$. This accounts for the remaining 144 terms; the analytic form which results is $$\frac{i\lambda }{8}\mathrm{\Delta }_F(x_1x_2)d^4x\mathrm{\Delta }_F(xx)^2,$$ (52) corresponding to the diagram in Figure 14(b). The denominator can be evaluated in a similar fashion; the Green’s function to order $`\lambda `$ is $$G^{(2)}(x_1,x_2)=\frac{\begin{array}{c}\{\mathrm{\Delta }_F(x_1x_2)\frac{i\lambda }{2}d^4x\mathrm{\Delta }_F(x_1x)\mathrm{\Delta }_F(xx)\mathrm{\Delta }_F(xx_2)\\ \frac{i\lambda }{8}\mathrm{\Delta }_F(x_1x_2)d^4x\mathrm{\Delta }_F(xx)^2+o(\lambda ^2)\}\end{array}}{1\frac{i\lambda }{8}d^4x\mathrm{\Delta }_F(xx)^2+o(\lambda ^2)}.$$ Since we have only computed the numerator and denominator to order $`\lambda `$, we can rewrite this expression in the following way: $$G^{(2)}(x_1,x_2)=\frac{\left(\begin{array}{c}\left\{\mathrm{\Delta }_F(x_1x_2)\frac{i\lambda }{2}d^4x\mathrm{\Delta }_F(x_1x)\mathrm{\Delta }_F(xx)\mathrm{\Delta }_F(xx_2)+o(\lambda ^2)\right\}\\ \times \left\{1\frac{i\lambda }{8}d^4x\mathrm{\Delta }_F(xx)^2+o(\lambda ^2)\right\}\end{array}\right)}{1\frac{i\lambda }{8}d^4x\mathrm{\Delta }_F(xx)^2+o(\lambda ^2)}.$$ We can now cancel the second factor in the numerator against the denominator, to order $`\lambda `$, resulting in $$G^{(2)}(x_1,x_2)=\mathrm{\Delta }_F(x_1x_2)\frac{i\lambda }{2}d^4x\mathrm{\Delta }_F(x_1x)\mathrm{\Delta }_F(xx)\mathrm{\Delta }_F(xx_2)+o(\lambda ^2).$$ This factorization of the numerator into a part containing no factors independent of the external position times the denominator occurs to all orders, as can be proven fairly cleanly via a combinatoric argument. The conclusion is that so-called disconnected parts (parts of diagrams not connected to any external line) cancel from Green’s functions, a fact which simplifies greatly the calculation of these objects. It cannot be overemphasized that there are only three ways to get accustomed to this formalism: practice, practice, and practice. Other reasonable exercises are the calculation of $`G^{(2)}`$ to order $`\lambda ^2`$ and the calculation of $`G^{(4)}`$ to order $`\lambda ^2`$. $`\varphi ^3`$ theory is also a useful testing ground for the techniques discussed in this section. ## 9 Instantons in Quantum Mechanics ### 9.1 General discussion It has already been briefly mentioned that in quantum mechanics certain aspects of a problem can be overlooked in a perturbative treatment. One example occurs if we have a harmonic oscillator with a cubic anharmonic term: $`V(q)=\frac{1}{2}m\omega ^2q^2+\lambda q^3`$ (Figure 15). We can calculate corrections to harmonic oscillator wave functions and energies perturbatively in $`\lambda `$, to any desired order, blissfully ignorant of a serious pathology in the model. As can be seen from Figure 15, this model has no ground state: the potential energy is unbounded as $`q\mathrm{}`$, a point completely invisible to perturbation theory. A second example is the double-well potential, $`V(q)=\frac{\lambda }{4!}(q^2a^2)^2`$ (Figure 16). There are two classical ground states. We can ignore this fact and expand $`V`$ about one of the minima; it then takes the form of a harmonic oscillator about that minimum plus anharmonic terms (both cubic and quartic). We can then compute perturbative corrections to the wave functions and energies, and never see any evidence of the other minimum. Were we to expand about the other minimum, we would produce an identical set of perturbative corrections. By symmetry the ground state energies calculated perturbatively to any order will be the same for the expansions about the two minima, so it appears that we have degenerate ground states. But in fact the ground state is not degenerate: a nonperturbative energy splitting separates the true ground state (an even function of $`q`$) from the first excited state (an odd function); this splitting is not seen in perturbation theory. We will examine this second example using PIs, the main goal being to calculate the energy splitting between the two candidate ground states. Let us first recall the PI expression for the Euclidean propagator: $$K_E(q^{},\frac{\beta }{2};q,\frac{\beta }{2})=q^{}\left|e^{\beta H/\mathrm{}}\right|q=𝒟qe^{S_E/\mathrm{}},$$ where $$S_E=_{\beta /2}^{\beta /2}𝑑\tau \left(\frac{1}{2}m\dot{q}^2+V(q)\right).$$ Henceforth, we will set $`m1`$. $`K_E`$ is useful because we can write it as $$K_E=\underset{n}{}q^{}|nn|qe^{\beta E_n/\mathrm{}};$$ (53) in the limit $`\beta \mathrm{}`$, this term will be dominated by the lowest-energy states. I say “states” here rather than “state” because we must calculate the two lowest-energy eigenvalues to get the splitting of the (perturbatively degenerate) lowest-energy states in the double-well potential. We will evaluate the PI using an approximation known as the semiclassical approximation, or alternatively as the method of steepest descent. To illustrate it, consider the following integral $$I=_{\mathrm{}}^{\mathrm{}}𝑑xe^{S(x)/\mathrm{}},$$ where $`S(x)`$ is a function with several local minima (Figure 17). Suppose we are interested in this integral as $`\mathrm{}0`$. Then the integral will be dominated by the minima of $`S`$; we can approximate it by a series of Gaussian integrations, one for each minimum of $`S`$. If $`x_i`$ is such a minimum, then in its vicinity $`S(x)S(x_i)+\frac{1}{2}(xx_i)^2S^{\prime \prime }(x_i)`$; we can write $$II_1+I_2+I_3+\mathrm{},$$ (54) where $`I_i`$ $`=`$ $`{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x\mathrm{exp}[S(x_i)+{\displaystyle \frac{1}{2}}(xx_i)^2S^{\prime \prime }(x_i)]/\mathrm{}`$ $`=`$ $`e^{S(x_i)/\mathrm{}}\sqrt{{\displaystyle \frac{2\pi \mathrm{}}{S^{\prime \prime }(x_i)}}}.`$ Anharmonicities of $`S`$ appear as corrections of order $`\mathrm{}`$ to $`I`$. (This can be easily seen, for example, by considering a specific case such as $`S(x)=ax^2+bx^4`$.) We will compute the PI (53) in the semi-classical approximation, where the analog of the $`x_i`$ in the above example will be classical paths (extremum of the action $`S_E[q]`$). Suppose, then, that $`q_c(\tau )`$ is the classical solution to the problem $$\frac{d^2}{d\tau ^2}q=\frac{V(q)}{q},q(\beta /2)=q,q(\beta /2)=q^{}.$$ We can write $`q(\tau )=q_c(\tau )+y(\tau )`$; the action is $`S_E[q_c+y]`$ $`=`$ $`{\displaystyle _{\beta /2}^{\beta /2}}𝑑\tau \left({\displaystyle \frac{1}{2}}\dot{q}_c+\dot{y}^2+V(q_c+y)\right)`$ (55) $`=`$ $`{\displaystyle _{\beta /2}^{\beta /2}}𝑑\tau \left({\displaystyle \frac{1}{2}}\dot{q}_{c}^{}{}_{}{}^{2}+V(q_c)\right)+(\text{linear in }y)`$ $`+{\displaystyle _{\beta /2}^{\beta /2}}𝑑\tau \left({\displaystyle \frac{1}{2}}\dot{y}^2+{\displaystyle \frac{1}{2}}V^{\prime \prime }(q_c)y^2\right)+\mathrm{}.`$ The term linear in $`y`$ vanishes for the usual reason, and the higher order terms not written down are of cubic or higher order in $`y`$. Neglecting these (which give order $`\mathrm{}`$ corrections to the PI), the propagator becomes $$K_E=𝒟qe^{S_E/\mathrm{}}=e^{S_E[q_c]/\mathrm{}}𝒟y\mathrm{exp}𝑑\tau \left(\frac{1}{2}\dot{y}^2+\frac{1}{2}V^{\prime \prime }(q_c)y^2\right)/\mathrm{}.$$ The functions $`y(\tau )`$ over which we integrate satisfy the boundary conditions $`y(\beta /2)=y(\beta /2)=0`$. The PI, being Gaussian, can be done exactly; it is not as straightforward as the harmonic oscillator PI since $`V^{\prime \prime }(q_c)`$ depends on $`\tau `$. While we have often managed to avoid evaluating PIs, here we must evaluate it. (Unfortunately, this is rather difficult.) To this end, we can use a generalization of the Fourier expansion technique mentioned in Section 2.2.2. We can rewrite the action as $$S_E=𝑑\tau \left(\frac{1}{2}\dot{y}^2+\frac{1}{2}V^{\prime \prime }(q_c)y^2\right)=\frac{1}{2}𝑑\tau y\left(\frac{d^2}{d\tau ^2}+V^{\prime \prime }(q_c)\right)y.$$ (56) The Schroedinger-like equation $$\left(\frac{d^2}{d\tau ^2}+V^{\prime \prime }(q_c)\right)y=\lambda y,y(\beta /2)=y(\beta /2)=0$$ has a complete, orthonormal set of solutions; let the solutions and eigenvalues be $`y_k(\tau )`$ and $`\lambda _k`$, respectively. The orthonormality relation is $$_{\beta /2}^{\beta /2}𝑑\tau y_k(\tau )y_l(\tau )=\delta _{kl}.$$ Then we can substitute $`y(\tau )=_ka_ky_k(\tau )`$ in (56), giving $$S_E=\frac{1}{2}𝑑\tau \underset{k}{}a_ky_k\left(\frac{d^2}{d\tau ^2}+V^{\prime \prime }(q_c)\right)\underset{l}{}a_ly_l=\frac{1}{2}\underset{k,l}{}a_ka_l\lambda _l𝑑\tau y_ky_l=\frac{1}{2}\underset{k}{}a_k^2\lambda _k.$$ The PI can now be written as an integral over all possible values of the coefficients $`\{a_k\}`$. This gives $$K_E=J^{}\underset{k}{}da_ke^{_ka_k^2\lambda _k/2\mathrm{}},$$ (57) where $`J^{}`$ is the Jacobian of the transformation from $`y(\tau )`$ to $`\{a_k\}`$. (57) is a product of uncoupled Gaussian integrals; the result is $$K_E=J^{}\underset{k}{}\left(\frac{2\pi \mathrm{}}{\lambda _k}\right)^{1/2}=J^{}\underset{k}{}(2\pi \mathrm{})^{1/2}(\underset{k}{}\lambda _k)^{1/2}=J^{}\underset{k}{}(2\pi \mathrm{})^{1/2}\stackrel{1/2}{det}\left(\frac{d^2}{d\tau ^2}+V^{\prime \prime }(q_c)\right),$$ where we have written the product of eigenvalues as the determinant of the Schroedinger operator on the space of functions vanishing at $`\pm \beta /2`$. We can write $`J=J^{}_k(2\pi \mathrm{})^{1/2}`$, giving $$K_E=J\stackrel{1/2}{det}\left(\frac{d^2}{d\tau ^2}+V^{\prime \prime }(q_c)\right)(1+o(\mathrm{})),$$ where the $`o(\mathrm{})`$ corrections can in principle be computed from the neglected beyond-quadratic terms in (55). We will not be concerned with these corrections, and henceforth we will drop the $`(1+o(\mathrm{}))`$. ### 9.2 Single Well in the Semiclassical Approximation Before looking at the double well, it is worthwhile examining the single well, defined be $$V(q)=\frac{1}{2}\omega ^2q^2+\frac{\lambda }{4!}q^4.$$ The classical equation is $$\frac{d^2}{d\tau ^2}q=V^{}(q).$$ Note that this is the equation of motion for a particle moving in a potential $`V(q)`$. If we choose the initial and final points $`q=q^{}=0`$, then the classical solution is simply $`q_c(\tau )=0`$; furthermore, $`V^{\prime \prime }(q_c)=V^{\prime \prime }(0)=\omega ^2`$, and $$K_E=J\stackrel{1/2}{det}\left(\frac{d^2}{d\tau ^2}+\omega ^2\right).$$ The evaluation of the determinant is not terribly difficult (the eigenvalues can be easily found; their product can be found in a table of mathematical identities); the result, for large $`\beta `$, is $$K_E=\left(\frac{\omega }{\pi \mathrm{}}\right)^{1/2}e^{\beta \omega /2}.$$ From (53), we can extract the ground state energy since, for large $`\beta `$, $`K_E\mathrm{exp}E_0\beta /\mathrm{}`$. We find $`E_0=\mathrm{}\omega /2`$ up to corrections of order $`\lambda \mathrm{}^2`$. We have discovered an incredibly complicated way of calculating the ground state energy of the harmonic oscillator! ### 9.3 Instantons in the Double Well Potential Let us now study a problem of much greater interest: the double well. We will see that configurations known as “instantons” make a non-perturbative correction to the energies. We wish to evaluate the PI $$K_E=_{q,\beta /2}^{q^{},\beta /2}𝒟qe^{S_E},$$ where $$S_E=𝑑\tau \left(\frac{1}{2}\dot{q}^2+\frac{\lambda }{4!}(q^2a^2)^2\right),$$ for $`\beta \mathrm{}`$. As explained above, the PI is dominated by minima of $`S_E`$, i.e., by classical solutions. The classical equation corresponds to a particle moving in the potential $`V(q)`$ (Figure 18); the “energy” $`E=\frac{1}{2}\dot{q}^2V(q)`$ is conserved. Let us examine classical solutions, taking the boundary values $`q,q^{}`$ of the classical solution corresponding to the maxima of $`V`$, $`\pm a`$. In the limit $`\beta \mathrm{}`$, these will be solutions of zero “energy”, since as $`\tau \pm \mathrm{}`$ both the kinetic and potential “energy” vanish. First, if $`q=q^{}=a`$ (an identical argument applies if $`q=q^{}=a`$), the obvious classical solution is $`q(\tau )=a`$; a quadratic approximation about this constant solution would be identical to the single-well case discussed above. But what if $`q=a`$ and $`q^{}=a`$ (or vice-versa)? Then the obvious classical solution corresponds to the particle initially sitting atop the maximum of $`V`$ at $`a`$, rolling towards $`q=0`$ after a very long (infinite, in the limit $`\beta \mathrm{}`$) time, and ending up at rest at the other maximum of $`V`$ as $`\tau \mathrm{}`$ (Figure 19). We can get the analytical form of this solution: setting $`E0`$, we have $$\frac{1}{2}\dot{q}^2=V(q),\text{or}\frac{dq}{d\tau }=\pm \sqrt{\frac{\lambda }{12}}(q^2a^2).$$ There are a family of solutions interpolating between $`a`$ and $`a`$: $$q(\tau )=a\mathrm{tanh}\frac{\omega }{2}(\tau \tau _0),$$ (58) where $`\omega =\sqrt{\lambda a^2/3}`$ and where $`\tau _0`$ is an integration constant which corresponds to the time at which the solution crosses $`q=0`$. This solution is much like a topological soliton in field theory, except that it is localized in time rather than in space. One could argue that the solution doesn’t appear to be localized: $`q`$ goes to different values as $`\tau \pm \mathrm{}`$. But these are just different, but physically equivalent, ground states, so we can say that the instanton is a configuration which interpolates between two ground states; the system is in a ground state except for a brief time – an “instant”. For this reason, the solution is known as an instanton. I called the two solutions $`q(\tau )=a`$ and $`q(\tau )=a\mathrm{tanh}\frac{\omega }{2}(\tau \tau _0)`$ the obvious classical solutions because there are an infinite number of approximate classical solutions which are potentially important in the PI. Since the instanton is localized in time, and since the total time interval $`\beta `$ is very large (in particular, much larger than the instanton width), a series of widely-separated instantons and anti-instantons (configurations interpolating between $`+a`$ and $`a`$) is also a solution, up to exponentially small interactions between neighbouring instantons and anti-instantons. Such a configuration is shown in Figure 20, where the horizontal scale has been determined by the duration of imaginary time $`\beta `$; on this scale the instanton and anti-instanton appear as step functions. It is clear than an instanton must be followed by an anti-instanton, and that if the asymptotic values of the position are $`+a`$ and $`+a`$ the classical solution must contain anti-instanton-instanton pairs whereas if they are $`a`$ and $`+a`$ we need an extra instanton at the beginning. Let us choose first limiting values $`q(\beta /2)=q(+\beta /2)=+a`$. Then we are interested in $$K_E=_{a,\beta /2}^{a,\beta /2}𝒟qe^{S_E}.$$ In the spirit of (54), in the steepest-descent approximation $`K_E`$ is equal to the sum of PIs evaluated about all classical solutions. The classical solutions are: $`q_c(\tau )=a`$; $`q_c=`$anti-instanton-instanton$`AI`$; $`q_c=AIAI`$; etc., where the positions of the $`A`$s and $`I`$s are not determined, and must be integrated over. Schematically, we may write $$K_E=K_E^0+K_E^2+K_E^4+\mathrm{},$$ (59) where the superscript denotes the total number of $`I`$s or $`A`$s. Let us discuss the first couple of contributions in some detail. $`q_c=a`$: This case is essentially equivalent to the single-well case discussed above, and we get $$K_E^0=\sqrt{\frac{\omega }{\pi \mathrm{}}}e^{\beta \omega /2},$$ where $`\omega =(\lambda a^2/3)^{1/2}`$ is the frequency of small oscillations about the minimum of $`V`$. $`q_c=AI`$: This case is rather more interesting (that is to say, complicated!). Let us suppose that the classical solution around which we expand consists of an anti-instanton at time $`\tau _1`$ and an instanton at $`\tau _2`$ (Figure 21); clearly $`\tau _2>\tau _1`$. Then we can write $`q=q_c+y`$, and $$S_E[q]=S_E[q_c]+S_E^{\mathrm{quad}}[y].$$ We can evaluate $`S_E[q_c]`$: it is twice the action of a single instanton (assuming the $`I`$ and $`A`$ are sufficiently far apart that any interaction is negligible): $`S_E[q_c]=2S_E^{\mathrm{inst}}`$. The one-instanton action $`S_E^{\mathrm{inst}}`$ is $$S_E^{\mathrm{inst}}=𝑑\tau \left(\frac{1}{2}\dot{q}^2+V(q)\right)|_{\mathrm{inst}}=2𝑑\tau V(q)|_{\mathrm{inst}}.$$ With the instanton profile given by (58), the result is $$S_E^{\mathrm{inst}}=\sqrt{\frac{\lambda }{3}}\frac{2a^3}{3}.$$ To evaluate the PI with the action $`S_E^{\mathrm{quad}}[y]`$, let us divide the imaginary time interval into two semi-infinite regions $`I`$ and $`II`$, where the boundary between the two regions is between and well away from the $`A`$ and the $`I`$ (Figure 22). Then we can write $$K_E^2=\frac{\beta ^2}{2}e^{2S_E^{\mathrm{inst}}}_{I+II}𝒟ye^{S_E^{\mathrm{quad}}/\mathrm{}}.$$ (60) Here the first factor represents integration over the positions of the $`A`$ and $`I`$ (remember that the $`A`$ must be to the left of the $`I`$!). The quadratic action can be written $$S_E^{\mathrm{quad}}=S_{E}^{\mathrm{quad}}{}_{I}{}^{}+S_{E}^{\mathrm{quad}}{}_{II}{}^{},$$ where $`S_{E}^{\mathrm{quad}}{}_{I}{}^{}`$ is the quadratic action in the presence of an anti-instanton and $`S_{E}^{\mathrm{quad}}{}_{II}{}^{}`$ is that in the presence of an instanton. Then the PI separates into two factors: $$_{I+II}𝒟ye^{S_E^{\mathrm{quad}}/\mathrm{}}=_I𝒟ye^{S_{E}^{\mathrm{quad}}{}_{I}{}^{}/\mathrm{}}_{II}𝒟ye^{S_{E}^{\mathrm{quad}}{}_{II}{}^{}/\mathrm{}},$$ (61) where there is an implied integration over the intermediate position at the boundary of the two regions. The quadratic no-instanton PI also separates into two factors: $$𝒟ye^{S_E^{\mathrm{quad},0}/\mathrm{}}=_I𝒟ye^{S_{E}^{\mathrm{quad},0}{}_{I}{}^{}/\mathrm{}}\times _{II}𝒟ye^{S_{E}^{\mathrm{quad},0}{}_{II}{}^{}/\mathrm{}},$$ (62) where the superscript “0” denotes that this is the PI about a no-instanton (constant) background. We can combine (61) and (62) to give: $$_{I+II}𝒟ye^{S_E^{\mathrm{quad}}/\mathrm{}}=𝒟ye^{S_E^{\mathrm{quad},0}/\mathrm{}}\frac{_I𝒟ye^{S_{E}^{\mathrm{quad}}{}_{I}{}^{}/\mathrm{}}}{_I𝒟ye^{S_{E}^{\mathrm{quad},0}{}_{I}{}^{}/\mathrm{}}}\frac{_{II}𝒟ye^{S_{E}^{\mathrm{quad}}{}_{II}{}^{}/\mathrm{}}}{_{II}𝒟ye^{S_{E}^{\mathrm{quad},0}{}_{II}{}^{}/\mathrm{}}}.$$ (63) But $$\frac{_I𝒟ye^{S_{E}^{\mathrm{quad}}{}_{I}{}^{}/\mathrm{}}}{_I𝒟ye^{S_{E}^{\mathrm{quad},0}{}_{I}{}^{}/\mathrm{}}}=\frac{𝒟ye^{S_E^{\mathrm{quad}}/\mathrm{}}}{𝒟ye^{S_E^{\mathrm{quad},0}/\mathrm{}}}$$ (64) and similarly for the last factor in (63), so we obtain $$_{I+II}𝒟ye^{S_E^{\mathrm{quad}}/\mathrm{}}=\sqrt{\frac{\omega }{\pi \mathrm{}}}e^{\beta \omega /2}R^2,$$ where $`R`$ is the ratio of the PI in the presence and absence of an instanton (or, equivalently, anti-instanton) given in (64). Substituting this into (60), $$K_E^2=e^{2S_E^{\mathrm{inst}}/\mathrm{}}\sqrt{\frac{\omega }{\pi \mathrm{}}}e^{\beta \omega /2}R^2\frac{\beta ^2}{2}.$$ A similar argument gives $$K_E^4=e^{4S_E^{\mathrm{inst}}/\mathrm{}}\sqrt{\frac{\omega }{\pi \mathrm{}}}e^{\beta \omega /2}R^4\frac{\beta ^4}{4!},$$ and so on for subsequent terms in the expansion (59). Summing these contributions, we get $`K_E`$ $`=`$ $`\sqrt{{\displaystyle \frac{\omega }{\pi \mathrm{}}}}e^{\beta \omega /2}\left(1+{\displaystyle \frac{\left(\beta Re^{S_E^{\mathrm{inst}}/\mathrm{}}\right)^2}{2!}}+{\displaystyle \frac{\left(\beta Re^{S_E^{\mathrm{inst}}/\mathrm{}}\right)^4}{4!}}+\mathrm{}\right)`$ $`=`$ $`\sqrt{{\displaystyle \frac{\omega }{\pi \mathrm{}}}}e^{\beta \omega /2}\mathrm{cosh}\left(\beta Re^{S_E^{\mathrm{inst}}/\mathrm{}}\right)`$ $`=`$ $`{\displaystyle \frac{1}{2}}\sqrt{{\displaystyle \frac{\omega }{\pi \mathrm{}}}}e^{\beta \omega /2}\left(e^{\beta Re^{S_E^{\mathrm{inst}}/\mathrm{}}}+e^{\beta Re^{S_E^{\mathrm{inst}}/\mathrm{}}}\right).`$ Now we must recall why we’re calculating this object in the first place. The propagator can be written as in (53): $$K_E=\underset{n}{}a|nn|ae^{\beta E_n/\mathrm{}}.$$ By comparing these two expressions we see that the lowest two energies are $$\frac{\mathrm{}\omega }{2}\mathrm{}Re^{S_E^{\mathrm{inst}}/\mathrm{}}\mathrm{and}\frac{\mathrm{}\omega }{2}+\mathrm{}Re^{S_E^{\mathrm{inst}}/\mathrm{}}.$$ So the energy splitting is given by $$\overline{)\mathrm{\Delta }E=2\mathrm{}Re^{S_E^{\mathrm{inst}}/\mathrm{}}.}$$ (65) $`\mathrm{\Delta }E`$ is clearly non-perturbative: it cannot be expanded as a power series in $`\mathrm{}`$ (or, equivalently, in $`\lambda `$). In principle, we should calculate the ratio $$R=\frac{(\text{instanton background PI})}{(\text{constant background PI})}\text{ratio of determinants},$$ but I don’t know how to compute it other than by doing a very arduous, technical calculation; luckily, time will not permit it. The interested reader can consult the book by Sakita for a discussion of this calculation. As a final note, we have calculated the PI with $`q=q^{}=a`$; a good exercise is to do the analogous calculation for $`q=a`$, $`q^{}=a`$. ### 9.4 Instantons in a Periodic Potential Consider a particle moving in a one-dimensional periodic potential (Figure 23). With two minima, as we have seen, instantons enable us to calculate the energy splitting between the lowest-energy states of even and odd parity. In a periodic potential, we will see that a continuum of energies arise. Let us label the classical minima of $`V`$ by an integer, $`j`$. Clearly this model will have solutions analogous to the instantons above, going from any minimum of $`V`$ to the adjacent minimum. We define an instanton as the classical solution going from any $`j`$ to $`j+1`$, and an anti-instanton as that going from $`j`$ to $`j1`$. Then the Euclidean PI to go from $`j=0`$ to $`j=0`$, for instance, can be computed in a manner similar to the calculation of the previous section. This time any number and any order of instantons and anti-instantons are possible, subject to the constraint that $`n_I=n_A`$. A calculation similar to that of the previous section results in the following expression for the propagator: $`K_E(0,\beta /2;0,\beta /2)`$ $`=`$ $`\sqrt{{\displaystyle \frac{\omega }{\pi \mathrm{}}}}e^{\beta \omega /2}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{n!^2}}\left(\underset{Q}{\underset{}{e^{S_E^{\mathrm{inst}}/\mathrm{}}R\beta }}\right)^{2n}`$ $`=`$ $`\sqrt{{\displaystyle \frac{\omega }{\pi \mathrm{}}}}e^{\beta \omega /2}{\displaystyle \underset{n,n^{}=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{Q^n}{n!}}{\displaystyle \frac{Q^n^{}}{n^{}!}}{\displaystyle _0^{2\pi }}{\displaystyle \frac{d\theta }{2\pi }}e^{i\theta (nn^{})}`$ $`=`$ $`\sqrt{{\displaystyle \frac{\omega }{\pi \mathrm{}}}}e^{\beta \omega /2}{\displaystyle _0^{2\pi }}{\displaystyle \frac{d\theta }{2\pi }}{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(Qe^{i\theta })^n}{n!}}{\displaystyle \underset{n^{}=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{(Qe^{i\theta })^n^{}}{n^{}!}}`$ $`=`$ $`\sqrt{{\displaystyle \frac{\omega }{\pi \mathrm{}}}}e^{\beta \omega /2}{\displaystyle _0^{2\pi }}{\displaystyle \frac{d\theta }{2\pi }}\mathrm{exp}\{Qe^{i\theta }\}\mathrm{exp}\{Qe^{i\theta }\}`$ $`=`$ $`\sqrt{{\displaystyle \frac{\omega }{\pi \mathrm{}}}}e^{\beta \omega /2}{\displaystyle _0^{2\pi }}{\displaystyle \frac{d\theta }{2\pi }}\mathrm{exp}\left\{2\beta Re^{S_E^{\mathrm{inst}}/\mathrm{}}\mathrm{cos}\theta \right\}.`$ The derivation of this is a worthwhile exercise. From it, we can read the energies: $$E(\theta )=\frac{\mathrm{}\omega }{2}2\mathrm{}Re^{S_E^{\mathrm{inst}}/\mathrm{}}\mathrm{cos}\theta .$$ The second factor is an expression of the well-known result that the degeneracy is broken nonperturbatively; the energies form a continuum, depending on the value of $`\theta `$. Another application of instantons in quantum mechanics is the phenomenon of tunneling (barrier penetration). The instanton method can be used to calculate the lifetime of a metastable state in a potential of the form depicted in Figure 15. We will not discuss this application. Instantons also appear in (and are by far most useful in) field theory. In certain field theories the space of finite-Euclidean-action configurations separates into distinct topological classes. An instanton is a nontrivial configuration of this type. The necessary topological requirements for this to occur are not hard to satisfy, and the list of theories that have instantons includes the Abelian Higgs model in 1+1 dimensions, the O(3) nonlinear $`\sigma `$-model in 1+1 dimensions, the Skyrme model in 2+1 dimensions, and (most significantly) QCD. Instantons give rise to a host of interesting phenomena depending on the model, including confinement (not in QCD though!), $`\theta `$-vacua, a solution of the U(1) problem in strong interactions, and the decay of a metastable vacuum. Unfortunately time does not permit discussion of these fascinating phenomena. ## 10 Summary and Gross Omissions In this set of lectures the subject of path integrals has been covered starting from scratch, emphasizing explicit calculations in quantum mechanics. This emphasis has its price: I have not had time to cover several things I would have liked to discuss. My hope is that having been subjected to calculations in gory detail for the most part in the relatively familiar context of quantum mechanics, you will be able to study more complicated and interesting applications on your own. Here is a list of some of the important aspects and applications of this subject which I didn’t have time to discuss: 1. Fermi fields and Grassmann functional integration; 2. Gauge theories (gauge fixing and ghosts arise in a particularly elegant way); 3. Feynman’s variational method and application to the polaron (electron moving in a crystal environment); 4. Derivation of the Landau-Ginsburg theory, including application to superconductivity; 5. Instantons in field theory; 6. Critical phenomena. I hope that, in spite of these unforgivable omissions, these lectures have been worthwhile. ## 11 Acknowledgements I wish to thank the Rencontres du Vietnam and the NCST, Hanoi for the invitation to give these lectures, and especially Patrick Aurenche for his boundless energy before, during, and after the School. I am grateful to Jean-Sébastien Caux and Manu Paranjape for discussions and advice during the preparation of these lectures, and to Patrick Irwin, who spotted many weaknesses and errors in this manuscript after I had considered it fit for human consumption. This work was financed in part by the Natural Sciences and Engineering Research Council of Canada. ## 12 References 1. P.A.M. Dirac, Physikalische Zeitschrift der Sowjetunion, Band 3, Heft 1 (1933). 2. R.P. Feynman, Reviews of Modern Physics 20, 367 1948. 3. R.P. Feynman and A.R. Hibbs, Quantum Mechanics and Path Integrals, McGraw-Hill, 1965. 4. R.P. Feynman, Statistical Mechanics: A Set of Lectures, Benjamin, 1972. 5. G. Baym, Lectures on Quantum Mechanics, Benjamin-Cummings, 1973. 6. L.S. Schulman, Techniques and Applications of Path Integration, John Wiley and Sons, 1981. 7. D.J. Amit, Field Theory, the Renormalization Group and Critical Phenomena, 2nd Edition, World Scientific, 1984. 8. L.H. Ryder, Quantum Field Theory, Cambridge University Press, 1985. 9. P. Ramond, Field Theory: A Modern Primer, Second Edition, Addison-Wesley, 1990. 10. E.S. Abers and B.W. Lee, Physics Reports 9C, 1, 1973. 11. B. Sakita, Quantum Theory of Many-Variable Systems and Fields, World Scientific, 1985. 12. S. Coleman, Aspects of Symmetry, Cambridge University Press, 1985. 13. A.M. Tsvelik, Quantum Field Theory in Condensed Matter Physics, Cambridge University Press, 1995. 14. V.N. Popov, Functional Integrals and Collective Excitations, Cambridge University Press, 1987. 15. E. Fradkin, Field Theories of Condensed Matter Systems, Addison-Wesley, 1991.
warning/0004/hep-ph0004035.html
ar5iv
text
# 1 Introduction ## 1 Introduction According to inflationary models , which were first considered to address the flatness, isotropy, and (depending on the particle physics model of the early universe) monopole problems of the hot big-bang model, the universe has undergone several stages during its evolution. During inflation, the energy density of the universe is dominated by the potential energy of the inflaton and the universe experiences a period of superluminal expansion. After inflation, the coherent oscillations of the inflaton, which behave like non-relativistic matter, dominate the energy density of the universe. At some later time these coherent oscillations decay to the fields to which they are coupled, and their energy density is transferred to relativistic particles, the reheating stage, which results in a radiation-dominated FRW universe. After the inflaton decay products are thermalized, the dynamics of the universe will be that of the hot big-bang model. Until recently, reheating was treated as a perturbative, one particle decay of the inflaton with the decay rate $`\mathrm{\Gamma }_d`$ (depending on the microphysics), leading to the simple estimate $`T_R(\mathrm{\Gamma }_dM_{Pl})^{\frac{1}{2}}`$ for the reheat temperature (assuming instant thermalization) . $`T_R`$ should be low enough so that the GUT symmetry is not restored and the original monopole problem is avoided. In supersymmetric models there are even stricter bounds on the reheat temperature. Gravitinos (the spin-$`\frac{3}{2}`$ superpartners of gravitons) with a mass in the range of 100 GeV-1 TeV (in agreement with low energy supersymmetry) decay after the big-bang nucleosynthesis. They are also produced in a thermal bath, predominantly through $`22`$ scatterings of gauge fields and gauginos. This results in the bound TR< 1081010subscript𝑇𝑅< superscript108superscript1010{T}_{R}{\ \lower-1.2pt\vbox{\hbox{\hbox to0.0pt{$<$\hss}\lower 5.0pt\vbox{\hbox{$\sim$}}}}\ }{10}^{8}-{10}^{10} GeV, in order to avoid gravitino overproduction which would destroy the successful predictions of big-bang nucleosynthesis . It has recently been noted that the initial stage of inflaton decay might occur through a complicated and non-perturbative process called parametric resonance, leading to an out-of-equilibrium distribution of final state particles with energies much higher than the inflaton mass (the preheating stage) . This may have important cosmological implications, including gravitino overproduction which is most relevnt to our discussion. Gravitinos can be produced either through the scattering of particles in the preheat distribution , or directly during inflaton oscillations. In particular, it has been recently noted that the second mechanism can lead to efficient production of helicity $`\pm \frac{1}{2}`$ gravitinos, which possibly dominate the abundance of thermally produced gravitinos by several orders of magnitude . However, it is believed that in a wide range of realistic inflationary models the final stage of inflaton decay occurs in the perturbative regime and that the reheat temperature is determined therein. Thus a correct treatment of this stage and the subsequent thermalization of decay products is necessary. After all, parametric resonance may not occur or may be modified, and a mechanism may be found for the dilution of relics which were produced in dangerous abundances during preheating. In any case, the thermal production of such relics, gravitinos in particular, must be in accordance with nucleosynthesis bounds. Our aim in this letter is to give a more detailed examination of thermalization after perturbative decay of the inflaton. In this regime, inflaton decay products have energies higher than the mean thermal energy (if they were instantly thermalized) and a number density lower than the thermal equilibrium value. This implies that thermalization occurs when the interactions which increase the number of particles are at equilibrium. We will note that the inflaton decay is not sudden, but rather a prolonged process which starts very shortly after the end of inflation. We then consider the energy distribution of particles in the plasma of inflaton decay products, and compare the rates for their decay and $`2N`$ particle scatterings (with $`N3`$), as well as $`22`$ particle scatterings which only redistribute the energy among the scattered particles. It will be shown that (cascade) decay is the dominant process which leads to thermalization. We find that, in general, a seed in the plasma of inflaton decay products is thermalized first and then the bulk of the plasma will be thermalized very rapidly by scattering off particles in the thermalized seed. It is emphasized that the thermalization time depends on the the typical mass scale of those inflaton decay products which have interactions of gauge strength. Finally, we will consider the case where the observable sector consists only of the MSSM matter content, and discuss the effect of flat direction vevs on thermalization. ## 2 The Energy Spectrum of Inflaton Decay Products After inflation the inflaton starts its coherent oscillations around the minimum of the scalar potential, and at some later time it will decay to the fields to which it is coupled. In the perturbative regime the decay occurs over many oscillations of the inflaton field. This means that the oscillating inflaton field behaves like non-relativistic matter consisting of a condensate of zero-mode bosons with the mass $`m_I`$. The inflaton decay rate $`\mathrm{\Gamma }_d`$ can then be calculated from the one particle decay channel of these bosons. The value of $`\mathrm{\Gamma }_d`$ depends on the nature of final state particles and their couplings to the inflaton. For example, if the inflaton has a Yukawa coupling $`h`$ to (practically) massless spin-$`\frac{1}{2}`$ fermions $`\mathrm{\Gamma }_d=\frac{h^2}{8\pi }m_I`$; while for the inflaton with gravitationally suppressed coupling to matter $`\mathrm{\Gamma }_d\frac{m_{I}^{}{}_{}{}^{3}}{M_{Pl}^{}{}_{}{}^{2}}`$. Efficient inflaton decay happens at $`t_d=\frac{2}{3H_d}`$ when $`H\mathrm{\Gamma }_d`$. At this time the bulk of the energy density in the coherent oscillations of the inflaton is transferred to relativistic particles with an energy of order $`m_I`$. From then on the universe is radiation-dominated, and the energy of relativistic particles is redshifted as $`a^1t^{\frac{1}{2}}`$; where $`a`$ is the scale factor of the universe. After the thermalization of inflaton decay products is completed the familiar hot big-bang universe is restored. However, inflaton decay does not suddenly happen at $`t_d`$. It is rather a prolonged process which starts once the inflaton oscillations start at $`t=t_I`$, when $`HH_I`$ ($`H_I`$ is the Hubble constant at the end of inflation). The comoving number density of zero-mode inflaton quanta at time $`t`$, $`n_c(t)`$, obeys the relation $`n_c(t)=n_I\mathrm{exp}(\mathrm{\Gamma }_dt)`$; where $`n_I`$ is the inflaton number density at $`t_I`$. The decay products have their largest number density at the earliest times, but they constitute only a tiny fraction of the energy density of the universe for $`t<t_d`$. The inflaton decay becomes efficient at the time $`t_d`$, and at this time the bulk of the energy density is carried by the decay products. Recently it has been noted that consideration of those decay products which were produced before $`t_d`$ can have important implications for thermalization , gravitino production , Affleck-Dine baryogenesis , and electroweak baryogenesis . For $`t_Itt_d`$ the universe is matter-dominated and hence $`H=\frac{2}{3t}`$. In the time interval between $`t`$ and $`t+\frac{2}{3}H^1=2t`$, inflaton decay products are produced and their physical number density $`n`$ at the end of this interval is $$n=n_I[\mathrm{exp}(\mathrm{\Gamma }_dt)\mathrm{exp}(2\mathrm{\Gamma }_dt)](\frac{H}{2H_I})^2.$$ (1) Here we have used the fact that for a matter-dominated universe the redshift factor for the number density of particles is $`(\frac{a}{a_I})^3=(\frac{H}{2H_I})^2`$ (the Hubble constant at the end of interval is $`\frac{H}{2}`$); where $`a_I`$ is the scale factor of the universe at the end of inflation. The redshift factor for the momentum of a particle (equivalently the energy of a relativistic particle) is $`(\frac{a}{a_I})^1=(\frac{H}{2H_I})^{\frac{2}{3}}`$. For $`t\frac{t_d}{2}`$ the above expression becomes $`n\frac{3H\mathrm{\Gamma }_d}{8H_{I}^{}{}_{}{}^{2}}n_I`$. The energy of these particles is between $`(\frac{1}{2})^{\frac{2}{3}}m_I`$ (for the particles produced at the beginning of interval) and $`m_I`$ (for those produced at the end of interval), and can be practically taken as $`m_I`$ for all particles. These particles have the highest energy in the spectrum of inflaton decay products at time $`2t`$. On the other hand, particles which were produced in the interval $`[t_I,2t_I]`$ have the lowest energy in the spectrum. At $`2t`$, their energy has been redshifted (from $`m_I`$) to $`(\frac{H}{H_I})^{\frac{2}{3}}m_I`$, while their number density has been redshifted from $`\frac{3\mathrm{\Gamma }_d}{H_I}n_I`$ to $`\frac{3H^2\mathrm{\Gamma }_d}{8H_{I}^{}{}_{}{}^{2}}n_I`$. After changing $`2t`$ back to $`t`$, and hence $`H`$ to $`2H`$, we find that for $`t<t_d`$ the plasma of inflaton decay products consists of particles with energy $`E`$ and number density $`n`$ $$\begin{array}{c}(\frac{H}{H_I})^{\frac{2}{3}}m_IEm_I\\ \\ \frac{3H^2\mathrm{\Gamma }_d}{4H_{I}^{}{}_{}{}^{3}}n_In\frac{3H\mathrm{\Gamma }_d}{4H_{I}^{}{}_{}{}^{2}}n_I\end{array}$$ (2) such that $$n=(\frac{E}{m_I})^{\frac{3}{2}}\frac{3H\mathrm{\Gamma }_d}{4H_{I}^{}{}_{}{}^{2}}n_I$$ (3) Particles with energy $`E_{Max}=m_I`$ and number density $`n_{Max}=\frac{3H\mathrm{\Gamma }_d}{4H_{I}^{}{}_{}{}^{2}}n_I`$ are produced until $`t_d`$. At this time, the inflaton decay is effectively completed, and almost all of the energy density is carried by relativistic particles. This implies that particles with energy $`m_I`$ are produced in the interval $`[\frac{t_d}{2},t_d]`$ for (practically) the last time. For $`H=\mathrm{\Gamma }_d`$, (2) gives the spectrum of inflaton decay products at $`t_d`$: $$\begin{array}{c}(\frac{\mathrm{\Gamma }_d}{H_I})^{\frac{2}{3}}m_IEm_I\\ \\ \frac{3\mathrm{\Gamma }_{d}^{}{}_{}{}^{3}}{2H_{I}^{}{}_{}{}^{3}}n_In\frac{3\mathrm{\Gamma }_{d}^{}{}_{}{}^{2}}{4H_{I}^{}{}_{}{}^{2}}n_I\end{array}$$ (4) From then on the universe is radiation-dominated, (practically) no more particles with energy $`m_I`$ are produced by inflaton decay, and the energy and the number density of all particles will be redshifted as $`t^{\frac{1}{2}}`$ and $`t^{\frac{3}{2}}`$ respectively. The spectrum of inflaton decay products for $`H<\mathrm{\Gamma }_d`$ is $$\begin{array}{c}(\frac{H}{\mathrm{\Gamma }_d})^{\frac{1}{2}}(\frac{2\mathrm{\Gamma }_d}{H_I})^{\frac{2}{3}}m_IE(\frac{H}{\mathrm{\Gamma }_d})^{\frac{1}{2}}m_I\\ \\ (\frac{H}{\mathrm{\Gamma }_d})^{\frac{3}{2}}\frac{3\mathrm{\Gamma }_{d}^{}{}_{}{}^{2}}{2H_{I}^{}{}_{}{}^{2}}n_In(\frac{H}{\mathrm{\Gamma }_d})^{\frac{3}{2}}\frac{3\mathrm{\Gamma }_{d}^{}{}_{}{}^{2}}{4H_{I}^{}{}_{}{}^{2}}n_I\end{array}$$ (5) and we also have $$n=(\frac{H}{\mathrm{\Gamma }_d})^{\frac{3}{4}}(\frac{E}{m_I})^{\frac{3}{2}}\frac{3\mathrm{\Gamma }_{d}^{}{}_{}{}^{2}}{4H_{I}^{}{}_{}{}^{2}}n_I$$ (6) Let us also find the occupation number of particles, $`f_E`$, as a function of their energy $`E`$ in the spectrum. It is seen from (2) and (5) that particles with energy $`E`$ at time $`t`$ were produced during a short interval $`t`$, at the time when the Hubble constant was $`H_p=(\frac{m_I}{E})^{\frac{3}{2}}`$. The particle momenta at the time of production were in the range $`m_Im_IH_ptpm_I`$, and their number density at that time, from (1), was $`nn_I\mathrm{\Gamma }_dt(\frac{H}{H_p})^2`$. This results in $`f_EE^{\frac{3}{2}}`$ (recall that $`f\frac{n}{p^2p}`$ and here $`p=m_IH_pt`$). It is clear that $`f_E`$ does not change in time since $`n`$ and $`p^2p`$ are both redshifted as $`a^3`$. ## 3 Interactions in the Plasma of Inflaton Decay Products It is clear from (3) and (6) that the number density and the energy density of the plasma is completely dominated by particles with the highest energy $`E_{Max}`$ in the spectrum, both before and after efficient decay of the inflaton. Particles with lower energy have a considerably smaller number density and hence carry a much smaller energy density. It has been pointed out however, that lower energy particles in the spectrum may have an important role in thermalization . Thermalization is a process during which the energy density $`\rho `$ of a distribution of particles remains constant, while their number density $`n`$ changes in such a way that the mean energy of particles reaches its equilibrium value $`T`$. For a distribution of relativistic particles which consists of $`n_B`$ bosonic degrees of freedom and $`n_F`$ fermionic degrees of freedom at thermal equilibrium, we have $`\rho =\frac{\pi ^2}{30}(n_B+\frac{7}{8}n_F)T^4`$ and $`n=\frac{\zeta (3)}{\pi ^2}(n_B+\frac{3}{4}n_F)T^3`$. Therefore, the ratios $`\frac{\rho ^{\frac{1}{4}}}{E}`$ and $`\frac{n^{\frac{1}{3}}}{E}`$ determine the deviation from thermal equilibrium. If these ratios are greater than $`O(1)`$, the number density of particles should decrease, and hence the mean energy increases in order to achieve thermal equilibrium. If they are less than $`O(1)`$, the number density should increase and the mean energy will decrease. In both cases, interactions which change the number of particles should be at equilibrium. Under the assumption that particles in the plasma decay very rapidly after scattering, thermal equilibrium is achieved when particle scatterings are efficient. If only scattering of particles with energy $`E_{Max}`$ (which also have the highest number density $`n_{Max}`$) is considered, the thermalization rate is $$\mathrm{\Gamma }_T\alpha ^2\frac{n_{Max}}{E_{Max}^{}{}_{}{}^{2}}$$ (7) where $`\alpha `$ is the gauge fine structure constant (an $`O(10^2)`$ number). After substituting for $`E_{Max}`$ and $`n_{Max}`$ from (5), we have $$\mathrm{\Gamma }_T\alpha ^2(\frac{H}{\mathrm{\Gamma }_d})^{\frac{1}{2}}\frac{\mathrm{\Gamma }_{d}^{}{}_{}{}^{2}}{H_{I}^{}{}_{}{}^{2}m_{I}^{}{}_{}{}^{2}}n_I$$ (8) Thermalization occurs when $`\mathrm{\Gamma }_TH`$ and is substantially delayed in general, resulting in a low reheat temperature which is consistent with the nucleosynthesis bound on the abundance of thermally produced gravitinos . The scattering rate of particles with energy $`E_s`$ and number density $`n_s`$ off each other is $`\alpha ^2\frac{n_s}{E_{s}^{}{}_{}{}^{2}}`$. From (6) we have $`\frac{n}{E^2}E^{\frac{1}{2}}`$, which implies that particles in a seed with energy $`E_sE_{Max}`$ are scattered off each other at a much higher rate and, therefore the seed could be thermalized much earlier than the bulk of the plasma. This has led to the notion of catalyzed thermalization of the bulk by a thermalized seed which happens if the number density of particles in the seed substantially increases after thermalization, and if particles in the bulk are rapidly scattered off these low energy particles. We have to bear in mind that $`22`$ scatterings do not change the number of particles and only redistribute the energy among the scattered particles. In order to increase the number density of particles in the plasma, the rate for one particle decay and/or $`2N`$ scatterings (with $`N3`$) should be at equilibrium. For a better analysis of thermalization it is therefore necessary to identify the relevant interactions and compare their rates. Here we list the important interactions of a particle with mass $`m`$ and energy $`E_s`$ (assuming $`E_sm`$, a brief note on this will come later): 1- $`22`$ scatterings off other particles in the plasma. It is seen from (3) and (6) that $`nE^{\frac{3}{2}}`$ and hence particles with less energy have also smaller number density and energy density. This implies that the energy density of a seed with energy $`E_s`$ can considerably change only by scattering off particles with energy $`E>E_s`$. For this to happen, it is also necessary that energy is redistributed among the scattred particles. Therefore we estimate the rate for scattering of a particle with energy $`E_s`$ off particles with energy $`E>E_s`$ and number density $`n`$, such that the transferred energy is $`E>E_s`$. The cross-section for this process is $`\sigma \frac{\alpha ^2}{E_sE}`$ <sup>1</sup><sup>1</sup>1Such scatterings occur as a result of effective contact interactions, $`t`$-channel processes or $`s`$-channel processes. Cross-section for the desired process, i.e. scatterings with energy transfer $`E>E_s`$, is readily found to be $`\sigma \frac{\alpha ^2}{E_sE}`$ for a contact interaction (e.g., the quartic coupling of bosons coming from the $`D`$-term part of the action in supersymmetric theories). If $`EE_s`$, the cross-section for the $`t`$-channel processes with energy transfer $`E_s<EE`$ is much greater than $`\frac{\alpha ^2}{E_sE}`$. However, these peocesses correspond to long range forces and are effectively screened in the plasma. Overall, $`\sigma \frac{\alpha ^2}{E_sE}`$ is a reasonable estimate when all processes are taken into account., and the resulting scattering rate will be of order $`\alpha ^2\frac{n}{E_sE}`$. From (3) and (6) we have $`\frac{n}{E}E^{\frac{1}{2}}`$, which implies that the highest scattering rate is off particles with energy $`E_{Max}`$ in he spectrum. Therefore the rate for $`22`$ scatterings which increase the energy of the particle $`E_s`$ is $`\mathrm{\Gamma }_{scatt}\alpha ^2\frac{n_{Max}}{E_sE_{Max}}`$ <sup>2</sup><sup>2</sup>2Since plasma contains particles of all energies in the spectrum we must verify that the inverse scattering is not important. In the perturbative regime of inflaton decay $`f_E<1`$, and the difference $`f_{E_s}f_{E_{Max}}f_{E_s+E}f_{E_{Max}E}`$ determines the direction in which the $`22`$ scatterings proceed. As we found earlier $`f_EE^{\frac{3}{2}}`$, which implies that scattering dominates over inverse scattering.. If $`\mathrm{\Gamma }_{scatt}H`$, particle will acquire an energy much greater than $`E_s`$. However, the energy of the scatterers remains almost unchanged because their number density dominates over the other particles in the plasma. It is seen from (2) and (5) that $`\frac{\mathrm{\Gamma }_{scatt}}{H}t^{\frac{1}{3}}`$ ($`t^{\frac{1}{2}}`$) before (after) efficient inflaton decay. Therefore if other interactions have negligible rates (compared with scatterings), all particles will finally have energies of order $`E_{Max}`$. 2- Decay to other particles through kinematically accessible channels with the rate $`\mathrm{\Gamma }_{decay}\alpha \frac{m^2}{E_s}`$ ($`\frac{m}{E_s}`$ is the time-dilation factor). Such decays increase the number of particles and trigger a chain of cascade decays when $`\mathrm{\Gamma }_{decay}H`$. If $`m^2<\frac{n_{Max}}{E_{Max}}`$ scatterings in (a) increase the energy of the particle before it can decay. These scatterings will also result in a plasma-induced mass-squared of order $`\alpha \frac{n_{Max}}{E_{Max}}`$ <sup>3</sup><sup>3</sup>3Actually $`\alpha ^2\frac{n}{E}`$ is integrtaed over the whole spectrum. It is seen from (3) and (6) that the contribution from particles with energy $`E_{Max}`$ dominates the integral.. Therefore $`m^2`$ is at least of order $`\alpha \frac{n_{Max}}{E_{Max}}`$, even if the mass of the particle in the absence of plasma effects $`m_0`$ is very small. As long as $`m_{0}^{}{}_{}{}^{2}<\alpha \frac{n_{Max}}{E_{Max}}`$ we have Γdecay< ΓscattsubscriptΓ𝑑𝑒𝑐𝑎𝑦< subscriptΓ𝑠𝑐𝑎𝑡𝑡{\Gamma}_{decay}{\ \lower-1.2pt\vbox{\hbox{\hbox to0.0pt{$<$\hss}\lower 5.0pt\vbox{\hbox{$\sim$}}}}\ }{\Gamma}_{scatt}, and $`22`$ scatterings dominate decays. 3- $`2N`$ scatterings off particles with energy $`E_{Max}`$ and number density $`n_{Max}`$, which happen at the rate Γ2N< α3nMaxEMaxEssubscriptΓ2𝑁< superscript𝛼3subscript𝑛𝑀𝑎𝑥subscript𝐸𝑀𝑎𝑥subscript𝐸𝑠{\Gamma}_{2\rightarrow N}{\ \lower-1.2pt\vbox{\hbox{\hbox to0.0pt{$<$\hss}\lower 5.0pt\vbox{\hbox{$\sim$}}}}\ }{\alpha}^{3}{{n}_{Max}\over{E}_{Max}{E}_{s}} <sup>4</sup><sup>4</sup>4The extra factor of $`\alpha `$ appears because at least one more vertex with gauge strength interaction is needed, and there are more $`\frac{1}{2\pi }`$ phase space factors from extra particles in the final state.. These scatterings increase the number of particles too. However, their rate is hierarchically smaller than those of $`22`$ scatterings and decays, and can be neglected in the analysis below. ## 4 Thermalization and Nucleosynthesis Bound on the Reheat Temperature The above arguments establish that the competition between $`22`$ scatterings and (cascade) decays determine the thermalization time and thus the reheat temperature. Depending on the value of $`m_0`$ and parameters of the model for inflation $`n_I`$, $`m_I`$, and $`\mathrm{\Gamma }_d`$, different situations may arise. Our aim here is to study these different possibilities in detail. 1- Consider the case where $`m_{0}^{}{}_{}{}^{2}<\alpha \frac{n_{Max}}{E_{Max}}`$ until very late times. In this case a particle with energy $`E_s`$ is scattered off particles with energy $`E_{Max}`$ before it can decay. Such scatterings increase the energy of the particle by an amount much greater than $`E_s`$. This increase lowers the particle decay rate (recall that $`\mathrm{\Gamma }_{decay}E_{s}^{}{}_{}{}^{1}`$) and renders the decay inefficient. The scatterings become important when $$\alpha ^2\frac{n_{Max}}{E_{min}E_{Max}}H$$ (9) At this time particles with energy $`E_{min}`$ are efficiently scattered and acquire much higher energies. Subsequently the rate for scattering of particles with successively higher energies comes to equilibrium, and they acquire much higher energies as time increases. All particles in the plasma have an energy of order $`E_{Max}`$ when $$\alpha ^2\frac{n_{Max}}{E_{Max}^{}{}_{}{}^{2}}H$$ (10) Shortly after that decay (now with the rate $`\mathrm{\Gamma }_{decay}\alpha ^2\frac{n_{Max}}{E_{Max}^{}{}_{}{}^{2}}`$) becomes effective, and a chain of cascade decays will lead to thermalization of the plasma. This yields $`H_{th}\alpha ^2\frac{n_{Max}}{E_{Max}^{}{}_{}{}^{2}}`$, which after replacing for $`E_{Max}`$ and $`n_{Max}`$ from (5), becomes $$H_{th}\frac{\alpha ^4\mathrm{\Gamma }_{d}^{}{}_{}{}^{3}}{H_{I}^{}{}_{}{}^{4}m_{I}^{}{}_{}{}^{4}}n_{I}^{}{}_{}{}^{2}$$ (11) and the reheat temperature is $$T_R(\frac{\alpha ^4\mathrm{\Gamma }_{d}^{}{}_{}{}^{3}M_{Pl}}{H_{I}^{}{}_{}{}^{4}m_{I}^{}{}_{}{}^{4}}n_I^2)^{\frac{1}{2}}$$ (12) This holds if $`m_{0}^{}{}_{}{}^{2}<\alpha \frac{n_{Max}}{E_{Max}}`$ for at least $`HH_{th}`$, giving $$m_{0}^{}{}_{}{}^{2}\frac{\alpha ^5\mathrm{\Gamma }_{d}^{}{}_{}{}^{4}}{H_{I}^{}{}_{}{}^{6}m_{I}^{}{}_{}{}^{5}}n_{I}^{}{}_{}{}^{3}$$ (13) Therefore if the typical mass scale of particles in the plasma of inflaton decay products satisfy (13), the reheat temperature is given by (12). 2- Consider the case where $`m_0`$ exceeds the bound in (14). In this case $`m_{0}^{}{}_{}{}^{2}`$ catches up with $`\alpha \frac{n_{Max}}{E_{Max}}`$ at an earlier time, and $`m^2m_{0}^{}{}_{}{}^{2}`$ subsequently. This happens for (again we use (5)) $$H_{eq}\frac{H_{I}^{}{}_{}{}^{2}m_I}{\alpha \mathrm{\Gamma }_dn_I}m_{0}^{}{}_{}{}^{2}$$ (14) At this time $`\mathrm{\Gamma }_{scatt}\alpha ^2\frac{n_{Max}}{E_sE_{Max}}`$ is at equilibrium for $$E_s\frac{\alpha ^2\mathrm{\Gamma }_d}{H_{I}^{}{}_{}{}^{2}m_I}n_I$$ (15) Particles with an (initial) energy less than $`E_s`$ in the spectrum have already had efficient scatterings, and therefore have acquired much higher energies. For particles with energies greater than $`E_s`$ decay dominates over scattering, and is efficient when $`\mathrm{\Gamma }_{decay}\alpha \frac{m_{0}^{}{}_{}{}^{2}}{E_s}H`$. At $`HH_{eq}`$ particles with energy $`E_s`$, given by (15), decay efficiently and a seed is self-thermalized after a chain of cascade decays <sup>5</sup><sup>5</sup>5Since the energy of decay products is smaller than the energy of the original particles subsequent decays in the chain occur in an even faster rate (recall that the decay rate is proportional to $`E^1`$). On the other hand, $`\mathrm{\Gamma }_{decay}>\mathrm{\Gamma }_{scatt}`$ regardless of the energy of the particle. Therefore scatterings are not important throughout the chain. At the end of the chain the number density of particles has increased while their energy has decreased. This implies that particles scatter off each other very efficiently, and therefore this can be defined as the time when the seed is thermalized.. The energy and the number density of particles in the thermalized seed are $`T\rho _{s}^{}{}_{}{}^{\frac{1}{4}}(E_sn_s)^{\frac{1}{4}}`$ and $`T^3\rho _{s}^{}{}_{}{}^{\frac{3}{4}}(E_sn_s)^{\frac{3}{4}}`$ respectively, and $`n_sT^3<n_{Max}`$. Two processes are now competing with each other: particles with an energy just above $`E_s`$ in the spectrum scatter off low energy particles in the thermalized seed at rate $`\alpha ^2\frac{T^3}{E_sT}`$, and particles in the thermalized seed scatter off particles with energy $`E_{Max}`$ at rate $`\alpha ^2\frac{n_{Max}}{TE_{Max}}`$. If the first process is more efficient, particles in the bulk loose energy through scatterings off the thermalized seed. This increases their decay rate, triggers cascade decay, and leads to their thermalization. The thermalized seed grows very rapidly in this way until all of the bulk is thermalized. This is called the catalyzed thermalization of the bulk by a thermalized seed . If the second process is more efficient, particles in the thermalized seed acquire energies of order $`E_{Max}`$, and therefore later scatterings off the seed cannot lower the energy of particles in the bulk. In this case, cascade decay of particles with energies greater than $`E_s`$ (hence their thermalization) starts after the Hubble expansion has sufficiently redshifted their energy. Therefore the thermalized seed grows slowly until the first process catches up with the second one. At this time catalyzed thermalization occurs and the bulk becomes thermalized very rapidly. For the first process to be efficient two conditions are necessary. First, its rate should be at equilibrium $$\alpha ^2\frac{T^3}{E_sT}H$$ (16) and, second, it should dominate the second process $$\alpha ^2\frac{T^3}{E_sT}\alpha ^2\frac{n_{Max}}{TE_{Max}}$$ (17) where $`T=(n_sE_s)^{\frac{1}{4}}`$ <sup>6</sup><sup>6</sup>6It is seen that (17) is satisfied when $`T^3\frac{E_s}{E_{Max}}n_{Max}`$. This implies that catalyzed thermalization can occur when $`T^3n_{Max}`$, and justifies the relation $`n_sT^3<n_{Max}`$ in above.. After substituting for $`E_{Max}`$, $`n_{Max}`$, $`E_s`$, and $`n_s`$ from (5) and (6), and using the fact that the seed becomes thermalized after a chain of cascade decays (i.e. when $`\alpha \frac{m_{0}^{}{}_{}{}^{2}}{E_s}H`$), (16) yields $$H(\frac{\alpha ^{18}\mathrm{\Gamma }_{d}^{}{}_{}{}^{5}m_{0}^{}{}_{}{}^{4}n_{I}^{}{}_{}{}^{4}}{m_{I}^{}{}_{}{}^{6}H_{I}^{}{}_{}{}^{6}})^{\frac{1}{7}}$$ (18) and (17) gives $$H(\frac{\alpha ^2m_{0}^{}{}_{}{}^{28}H_{I}^{}{}_{}{}^{8}}{\mathrm{\Gamma }_dm_{I}^{}{}_{}{}^{2}n_{I}^{}{}_{}{}^{4}})^{\frac{1}{21}}$$ (19) Thermalization of the bulk occurs at the smaller $`H`$ from (19) and (20), and for $`HH_{eq}`$ (the seed cannot be thermalized for $`H>H_{eq}`$ because scattering dominates over decay then) <sup>7</sup><sup>7</sup>7In reality catalyzed thermalization of the bulk occurs after a number of efficient scatterings off the thermalized seed. It would be safer then to have a factor of 10 on the right-hand side of (16) and (17). However, since $`T_RH_{th}^{}{}_{}{}^{\frac{1}{2}}`$, $`T_R`$ remains within the same order of magnitude even if $`H_{th}`$ changes by a factor of 10.: $$H_{th}=min[H_{eq},(\frac{\alpha ^{18}\mathrm{\Gamma }_{d}^{}{}_{}{}^{5}m_{0}^{}{}_{}{}^{4}n_{I}^{}{}_{}{}^{4}}{m_{I}^{}{}_{}{}^{6}H_{I}^{}{}_{}{}^{6}})^{\frac{1}{7}},(\frac{\alpha ^2m_{0}^{}{}_{}{}^{28}H_{I}^{}{}_{}{}^{8}}{\mathrm{\Gamma }_dm_{I}^{}{}_{}{}^{2}n_{I}^{}{}_{}{}^{4}})^{\frac{1}{21}}]$$ (20) This holds if $`m_0`$ is small enough such that $`m_{0}^{}{}_{}{}^{2}<\alpha \frac{n_{Max}}{E_{Max}}`$ before efficient inflaton decay, which, after using (4), gives $$\frac{\alpha ^5\mathrm{\Gamma }_{d}^{}{}_{}{}^{4}}{H_{I}^{}{}_{}{}^{6}m_{I}^{}{}_{}{}^{5}}n_{I}^{}{}_{}{}^{3}<m_{0}^{}{}_{}{}^{2}<\frac{\alpha \mathrm{\Gamma }_{d}^{}{}_{}{}^{2}}{H_{I}^{}{}_{}{}^{2}m_I}n_I$$ (21) Therefore if $`m_0`$ satisfies (21), $`H_{th}`$ is determined from (20) and $`T_R(H_{th}M_{Pl})^{\frac{1}{2}}`$. 3- Consider the case where $`m_0`$ exceeds the bound in (21). In this case $`m_{0}^{}{}_{}{}^{2}`$ dominates over $`\alpha \frac{n_{Max}}{E_{Max}}`$ before efficient inflaton decay (i.e., for $`H\mathrm{\Gamma }_d`$). In order to find $`H_{th}`$, the above steps are repeated but the values of $`E_{Max}`$, $`n_{Max}`$, $`E_s`$, and $`n_s`$ are replaced from (2) and (3), instead of (5) and (6). This leads to $$H_{eq}\frac{H_{I}^{}{}_{}{}^{2}m_Im_{0}^{}{}_{}{}^{2}}{\alpha \mathrm{\Gamma }_dn_I}$$ (22) The conditions for catalyzed thermalization of the bulk give $$H(\frac{\alpha ^9\mathrm{\Gamma }_{d}^{}{}_{}{}^{2}m_{0}^{}{}_{}{}^{2}n_{I}^{}{}_{}{}^{2}}{m_{I}^{}{}_{}{}^{3}H_{I}^{}{}_{}{}^{4}})^{\frac{1}{3}}$$ (23) instead of (18), and $$H(\frac{\alpha ^7H_{I}^{}{}_{}{}^{4}m_{0}^{}{}_{}{}^{14}}{m_I\mathrm{\Gamma }_{d}^{}{}_{}{}^{2}n_{I}^{}{}_{}{}^{4}})^{\frac{1}{9}}$$ (24) instead of (19). Therefore in this case $$H_{th}=min[(\frac{\alpha ^9\mathrm{\Gamma }_{d}^{}{}_{}{}^{2}m_{0}^{}{}_{}{}^{2}n_{I}^{}{}_{}{}^{2}}{m_{I}^{}{}_{}{}^{3}H_{I}^{}{}_{}{}^{4}})^{\frac{1}{3}},(\frac{\alpha ^7H_{I}^{}{}_{}{}^{4}m_{0}^{}{}_{}{}^{14}}{m_I\mathrm{\Gamma }_{d}^{}{}_{}{}^{2}n_{I}^{}{}_{}{}^{4}})^{\frac{1}{9}}]$$ (25) It is also necessary that $`H_{th}\mathrm{\Gamma }_d`$ for thermalization to occur before efficient decay of the inflaton (otherwise we are back to case 2 in the above) which implies that $`m_0`$ must be indeed large. In this case the plasma is thermal when inflaton decay is completed at $`H\mathrm{\Gamma }_d`$ and $`T_R(\mathrm{\Gamma }_dM_{Pl})^{\frac{1}{2}}`$. In summary, depending on the value of $`m_0`$ and inflationary model parameters $`n_I`$, $`m_I`$, and $`\mathrm{\Gamma }_d`$, we have (α4Γd3MPlHI4mI4nI2)12< TR< (ΓdMPl)12superscriptsuperscript𝛼4superscriptsubscriptΓ𝑑3subscript𝑀𝑃𝑙superscriptsubscript𝐻𝐼4superscriptsubscript𝑚𝐼4superscriptsubscript𝑛𝐼212< subscript𝑇𝑅< superscriptsubscriptΓ𝑑subscript𝑀𝑃𝑙12{({{\alpha}^{4}{{\Gamma}_{d}}^{3}{M}_{Pl}\over{{H}_{I}}^{4}{{m}_{I}}^{4}}}{{n}_{I}^{2})}^{1\over 2}{\ \lower-1.2pt\vbox{\hbox{\hbox to0.0pt{$<$\hss}\lower 5.0pt\vbox{\hbox{$\sim$}}}}\ }{T}_{R}{\ \lower-1.2pt\vbox{\hbox{\hbox to0.0pt{$<$\hss}\lower 5.0pt\vbox{\hbox{$\sim$}}}}\ }{({\Gamma}_{d}{M}_{Pl})}^{1\over 2} (26) Big-bang nucleosynthesis yields the bound TR< 1081010subscript𝑇𝑅< superscript108superscript1010{T}_{R}{\ \lower-1.2pt\vbox{\hbox{\hbox to0.0pt{$<$\hss}\lower 5.0pt\vbox{\hbox{$\sim$}}}}\ }{10}^{8}-{10}^{10} GeV, which constrains the parameters of the model. This ensures that gravitinos are not produced in dangerous abundances in the thermal bulk. Even before thermalization of the bulk gravitinos are produced in the non-thermal bulk and the thermal seed. One may wonder whether the requirement TR< 1081010subscript𝑇𝑅< superscript108superscript1010{T}_{R}{\ \lower-1.2pt\vbox{\hbox{\hbox to0.0pt{$<$\hss}\lower 5.0pt\vbox{\hbox{$\sim$}}}}\ }{10}^{8}-{10}^{10} GeV also keeps the abundance of these gravitinos at a safe level. The number density of gravitinos which are produced in the plasma is in general proportional to $`n^2`$ ($`n`$ is the typical number density of particles in the plasma) <sup>8</sup><sup>8</sup>8One factor of $`n`$ comes from the scattering rate and the other one is from the number density of particles in the plasma., which is much smaller before thermalization of the bulk. On the other hand, thermalization of the bulk releases a huge entropy and dilutes everything, including gravitinos which were produced in the thermal seed. In the case that thermalization occurs before efficient inflaton decay (case 3 in the above) it has been shown that the bound TR< 1081010subscript𝑇𝑅< superscript108superscript1010{T}_{R}{\ \lower-1.2pt\vbox{\hbox{\hbox to0.0pt{$<$\hss}\lower 5.0pt\vbox{\hbox{$\sim$}}}}\ }{10}^{8}-{10}^{10} GeV guarantees successful nucleosynthesis, despite the fact that the instantaneous temperature of the plasma is much higher at earlier times . We can expect that the same conclusion also holds for cases 1 and 2 above and therefore the strongest constraint on the model parameters is, in general, due to gravitino production in the thermal bulk. Two clarifying comments are to be made here. We have actually studied the thermalization of particles with mass $`m_0`$. After their thermalization, the lighter particles reach thermal equilibrium through gauge-strength scatterings of these particles (e.g., two squarks exchange a gaugino and scatter to two quarks). Therefore only the typical mass scale of the sector is important for its thermalization and not the individual particle masses. Finally, the above analysis was done for relativistic particles and its consistency requires that $`T_Rm_0`$. If $`T_R<m_0`$, particles with mass $`m_0`$ become non-relativistic before thermalization occurs. These particles will then decay to other (relativistic) particles whose subsequent (cascade) decay leads to thermalization and determines the reheat temperature. The same analysis as above can be done in this case, but $`m_I`$ and $`m_0`$ should be replaced by $`m_0`$ and the mass scale of secondary decay products respectively. ## 5 The MSSM Example and the Role of Flat Directions Here we consider a model which consists of the Minimal Supersymmetric Standard Model (MSSM) sector, the inflaton sector, and the low-energy supersymmetry breaking sector, where the latter two interact with the MSSM sector through gravitationally suppressed couplings. To keep the matter content minimal, no right-handed neutrinos or gauge sectors at the intermediate scale are introduced, and therefore the MSSM sector is the only one with gauge strength couplings. We consider a chaotic inflation model with the potential $`\frac{1}{2}m_{I}^{}{}_{}{}^{2}\varphi ^2`$ for the inflaton $`\varphi `$, where the inflaton decays to other particles via gravitationally suppressed couplings (and hence there is no stage of parametric resonance decay in this model). From the COBE data on the isotropy of the microwave background radiation, $`m_I10^{13}`$ GeV . At the end of inflation $`\varphi 10^{19}`$ GeV which yields $`H_I{\displaystyle \frac{m_I\varphi }{M_{Pl}}}10^{13}\mathrm{GeV}`$ $`n_I=m_I\varphi ^210^{51}\mathrm{GeV}^3`$ $`\mathrm{\Gamma }_d{\displaystyle \frac{m_{I}^{}{}_{}{}^{3}}{M_{Pl}^{}{}_{}{}^{2}}}10\mathrm{GeV}`$ Also for the MSSM sector, $`m_010^2`$ GeV. It is easily seen that with these values, (13) is satisfied and the reheat temperature derived from (12) is $`T_R10^6`$ GeV, which is far below the limit set by the nucleosynthesis bound. In the early universe however, it is possible to have much larger mass scales in the MSSM sector. Flat directions in the scalar potential of the MSSM (denoted as $`\phi `$ here) can acquire very large vevs during inflation, both in the minimal models and in the no-scale models of supergravity. At the end of inflation $`\phi (H_IM^{n3})^{\frac{1}{n2}}`$ , where $`n`$ (not to be confused with the number density) is the order of nonrenormalizable superpotential term which lifts the flat direction, and $`M`$ is the scale of new physics which induces the nonrenormalizable terms. The non-zero energy density of the early universe strongly breaks supersymmetry and induces a negative mass-squared of order $`H^2`$ for the flat direction. So long as other supersymmetry breaking sources do not induce a positive mass-squared of the same order, $`\phi (HM^{n3})^{\frac{1}{n2}}`$ and the flat direction vev decreases slowly. When other sources become dominant, the flat direction mass-squared becomes positive, oscillations start, and from then on $`\phi `$ is Hubble redshifted more rapidly. A non-zero vev for the flat direction breaks a subgroup of the MSSM gauge group. Those MSSM fields which have $`F`$\- and $`D`$-term couplings to the flat direction acquire masses of order $`g\phi `$, where $`g`$ is a gauge or Yukawa coupling. This implies that in the early universe the mass scale of those fermions and scalars which are coupled to the flat direction, as well as the gauge fields and gauginos of the broken subgroup, can be much larger than $`10^2`$ GeV. Thermalization in the MSSM sector can then happen earlier (case 2 or 3 in the above), leading to a reheat temperature higher than $`10^6`$ GeV. For example, if $`m_010^{10}`$ GeV (which can be easily induced by the flat direction vev), thermalization occurs before efficient inflaton decay and $`T_R10^{10}`$ GeV. On the other hand, the flat direction vevs, and thus the induced masses, are rapidly redshifted after the flat direction oscillations start. The question is whether they remain large enough for a sufficiently long time so that thermal equilibrium is achieved. This depends on the time when oscillations start and the flat direction vev at the onset of oscillations. In the standard approach , oscillations start at $`Hm_{\frac{3}{2}}10^210^3`$ GeV when the low-energy supersymmetry breaking takes over the Hubble-induced one. Recently it has been shown that many flat directions can start their oscillations much earlier, due to plasma effects . Therefore the mass scale of the model at early times, and hence the thermalization dynamics, has a dependence on the nature of the flat direction and its initial vev. This may suggest another role for the supersymmetric flat directions besides baryogenesis and dark matter candidates : they may also assist thermalization to happen earlier and lead to a higher reheat temperature. In conclusion the reheat temperature $`T_R`$ may be below the upper limit $`(\mathrm{\Gamma }_dM_{Pl})^{\frac{1}{2}}`$ by several orders of magnitude, in which case the constraints on the parameters of the model are considerably relaxed <sup>9</sup><sup>9</sup>9For example, in models of $`D`$-term inflation or supersymmetric models for new inflation the inflaton mass can be be as large as $`10^{15}10^{16}`$ GeV . In such models the upper limit on $`T_R`$ can be as high as $`10^{12}`$ GeV. If $`m_0`$ is small enough, the reheat temperature will be far below its upper limit and hence in agreement with the nucleosynthesis bound.. However, flat directions with large vevs may result in earlier thermalization and push $`T_R`$ towards its upper limit. ## 6 Conclusion We have considered thermalization after perturbative decay of the inflaton. In this regime the number density of inflaton decay products is smaller than the thermal equilibrium value, while the mean energy of particles is larger than its thermal value. Particles with the highest energy $`E_{Max}`$ dominate the number density and the energy density of the plasma of inflaton decay products, and $`22`$ scatterings off these particles increase the energy of other particles. We compared the rate for $`22`$ scatterings, particle decay, and $`2N`$ scatterings and found that the first two are the important ones. Thermalization occurs when decays dominate scatterings and, depending on the model parameters, different situations may arise. We showed that the thermalization time and the reheat temperature has a dependence on the typical mass scale of inflaton decay products. If this mass scale is sufficiently large, a seed in the plasma is self-thermalized and the scattering of particles in the plasma off the thermalized seed leads to catalyzed thermalization of the bulk. Otherwise, the bulk is self-thermalized at late times, when all particles in the plasma have energies of order $`E_{Max}`$. As a result, the reheat temperature can vary in a wide range. The strongest constraint (from nucleosynthesis) on the model parameters is derived for a very large mass scale when thermalization occurs before efficient inflaton decay. In general, a small mass scale considerably relaxes this constraint. We also considered the MSSM case as an example. It was shown that the mass scale of the model can be much larger in the early universe due to particle couplings to the flat directions with very large vevs. If the flat direction oscillations do not start early, these vevs stay large enough for a sufficiently long time. In this case, thermalization may occur earlier, thus leading to a higher reheat temperature and resulting in a stronger constraint on the model parameters. Finally, the dynamics of thermalization seems to be very detailed even in the perturbative regime of inflaton decay. Acknowledgements The author wishes to thank B. A. Campbell for useful discussions and comments. He also thanks A. N. Kamal for discussions and K. Kaminsky for careful reading of the manuscript. This work was supported in part by the Natural Sciences and Engineering Research Council of Canada.
warning/0004/astro-ph0004264.html
ar5iv
text
# Unavoidable Selection Effects in the Analysis of Faint Galaxies in the Hubble Deep Field: Probing the Cosmology and Merger History of Galaxies ## 1. Introduction Number counting of faint galaxies is one of the most fundamental observational tests with which the formation/evolution of galaxies as well as the geometry of the universe is probed. The best view to date of the optical sky to faint flux levels is given by the Hubble Deep Field (HDF, Williams et al. 1996), and it provides a valuable information to a wide range of studies on galaxies and cosmology. A comprehensive study of the HDF galaxy counts has been performed by Pozzetti et al. (1998), and they found that a simple model of pure luminosity evolution (PLE), in which galaxies evolve passively due to star formation histories without mergers or number evolution, gives a reasonable fit to the HDF counts in all the four passbands of $`U`$, $`B`$, $`V`$, and $`I`$, when an open universe with $`\mathrm{\Omega }_0=0.1`$ is assumed. The increase of the number of galaxies with their apparent magnitude was originally proposed as a measure of the geometry of the universe (Sandage 1961), and considerable efforts have been made along this line (e.g., Yoshii & Takahara 1988; Fukugita, Takahara, Yamashita & Yoshii 1990; Yoshii & Peterson 1991). However, the obtained constraints on cosmological parameters based on the PLE model have not been thought deterministic because of possible number evolution of galaxies by mergers. Particularly, when the PLE model is used, the Einstein-de Sitter (EdS) universe ($`\mathrm{\Omega }_0=1`$) underpredicts the observed galaxy counts at faint magnitudes, but a simple model of galaxy number evolution can reproduce the observed counts and save the EdS universe (Rocca-Volmerange & Guiderdoni 1990; Pozzetti et al. 1996). This degeneracy between the effects of galaxy evolution and cosmology has been a major problem when one uses the galaxy number count to determine the geometry of the universe. The information of redshifts is able to break such a degeneracy, because luminous galaxies at great distance are distinguishable from dwarf galaxies in a local universe. Although most of the HDF galaxies are too faint to measure the spectroscopic redshifts, several catalogs of their photometric redshifts have been published (Sawicki, Lin, & Yee 1997; Wang, Bahcall, & Turner 1998; Fernández-Soto, Lanzetta, & Yahil 1999). The follow-up studies based on these catalogs show that the photometric redshifts give reasonably reliable estimates of spectroscopic redshifts and are useful for a statistical study of the HDF galaxies. Here we give a combined analysis for the HDF counts and redshifts and constrain the cosmological parameters separately from the merger history of galaxies. Both the number count of faint galaxies and their redshift distribution are significantly affected by the selection effects inherent in the method of detecting galaxies in faint surveys, but these important effects have been ignored in almost all previous studies except for Yoshii & Fukugita (1991) and Yoshii (1993). It is well known that the surface brightness of galaxies rapidly becomes dimmer with increasing redshift as $`\left(1+z\right)^4`$ (Tolman 1934), and this cosmological dimming makes many high-redshift galaxies remain undetected below the threshold value of surface brightness adopted in a galaxy survey (Pritchet & Kline 1981; Tyson 1984; Ellis, Sievers, & Perry 1984). The seeing or smoothing of an image furthermore lowers its surface brightness, and the photometry scheme used in a survey heavily affects a magnitude estimate of the faintest galaxies. Some observers apply corrections to raw counts of faint galaxies for those undetected, but it is in principle difficult and heavily model-dependent to estimate the number of undetected galaxies. Rather, the best way is to make theoretical predictions with the selection effects taken into account and then compare them directly to raw counts (Yoshii 1993). This paper is the first analysis of the HDF galaxies in which the above selection bias against high-redshift galaxies is explicitly incorporated. We use a standard PLE model of galaxies including the effects of internal dust obscuration and intergalactic HI absorption. Number evolution of galaxies is also allowed for with simple modifications to the PLE model. Throughout this paper, we use the AB photometry system with the notation of $`U_{300}`$, $`B_{450}`$, $`V_{606}`$, and $`I_{814}`$ (Williams et al. 1996). In §2, we present a detailed description for models of galaxy evolution and formulations to calculate galaxy counts and redshift distribution with the selection effects taken into account. Extensive calculations of number count predictions and comparison to the HDF counts are given in §3, checking in great detail the uncertainties arising from the prescribed properties of local galaxies and their evolution. We will give the comparison of the model predictions with the observed photometric redshift distribution in §4. We discuss the results in §5. The summary and conclusion of this paper are given in §6. ## 2. The Model of Galaxies and Detection in the HDF First we describe the basic ingredients involved in our theoretical modeling such as the local luminosity function, galaxy evolution in luminosity and number, internal and intergalactic absorption, and the selection effects. Then we will present the formulations to calculate the number count of faint galaxies and their redshift distribution. ### 2.1. Galaxies at Present, and Their Evolution We use a standard PLE model of galaxy evolution, in which galaxies are classified into five morphological types of E/S0, Sab, Sbc, Scd, and Sdm. Spectral energy distributions (SEDs) and their evolution are calculated by using the galaxy evolution model of Arimoto & Yoshii (1987) for elliptical galaxies and the I1 model of Arimoto, Yoshii, & Takahara (1992) for spiral galaxies. These models are constructed to reproduce the photometric and chemical properties of present-day galaxies. In order to see the systematic uncertainty in evolution models, we will also use an updated version of these models by Kobayashi et al. (1999) using the latest database of stellar populations compiled by Kodama & Arimoto (1997). We set the epoch of galaxy formation at $`z_F=5`$ as a standard and change this value to see the systematic uncertainty. The luminosity function of local galaxies is also important in predicting the number count and redshift distribution of faint galaxies. We use the type-dependent (E/S0, Spiral, and Irr) $`B`$-band luminosity function derived from the Second Southern Sky Redshift Survey (SSRS2, Marzke et al. 1998). We associate the Sab, Sbc and Scd models to be assigned to spiral galaxies, whereas the Sdm model assigned to irregular galaxies. The relative proportions of Sab, Sbc, and Scd are taken from Pence (1976). In order to check the systematic uncertainty related to the luminosity function, we also use the type-independent luminosity function of Stromlo-APM redshift survey (Loveday et al. 1992) and the type-dependent luminosity function from the Center for Astrophysics (CfA) redshift survey (Huchra et al. 1983). Their Schechter parameters are tabulated in Table 1 (see also Efstathiou, Ellis, & Peterson 1988 and Yoshii & Takahara 1989). ### 2.2. Absorption The above models of galaxy evolution do not include the absorption by interstellar dust which becomes significant for high-$`z`$ galaxies when observed in optical bands. In order to take this effect into account, we make a physically natural assumption that the dust optical depth is proportional to the column density and metallicity of the gas. In fact, it is well known that the Galactic extinction is well correlated to the column density of the HI gas (e.g., Burstein & Heiles 1982). It is also known that the dust opacity becomes smaller in order of decreasing metallicity from the Galaxy to the Large and then Small Magellanic Clouds, when the gas column density is fixed (e.g., Pei 1992). Since the galaxy evolution models give the gas fraction $`f_g`$ and the metallicity $`Z_g`$ in the gas, the dust optical depth is calculated from $`\tau _{\mathrm{dust}}=\kappa f_gZ_gr_e^2\left(M/L_B\right)L_B`$, where $`r_e`$, $`M`$, and $`L_B`$ are the effective radius, the baryon mass, and the $`B`$-band luminosity of a galaxy, respectively. (We will describe the treatment of galaxy size in §2.3.) The proportionality constant $`\kappa `$ is chosen to be consistent with the present-day, average extinction of $`A_V0.17`$ taken from the Galactic extinction map (Burstein & Heiles 1982; Schlegel, Finkbeiner, & Davis 1998) and a theoretical estimate (Hatano, Branch, & Deaton 1998). The standard extinction curve of our Galaxy (e.g., Pei 1992) is used for the wavelength dependence of the optical depth. Given the optical depth, the attenuation of emerging stellar lights depends on the spatial dust distribution. Following Disney, Davies, & Phillipps (1989), there are two extreme cases such as the screen model in which the dust is distributed on the line of sight to stars, and the slab model in which the dust has the same distribution with stars. Neglecting the scattering of lights by dust, the attenuation factor of stellar lights is given by $`\mathrm{exp}\left(\tau _{\mathrm{dust}}\right)`$ for the screen model and $`\left[1\mathrm{exp}\left(\tau _{\mathrm{dust}}\right)\right]/\tau _{\mathrm{dust}}`$ for the slab model. In fact, the galaxy evolution model used here has been made to reproduce the present-day SED of galaxies which has been already affected by dust obscuration. We take into account this point and hence correct the above attenuation factors by using the optical depth at present. Therefore the present-day SEDs of model galaxies are the same for all the prescriptions of dust-free, screen, and slab models. In the slab model the apparent reddening reaches an asymptote when the optical depth becomes much larger than unity, because the observed lights are emitted from surface regions of a galaxy where the optical depth to an observer is low. However, the observed correlation between the power-law index of UV spectra and the Balmer line ratio, both of which are a reddening indicator, extends well beyond the asymptote. This indicates that the observed reddening of starburst galaxies is larger than expected from the slab model, and at least some fraction of dust should behave like a screen (for detail see Calzetti, Kinney, & Storchi-Bergmann 1994). We then use the screen model as a standard, considering that UV and optical observations of starburst galaxies favor the screen dust. It may be an extreme prescription that all dust is distributed as a screen, but note that the emergent lights are rapidly attenuated exponentially once a considerable fraction of dust contributes to the screen. Therefore the screen model is more appropriate than the slab model in which the attenuation factor decreases only moderately like $`\tau _{\mathrm{dust}}^1`$ when $`\tau _{\mathrm{dust}}\mathrm{}`$. The spectral energy distributions (SEDs) of galaxies of various types are given in Fig. 1 at several epochs of galaxy evolution for three cases of dust extinction such as the screen model (solid line), the slab model (dashed line), and the no-extinction model (dotted line). The effect of dust extinction is especially important for elliptical galaxies at high redshifts, where UV radiation is quite strong because of the initial starbursts supposed in the galactic wind model of elliptical galaxies (Arimoto & Yoshii 1987). According to the method described above, the dust optical depth during the initial starburst phase of elliptical galaxies is estimated to be much larger than unity ($`\tau \stackrel{>}{}10`$) for UV photons at $``$ 2000 Å in the restframe. On the other hand, the evolution of dust obscuration makes UV luminosity of late-type spiral galaxies brighter at early epochs than that without the dust effect, because of the lower metal abundance than the present-day galaxies. However, this effect in late-type spiral galaxies is not significant because they are not heavily obscured at present. In addition to the internal absorption by dust, the intergalactic absorption significantly affects the visibility of high-redshift galaxies (Yoshii & Peterson 1994; Madau 1995). Lights from a distant galaxy at rest wavelengths below the Lyman limit (912 Å) and those below the Lyman $`\alpha `$ line (1216 Å) are extinguished by Lyman continuum absorption and Lyman series line absorption, respectively, in intergalactic HI clouds along the line of sight. We include this effect consistently in our theoretical calculations making use of the intergalactic optical depth calculated by Yoshii & Peterson (1994). The optical depth of this absorption is shown in Fig. 2, as a function of observed wavelength for various source redshifts. ### 2.3. Selection Effects Apparent surface brightness and size of an image in a survey observation are the essential quantities for it to be detected as a real galaxy. We calculate these quantities of a model galaxy assuming its intrinsic luminosity profile and size and taking into account the cosmological dimming and the observational seeing. For the details of the formulations, see Yoshii (1993). #### 2.3.1 Galaxy Sizes and Luminosity Profile In our PLE model, we assume that the galaxy size does not evolve except for the case of mergers of galaxies, and use the empirical relation between the effective radius $`r_e`$ and absolute luminosity $`L_B`$ for local galaxies. If we allow for the number evolution of galaxies, we must take into account the change of galaxy sizes, and we will discuss this in §2.4. Fig. 3 shows the size-luminosity relation of local elliptical and spiral galaxies. The data of elliptical galaxies are taken from Bender et al. (1992), and those of spiral galaxies from Impey et al. (1996). Although there is a significant scatter in this relation, we use a simple power-law relation for this relation as $$r_eL_B^{2.5/p},$$ (1) or, if expressed in terms of the absolute $`B`$ magnitude, $$M_B=p\mathrm{log}r_e+q+\left(p5\right)\mathrm{log}\left(H_0/50\mathrm{km}/\mathrm{s}/\mathrm{Mpc}\right).$$ (2) Elliptical galaxies form two distinct families which follow the well-separated sequences at low luminosities in the $`r_e`$-$`L_B`$ diagram. One is the ordinary sequence from giant through dwarf elliptical (GDE) galaxies, while the other is the bright sequence from giant through compact elliptical (GCE) galaxies (see Fig 3). Since the predictions of galaxy number count are not sensitive to whichever sequence is used in the analysis (see §3.2 and Fig 9), we take the GDE sequence as the standard size-luminosity relation of elliptical galaxies in this paper. The $`r_e`$-$`L_B`$ relations fitted to the data in Fig.3 yield $`(p,q)=(6.0,16)`$ and (3.5,18.7) for the GDE and GCE sequences of elliptical galaxies, respectively, and (9.4, 12) for spiral galaxies. In order to examine the uncertainty due to the significant scatter in the $`r_e`$-$`L_B`$ relation, we derive the standard deviation in $`\mathrm{\Delta }\left(\mathrm{log}r_e\right)`$ from the best-fit relation and repeat calculations with the shifted relations shown in Fig. 3 by the dashed lines. The radial distribution of surface brightness is assumed to follow de Vaucouleurs’ (1962) profile ($`S\mathrm{exp}[(r/r_e)^{1/4}`$\] ) for elliptical galaxies and an exponential profile for spiral galaxies (Freeman 1970). Then we can calculate the radial distribution of surface brightness of a galaxy at given redshift in any passband, when the galaxy type, the present-day $`B`$ luminosity, and the evolution model are specified. This surface brightness profile should be convolved with a Gaussian point-spread function (PSF) having the same dispersion with the observational seeing. #### 2.3.2 Object Detection Let $`S_{\mathrm{th}}`$ be the surface brightness threshold adopted in a galaxy survey. Strictly, this threshold could change due to different noise levels within a survey field, but we use a single value for the simplicity. When the observed profile of a galaxy image is calculated as above, we can estimate the isophotal size which encircles a region of a galaxy image with surface brightness brighter than $`S_{\mathrm{th}}`$. If the isophotal size of a galaxy is zero, i.e., the central surface brightness is fainter than $`S_{\mathrm{th}}`$, this galaxy can not be detected in the survey. Usually a minimum isophotal diameter $`D_{\mathrm{min}}`$, which is comparable to the seeing size, is adopted as a condition for an image to be detected as a galaxy. We can calculate the isophotal diameter for a model galaxy, then it is easy to check whether this galaxy meets the detection criterion in the galaxy survey. #### 2.3.3 Photometry Scheme There are three photometry schemes to evaluate the apparent magnitude of galaxies such as isophotal, aperture, and pseudo-total magnitudes. These magnitudes may significantly differ especially near the detection limit, and this difference should be included in the theoretical modeling of the selection effects inherent in the method of detecting faint galaxies. The isophotal magnitude is the flux within the isophotal size of a galaxy image. The aperture magnitude is the flux within a fixed aperture which is adopted by observers. Some observers often make corrections to these magnitudes into pseudo-total magnitudes, which are intended to mimic an ideal total flux of a galaxy. However, such corrections use a model luminosity profile to evaluate the flux from an ‘undetected’ part of a galaxy image. It is inappropriate to make a model-dependent correction to the observed quantities with which various theoretical models are compared. The best way is, on the contrary, to incorporate all the necessary corrections in the theoretical models to be compared directly with the observed quantities. We therefore suggest that observers should also present raw counts and magnitudes, in addition to presenting corrected quantities in their papers. #### 2.3.4 Detection and Photometry of the HDF Galaxies Here we describe the detection processes of the HDF galaxies, following Williams et al. (1996), and see also Bouwens, Broadhurst, & Silk (1998). Object detection is performed in the combined $`V_{606}+I_{814}`$ image. It is first convolved with a fixed smoothing kernel of 25 pixels (=0.04 square arcseconds), then pixels having values higher than a fixed threshold above a local sky background are marked as potentially being part of an object. After thresholding, regions consisting of more than contiguous 25 pixels are counted as sources. It corresponds to an isophotal diameter limit of $`D_{\mathrm{min}}`$ 0.2 arcsec, about 1.6 times larger than the FWHM of the PSF. The surface brightness threshold is not clearly indicated in Williams et al. (1996), but we can evaluate this value from the surface brightness distribution of the HDF galaxies. In Fig. 4, we plot their apparent magnitudes versus average surface brightness in the four passbands of $`U_{300}`$, $`B_{450}`$, $`V_{606}`$, and $`I_{814}`$ by using the published sizes and magnitudes of the HDF catalog. Here the size refers to the isophotal size of the combined $`V_{606}+I_{814}`$ image, and the magnitude refers to the isophotal magnitude measured within the isophotal size. First we consider the $`V_{606}`$ and $`I_{814}`$ bands. Figure 4 shows that no galaxies are detected when the average surface brightness is fainter than $`S_{\mathrm{th}}=27.5`$ mag arcsec<sup>-2</sup> in $`V_{606}`$ and $`27.0`$ mag arcsec<sup>-2</sup> in $`I_{814}`$. The faintest surface brightness detected in the HDF for galaxies with a fixed isophotal magnitude becomes brighter for brighter galaxies. This trend occurs because the size of galaxies is larger for brighter galaxies. The edge of an isophotal image corresponds to the isophotal limit, and hence its average surface brightness within the isophotal area is always brighter than its threshold when the galaxy image have a bright central part and therefore a larger size. Consequently the faintest surface brightness at the faintest isophotal magnitudes, i.e., $`S_{\mathrm{th}}=27.5`$ mag arcsec<sup>-2</sup> in $`V_{606}`$ and $`27.0`$ mag arcsec<sup>-2</sup> in $`I_{814}`$, gives the surface brightness threshold for object detection. In our calculations the above threshold values in the $`V_{606}`$ and $`I_{814}`$ bands are used respectively, although in reality the detection was done by the combined $`V_{606}+I_{814}`$ image in the HDF catalog. On the other hand, the surface brightness threshold is not clear in the $`U_{300}`$ and $`B_{450}`$ bands, because the object detection was done without these bands. It should also be noted that the isophotal $`U_{300}`$ and $`B_{450}`$ magnitudes are defined as the flux within the isophotal size in the combined $`V_{606}+I_{814}`$ image. Among the objects to which these isophotal magnitudes are assigned, those with $`S/N>2`$ are detected as galaxy images in the $`U_{300}`$ and $`B_{450}`$ bands (Williams et al. 1996). We see a clear boundary indicated by dot-dashed line running from upper-left to lower-right in the $`U_{300}`$ and $`B_{450}`$ panels of Fig. 4, and this corresponds to the line of $`S/N=2`$. It is easy to show that this is equivalent to a condition of $`S+m`$ = const if the noise level is proportional to $`A^{1/2}`$ (Poisson type noise), where $`A`$ is the isophotal area. We have also confirmed that the galaxies in $`U_{300}`$ and $`B_{450}`$ with $`S/N2`$ are actually on the dot-dashed line of Fig. 4 <sup>1</sup><sup>1</sup>1This is true for the three WF fields, but not for the PC field, because of the different surface brightness threshold employed (Williams et al. 1996). The number of galaxies in the PC field is negligible compared with the WF fields.. In our calculations the galaxies in the $`U_{300}`$ and $`B_{450}`$ bands are detected if they are detected in $`I_{814}`$ band and furthermore meet the criterion of $`S/N>2`$ within the isophotal area in the $`I_{814}`$ band (i.e., those below the dot-dashed line in Fig. 4). ### 2.4. Merger and Number Evolution Currently the most popular theory for the structure formation in the universe is the bottom-up scenario with the cold dark matter which dominates the total mass density of the universe, in which smaller mass objects form earlier and then merge into larger objects (e.g., Blumenthal et al. 1984). Merging history of dark matter haloes is relatively well studied by analytical methods as well as $`N`$-body simulations, but merging history of galaxies could significantly differ from that of dark matter haloes and is poorly known. Since a number evolution of galaxies caused by galaxy mergers significantly affects the number count of faint galaxies, we investigate this effect by using a simple merging model in which the luminosity density of galaxies is conserved. A common practice for this is to introduce the redshift-dependent parameters of Schechter-type luminosity function of galaxies such as $`\varphi ^{}\left(z\right)`$ $`=`$ $`\varphi ^{}\left(0\right)\left(1+z\right)^\eta `$ $`L^{}\left(z\right)`$ $`=`$ $`L^{}\left(0\right)\left(1+z\right)^\eta .`$ (3) We adopt a single value of $`\eta `$ for all types of galaxies for simplicity. In the analysis of this paper, the size of a galaxy is crucially important for evaluating the selection effects. Merger of galaxies should change their size, and this should also be taken into account. The empirical relation $`r_eL_B^{2.5/p}`$ may not hold at high redshifts, depending on how $`r_e`$ and $`L_B`$ change during the merger process. We assume that the change of $`L_B`$ and $`r_e`$ during merger processes always satisfy a relation $`L_Br_e^\xi `$. (Note that this relation is physically different from equation 1 which is the relation of galaxies at a fixed time, but describing the change of luminosity and size of a test galaxy during merger processes.) If $`\xi =2`$, the surface brightness of galaxies is conserved during mergers, and if $`\xi =3`$, the luminosity density in each galaxy is conserved. Then the luminosity and size of a $`z=0`$ galaxy evolve as $`L_B`$ $``$ $`L_B\left(1+z\right)^\eta `$ (4) $`r_e`$ $``$ $`r_e\left(1+z\right)^{\eta /\xi }.`$ (5) By applying this transformation to the empirical $`r_e`$-$`L_B`$ relation at $`z=0`$, it is straightforward to give the $`r_e`$-$`L_B`$ relation as a function of redshift: $$r_e(L_B,z)=r_e(L_B,0)\times \left(1+z\right)^{\frac{\eta }{p}\left(\frac{p}{\xi }2.5\right)},$$ (6) which is a generalization of eq. 1. The value of $`\xi `$ depends on the physical process of mergers. Generally, merger products are expected to become more compact than pre-merger progenitors in gas-rich mergers because of efficient cooling and dissipation. On the other hand, merger products become less compact in gas-less mergers because the relative translational energy of pre-merger stellar progenitors is converted into the internal kinetic energy of a merged stellar system. The former corresponds to larger $`\xi `$, while the latter to smaller $`\xi `$. We use $`\xi =3`$ as a standard value and examine the effect of changing this value. The assumed conservation of the total luminosity density of all galaxies (i.e., $`\varphi ^{}\left(z\right)L^{}\left(z\right)`$ = constant) will also be discussed when we draw our conclusion of this paper. ### 2.5. Formulations In the following we describe the formulations necessary to calculate the number count and redshift distribution of faint galaxies. We denote $`\lambda `$ as the observed wavelength, and $`\lambda _D`$ as the wavelength at which the object detection is performed. For example, we use $`\lambda =\lambda _D=V_{606}`$ or $`I_{814}`$ for counts in the $`V_{606}`$ and $`I_{814}`$ bands, whereas $`\lambda _D=I_{814}`$ and $`\lambda =U_{300}`$ or $`B_{450}`$ for the $`U_{300}`$ and $`B_{450}`$ bands (see §2.3.4). The number of galaxies per unit steradian, per unit apparent magnitude ($`m_\lambda `$), and per unit redshift is written as $$\frac{d^3N}{dm_\lambda dzd\mathrm{\Omega }}=H\left(x\right)\frac{d^2V}{dzd\mathrm{\Omega }}\underset{i}{}\varphi _i(M_B,z)\frac{dM_B}{dm_\lambda },$$ (7) where $`x=D_{\lambda _D}D_{\mathrm{min}}`$, $`D_{\lambda _D}`$ is the isophotal diameter of galaxies measured in the $`\lambda _D`$ band, $`d^2V/dzd\mathrm{\Omega }`$ is the comoving volume element which depends on the cosmological parameters, $`\varphi _i`$ is the luminosity function per unit $`M_B`$ with the Schechter parameters given in equation (3), and $`H`$ is the step function \[$`H\left(x\right)=1`$ and 0 for $`x0`$ and $`<0`$, respectively\]. The absolute $`B`$ magnitude $`M_B`$ is that of the present-day galaxies, which is related with $`z`$ and $`m_\lambda `$ by $`K`$-correction and evolutionary $`\left(E\right)`$ correction (see equation 12 below). The subscript $`i`$ denotes the galaxy type. The quantity $`d^3N/dm_\lambda dzd\mathrm{\Omega }`$ gives the redshift distribution of galaxies, and the integration over $`z`$ gives the number count or the number-magnitude relation. We calculate $`D_{\lambda _D}`$ and $`M_B`$ as a function of $`m_\lambda `$ and $`z`$, taking into account the selection effects and the photometry scheme. This is carried out as follows: For a given set of $`M_B`$ and $`z`$, by using the assumed luminosity profile and the $`L_B`$-$`r_e`$ relation, we first calculate the surface brightness distribution of a galaxy image in the $`\lambda _D`$ band. Comparing this surface brightness distribution with the adopted threshold $`S_{\mathrm{th}}`$ in the $`\lambda _D`$ band, we then calculate the isophotal diameter $`D_{\lambda _D}`$. Given this diameter, we calculate the isophotal magnitude in the $`\lambda `$ band. (If the aperture magnitude is used in a survey, we should use the fixed aperture here. The pseudo-total magnitude is easily calculated simply from the total absolute magnitude without using the surface brightness distribution.) In this way we finally obtain $`m_\lambda `$ as a function of $`M_B`$ and $`z`$, or conversely $`M_B`$ can be related to $`m_\lambda `$. It is obvious that $`D_{\lambda _D}`$ is automatically obtained in this process. In the following we give detailed numerical formulations necessary for the above calculations. Let $`g\left(\beta \right)`$ be radially symmetric luminosity profile (surface brightness distribution) of a galaxy, where $`\beta =r/r_e`$ is the radius from the center normalized by the effective radius of the galaxy. The adopted form of the profile is given by $$g\left(\beta \right)=\mathrm{exp}\left(a_n\beta ^{1/n}\right),$$ (8) and the integrated profile out to $`\beta `$ is given by $`G\left(\beta \right)2\pi _0^\beta g\left(\beta ^{}\right)\beta ^{}𝑑\beta ^{}.`$ We assume de Vaucouleurs’ profile ($`n=4`$) for elliptical galaxies and an exponential profile ($`n=1`$) for spiral galaxies, as mentioned earlier. An effective radius $`r_e`$ is defined as the radius within which a half of total luminosity is encircled, and by this definition the coefficient $`a_n`$ is given by $`a_4=7.67`$ and $`a_1=1.68`$. In order to incorporate the effect of observational seeing, we convolve this profile function with a Gaussian PSF with dispersion $`\sigma _t`$: $`\stackrel{~}{g}\left(\beta \right)`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{d\xi \xi }{\sigma _t^2}}g\left(\xi \right)\left\{I_0\left({\displaystyle \frac{\beta \xi }{\sigma _t^2}}\right)\mathrm{exp}\left({\displaystyle \frac{\beta \xi }{\sigma _t^2}}\right)\right\}`$ (9) $`\times `$ $`\mathrm{exp}\left({\displaystyle \frac{\left(\beta \xi \right)^2}{2\sigma _t^2}}\right),`$ where $`I_0\left(x\right)`$ is the modified Bessel function of the first kind (e.g., Press et al. 1992). The seeing FWHM in units of radian is now related to $`\sigma _t`$ as $`\sigma _t=\left({\displaystyle \frac{\mathrm{seeing}\mathrm{FWHM}}{2.35}}\right){\displaystyle \frac{d_A\left(z\right)}{r_e}},`$ (10) where $`d_A\left(z\right)`$ is the standard angular diameter distance. The surface brightness distribution of a galaxy image is given by $`S_\lambda \left(\theta \right)\left[\mathrm{mag}\mathrm{arcsec}^2\right]`$ $`=`$ $`M_B+\left(\lambda B\right)_0+K_\lambda \left(z\right)+E_\lambda \left(z\right)`$ (11) $`+`$ $`2.5\tau _{\mathrm{HI}}(\lambda ,z)\mathrm{log}e`$ $`+`$ $`5\mathrm{log}\left[r_e(M_B,z)\left(1+z\right)^2/10\mathrm{p}\mathrm{c}\right]`$ $`+`$ $`26.5721`$ $``$ $`2.5\mathrm{log}\left[\stackrel{~}{g}\left(\beta \right)/G\left(\mathrm{}\right)\right],`$ where $`\theta `$ is the angular radius from the center of a galaxy image, $`K_\lambda `$ is the $`K`$-correction, $`E_\lambda `$ is the $`E`$-correction including internal absorption by dust, and $`\tau _{\mathrm{HI}}`$ is the optical depth of intergalactic absorption by HI clouds. The present-day color of galaxies of a given type, $`\left(\lambda B\right)_0`$, is calculated by the AB colors based on the present-day SEDs of model galaxies which reproduce the observed colors, and we use $`B_{450}B=0.153`$ to relate the AB magnitudes to the $`B`$ magnitudes of the luminosity function. The radius parameter $`\beta `$ is related to $`\theta `$ as $`\beta =d_A\theta /r_e`$, and $`G\left(\mathrm{}\right)=\left(2n\right)!\pi /a_n^{2n}`$. The isophotal size $`D_{\lambda _D}`$ can be derived from solving the equation $`S_{\lambda _D}\left(D_{\lambda _D}/2\right)=S_{\mathrm{th}}`$. Similarly, the apparent magnitude encircled within $`\theta `$ is given by $`m_\lambda \left(\theta \right)`$ $`=`$ $`M_B+\left(\lambda B\right)_0+K_\lambda (\lambda ,z)+E_\lambda (\lambda ,z)`$ (12) $`+`$ $`2.5\tau _{\mathrm{HI}}(\lambda ,z)\mathrm{log}e+5\mathrm{log}\left[d_L/10\mathrm{p}\mathrm{c}\right]`$ $``$ $`2.5\mathrm{log}\left[\stackrel{~}{G}\left(\beta \right)/G\left(\mathrm{}\right)\right],`$ where $`d_L`$ is the standard luminosity distance, and $`\stackrel{~}{G}\left(\beta \right)`$ is the integrated profile of $`\stackrel{~}{g}\left(\beta \right)`$. The isophotal, aperture, and pseudo-total magnitudes are obtained with $`\theta =D_{\lambda _D}/2`$, aperture/2, and $`\mathrm{}`$, respectively. (Note that $`\stackrel{~}{G}\left(\mathrm{}\right)=G\left(\mathrm{}\right)`$.) The $`K`$ and $`E`$ corrections are calculated from the time-dependent models of spectral energy distribution per unit wavelength $`f_\lambda \left(t\right)`$, after modified to include the dust absorption. They can be written as $`K_\lambda \left(z\right)`$ $`=`$ $`2.5\mathrm{log}\left\{{\displaystyle \frac{1}{\left(1+z\right)}}{\displaystyle \frac{f_{\lambda /\left(1+z\right)}\left(t_0\right)}{f_\lambda \left(t_0\right)}}\right\},`$ (13) $`E_\lambda \left(z\right)`$ $`=`$ $`2.5\mathrm{log}\left\{{\displaystyle \frac{f_{\lambda /\left(1+z\right)}\left(t_z\right)}{f_{\lambda /\left(1+z\right)}\left(t_0\right)}}\right\},`$ (14) where $`t_z`$ is the age of a galaxy at redshift $`z`$ which was formed at $`z_F`$, and $`t_0`$ is its present age. With all these prescriptions, we can numerically solve $`M_B`$ for a given set of $`m_\lambda `$ and $`z`$, and then the number count and redshift distribution of galaxies by means of equation (7). ## 3. Faint Galaxy Number Counts ### 3.1. Importance of the Observational Selection Effects Figure 5 shows the predictions of galaxy number count in the HDF based on the PLE model of galaxy evolution, including the effects of dust and intergalactic absorptions, and the selection effects. Here we have used a ‘standard’ set of model parameters in this paper: $`(h,\mathrm{\Omega }_0,\mathrm{\Omega }_\mathrm{\Lambda })=(0.7,0.2,0.8)`$, $`z_F=5`$, the local luminosity function of the SSRS2 survey, and the screen model of dust. \[Here, $`h=H_0`$/(100km/s/Mpc) as usual.\] The solid line is the total number count of all galaxy types, and the other five lines are those of respective galaxy types. In this paper we will present many calculations of galaxy counts, changing various parameters in order to check the systematic uncertainties which may affect our conclusions. The summary of count calculations is presented in Table 2 with references to the number of figures in this paper. In the following we compare these calculated counts with the HDF data as well as the ground-based data transformed into the AB magnitude system. Here we note that the HDF bandpass filters for the $`U_{300}`$ and $`V_{606}`$ are significantly different from those for the $`U`$, $`V`$, and $`R`$ used in the ground-based observations. We have corrected the magnitudes of such ground-based data, making use of the central wavelength of bandpass filters, the present-day SEDs of different galaxy types, and the relative proportions among the galaxy types. The HDF counts are those of isophotal magnitudes, while published total-magnitude counts are used for other ground-based data for which the selection effects are not as important as for the HDF data. Figure 6 shows the effect of cosmological parameters and the selection effects in the predictions of galaxy number count. The predictions with the selection effects are presented by solid lines, while those without the selection effects by dashed lines. In either cases, the three lines from bottom to top correspond to the EdS universe with $`(h,\mathrm{\Omega }_0,\mathrm{\Omega }_\mathrm{\Lambda })=(0.5,1,0)`$, an open universe with $`(h,\mathrm{\Omega }_0,\mathrm{\Omega }_\mathrm{\Lambda })=(0.6,0.2,0)`$, and a $`\mathrm{\Lambda }`$-dominated flat universe with $`(h,\mathrm{\Omega }_0,\mathrm{\Omega }_\mathrm{\Lambda })=(0.7,0.2,0.8)`$. The dotted line is a prediction without the selection effects in an open universe with $`\mathrm{\Omega }_0=0.1`$. The prescriptions used here are the same as those in the prediction by Pozzetti et al. (1998) which they found to agree with the observed HDF counts. Our prediction shown by this dotted line also gives a good fit to the HDF data, and we have confirmed Pozzetti et al.’s result where no selection effects are taken into account. This suggests that the difference between the PLE models of different authors is not significant in the number count predictions. However, this figure clearly demonstrates the importance of the selection effects in comparison between the predicted and observed number counts of galaxies. The difference between the predicted counts with and without the selection effects attains up to a factor of about 4 at the faintest magnitudes. As a result, the observed HDF counts in excess of the PLE predictions with the selection effects are much larger than previously considered. It is striking that such PLE predictions seem to be short of the observed counts in all the passbands, even in the $`\mathrm{\Lambda }`$-dominated flat universe with $`\mathrm{\Omega }_\mathrm{\Lambda }`$ = 0.8 in which the number count is close to the maximum. In the EdS universe, such a predicted deficit in the PLE model attains up to a factor of more than 10. This large excessive number of HDF galaxies may suggest the number evolution of galaxies, but it depends heavily on the cosmology how much the number evolution is required to explain the HDF data. ### 3.2. Dependence on Model Parameters Before we investigate the effect of number evolution quantitatively, it is necessary to see the uncertainties in the PLE predictions. Figure 7 shows how the PLE predictions depend on models of luminosity evolution and absorptions. The solid line is the standard model presented in Fig. 5. The dotted line is the same as the standard model but with no luminosity evolution. The dot-dashed line is the prediction where the updated luminosity evolution model of Kobayashi et al. (1999) is used rather than the standard model of Arimoto & Yoshii (1987) and Arimoto, Yoshii, & Takahara (1992). The short dashed line is the prediction where the slab distribution of dust is assumed rather than the screen by dust. The long-dashed line is the prediction where no intergalactic absorption is taken into account. This line almost overlaps with the solid line and is not visible except in the $`U_{300}`$ band. Figure 8 shows the effect of changing the galaxy formation epoch $`z_F`$ and local luminosity function. The solid line is the standard model presented in Fig. 5. The short- and long-dot-dashed lines are the predictions with $`z_F`$ = 3 and 10, respectively, rather than $`z_F=5`$ in the standard model. The short- and long-dashed lines are the predictions where the luminosity function of the Stromlo-APM and the CfA redshift surveys are used, respectively, with $`z_F=5`$. Figure 9 shows how the uncertainty in the size-luminosity relation affects the predictions of galaxy number count. The solid line is the standard model presented in Fig. 5, while the dashed line is the prediction when the selection effects are ignored, as shown in Fig. 6. The short- and long-dot-dashed lines are the counts when the $`r_e`$$`L_B`$ relation is shifted by $`\pm `$ 1 $`\sigma `$ in $`\mathrm{\Delta }\left(\mathrm{log}r_e\right)`$. The dotted line is the prediction using the GCE sequence instead of the standard GDE sequence as the size-luminosity relation of elliptical galaxies. (See §2.3.1 for detail.) From these three figures we understand a range of the systematic uncertainties in the PLE predictions, which turns out to be not large enough to save the EdS universe nor an open universe with $`\mathrm{\Omega }_0>0.2`$. These uncertainties are significant at the faintest magnitudes where the $`N`$-$`m`$ relation starts to turn over. However, we point out that the effect of cosmological parameters becomes apparent at brighter magnitudes where the uncertainties remain much less significant. In fact, we will show that the slope of the $`N`$-$`m`$ relation at $`B_{450}`$ = 22–26 can be used to discriminate between the effects of cosmological parameters and galaxy number evolution. In order to demonstrate the above statement quantitatively, the systematic model uncertainties at $`m`$= 25 and 28 are summarized in Table 3. The uncertainty at $`m=25`$, to be compared with the effect of cosmological parameters, is dominated by the dust distribution model, but we note that the estimated change from a standard screen model to the slab model is somewhat overestimated because the slab model is clearly inconsistent as a model of dust distribution in starburst galaxies, as mentioned in §2.2. ### 3.3. Mergers and Number Evolution The PLE predictions fall considerably short of the observed HDF counts in the EdS universe and an open universe, and it is still the case even in a $`\mathrm{\Lambda }`$-dominated flat universe. Here, by using a simple model of mergers introduced in §2.4, we investigate whether the number evolution explains the large number of faint HDF galaxies. Figure 10 shows the effect of introducing such a number evolution model in a $`\mathrm{\Lambda }`$-dominated flat universe. The solid line is the prediction with the merger parameters of $`(\eta ,\xi )=(1,3)`$, while the dotted line is the standard PLE prediction with no number evolution (Fig. 5). The short- and long-dot-dashed lines are the predictions with $`(\eta ,\xi )=(1,2)`$ and (1, 4), respectively, showing that the effect of changing $`\xi `$ is not significant. This result indicates that a modest number evolution with $`\eta 1`$ \[$`\varphi ^{}\left(1+z\right)^\eta `$\] is sufficient to explain the observed HDF counts in a $`\mathrm{\Lambda }`$-dominated flat universe with $`\mathrm{\Omega }_\mathrm{\Lambda }=0.8`$, and an even stronger number evolution with $`\eta \stackrel{>}{}1`$ is rejected by the data. Next we consider the EdS universe where the PLE count prediction is by more than one order of magnitudes smaller than those observed in the HDF (Fig. 11). The predictions with $`\eta `$ = 2, 3, 4, and 5 are shown by four solid lines in order from bottom to top, with a fixed value of $`\xi =3`$. The dotted line is the PLE prediction without number evolution. This result indicates that a strong number evolution with $`\eta \stackrel{>}{}`$ 3–4 is necessary to explain the HDF counts in the EdS universe. We note that, while the strong number evolution explains the counts at the faintest magnitudes, it fails to explain the overall shape or slope of the $`N`$-$`m`$ relation. This failure is clearly seen in the $`B_{450}`$ band, where the strong number evolution makes the count slope less steep and deviate from the observations most prominently at $`B_{450}`$ = 22–26. This argument is quite robust, because there should be an upper bound in the steepness of the $`N`$-$`m`$ slope ($`d\mathrm{log}N/dm<`$ 0.4), when the total luminosity density of all galaxies is conserved during the merger process. In a more realistic case such as gas-rich mergers inducing starbursts, an even flatter slope is predicted, because galaxies before mergers are always fainter than those in the case of luminosity-density conservation. If one tries to explain the observed steep slope by mergers, it is necessary to contribe a merger process where galaxies before mergers are always brighter, in other words, the luminosity density of galaxies increases if the merger process is traced backwards—which we consider quite unrealistic. Therefore, we conclude that a strong number evolution in the EdS universe is unlikely to explain the observed counts over the whole range of apparent magnitudes. Figure 12 shows the effect of introducing a number evolution model in an open universe with $`\mathrm{\Omega }_0=0.2`$. The lines in this figure have the same meanings as in Fig. 11, but the four solid lines are the predictions with $`\eta `$ = 1, 2, 3, and 4 in order from bottom to top. In this open universe, a number evolution with $`\eta \stackrel{>}{}2`$ is necessary to explain the faintest counts, but again it can not explain the steep slope of the $`N`$-$`m`$ relation in the $`B_{450}`$ and $`V_{606}`$ bands. This indicates that an open universe is also difficult to explain the HDF counts if $`\mathrm{\Omega }_0>0.2`$. However, a lower-density open universe, for example, with $`\mathrm{\Omega }0.1`$ might give a similar result with a $`\mathrm{\Lambda }`$-dominated flat universe with $`\mathrm{\Omega }_0=0.2`$ (see Fig. 6), and such an open universe could also explain the HDF counts if a modest number evolution of galaxies is taken into account. ## 4. Photometric Redshift Distribution Although not as reliable as spectroscopic redshifts, photometric redshifts of galaxies are useful for statistical studies of high-redshift galaxies. Several groups have published catalogs of photometric redshifts for the HDF galaxies (e.g., Sawicki, Lin, & Yee 1997; Wang, Bahcall, & Turner 1998; Fernández-Soto, Lanzetta, & Yahil 1999). Here we compare our theoretical model with the photometric redshift distributions reaching $`I_{814}=28`$ derived by Fernández-Soto et al. (1999), which utilizes not only the optical photometry of the HDF but also the information of the near-infrared $`J`$, $`H`$, and $`K`$ bands. We have also compared our model with the other two catalogs of photometric redshifts by Sawicki et al. (1997) and Wang et al. (1998), and confirmed that the following result is hardly changed. Figure 13 shows the observed redshift distribution in three $`I_{814}`$ magnitude ranges. The model curves are our predicted redshift distributions of galaxies in a $`\mathrm{\Lambda }`$-dominated flat universe with $`\mathrm{\Omega }_\mathrm{\Lambda }=0.8`$. The area over which the models are calculated is chosen to coincide with the sky area covered by the observational analysis of Fernández-Soto et al. (1999): 5.31 arcmin<sup>2</sup> for $`I_{814}<26`$ and 3.92 arcmin<sup>2</sup> for $`I_{814}>26`$. Therefore, not only the shape but also the normalization of the predicted redshift distributions can be compared directly with the data. The solid and dashed lines are the models with $`(\eta ,\xi )=(0,3)`$ and (1, 3), respectively, with the selection effects taken into account. The dot-dashed line is the same as the dashed line with $`(\eta ,\xi )=(1,3)`$, but no selection effects are taken into account. It is clear that the selection effects give a bias against high-redshift galaxies, and this selection bias is significant especially at the faintest magnitudes. It is inevitable to include these effects when one uses the redshift distribution as a probe of number evolution of galaxies. Comparison with the data shows that a modest merger model with $`\eta 1`$ in a $`\mathrm{\Lambda }`$-dominated flat universe gives a reasonable fit to the photometric redshift distribution as well as galaxy counts, provided that the selection effects are properly taken into account. Figure 14 compares the observed redshift distribution with the predictions in the EdS universe. The solid line is the prediction without number evolution, while other lines are those with $`\eta `$ = 3, 4, and 5 with a fixed value of $`\xi `$=3. The selection effects are taken into account in all curves. A strong number evolution predicts that most of galaxies have lower redshifts of $`z\stackrel{<}{}`$ 1, deviating significantly from the observed distribution for $`23<I_{814}<26`$. If the assumption of conserved luminosity density is relaxed, the predicted distribution becomes peaked at an even lower redshift, because pre-merger galaxies in more realistic merger models are fainter than expected from the conserved luminosity density, as discussed in the previous section. This gives another argument that the EdS universe can not explain the observed number of HDF galaxies even if a strong merger is invoked, in addition to the argument in the previous section against the EdS universe based on the slope of the observed $`N`$-$`m`$ relation. Figure 15 is similar to Fig. 14, but for an open universe with $`\mathrm{\Omega }_0=0.2`$. The predicted redshift distribution is still peaked at a lower redshift compared with the observed distribution. This discrepancy is significant for $`23<I_{814}<26`$, but not as serious as in the EdS universe. ## 5. Discussion In this paper we have shown that a strong number evolution with $`\eta \stackrel{>}{}`$ 3–4 is necessary to explain the HDF counts at the faintest magnitudes in the EdS universe. The photometric redshift distribution of HDF galaxies suggests that a significant number of the faintest galaxies are at $`z<1`$, and hence there must be a strong number evolution already operated at $`z<1`$ to increase the number of galaxies by factor of about 10. Therefore, further argument for or against the EdS universe with strong number evolution of galaxies can be made from the observational constraints based on spectroscopic redshift surveys at $`z<1`$. Totani & Yoshii (1998) argued that such a strong number evolution at $`z<1`$ is clearly inconsistent with the spectroscopic catalogs of galaxies reaching $`z1`$, at least for giant galaxies with $`LL^{}`$. This is based on a $`V/V_{\mathrm{max}}`$ test for the galaxies in the Canada-France Redshift Survey (CFRS, Lilly et al. 1995), and an important finding is that the PLE model is not inconsistent with the spectroscopic data, giving a constraint on number evolution as $`\eta =1.8\pm 0.7`$, $`1.1\pm 0.7`$, and $`0.5\pm 0.7`$ for the EdS universe, an open $`\mathrm{\Omega }_0=0.2`$ universe, and a flat $`\mathrm{\Omega }_\mathrm{\Lambda }=0.8`$ universe, respectively. Although the $`V/V_{\mathrm{max}}`$ test favors a larger $`\eta `$ in the EdS universe, this value seems smaller than that required to explain the HDF counts at the faintest magnitudes. These results have been confirmed for elliptical galaxies by Shade et al. (1999), in which they found that the population of massive early-type galaxies was largely in place by $`z1`$. An independent constraint on galaxy mergers at $`z<1`$ comes from statistical studies of merging galaxies inferred from high-resolution images. Recently, Le Févre et al. (2000) derived the evolution of merger rate from the HST images of the CFRS and LDSS galaxies. Their result suggests that $`L^{}`$ galaxies on the average have undergone about one merger event from $`z=1`$ to 0, which corresponds to $`\eta 1`$ and hence this is consistent with the number evolution in a $`\mathrm{\Lambda }`$-dominated flat universe suggested by this paper. Pozzetti et al. (1998) claimed that the PLE model can not explain all the observed data, although it well explains the HDF galaxy counts. The major discrepancies between the PLE model and the observed data were found in the evolution of the luminosity density in the universe. That is, the observed luminosity density increases more steeply to $`z1`$ than the PLE model prediction, and on the other hand, the PLE model predicts too high UV luminosity density at $`3.5<z<4.5`$ compared with the observation, because of intense starbursts in elliptical galaxies. However, Totani, Yoshii, & Sato (1997) had already pointed out that the observed steep evolution to $`z1`$ is explained in a $`\mathrm{\Lambda }`$-dominated flat universe. It has already been argued that the PLE model is consistent with the observed luminosity density evolution if initial starbursts in high-redshift elliptical galaxies are obscured or not existent. In fact, our calculation of the redshift distribution in this paper shows that initial starbursts are not detected at $`z\stackrel{>}{}3`$, because of the dust obscuration and the selection effects. Therefore, the major problems of the PLE model claimed by Pozzetti et al. (1998) are resolved, and the PLE model in a $`\mathrm{\Lambda }`$-dominated flat universe gives a reasonable fit to the observed HDF data, allowing only for a modest number evolution of galaxies with $`\eta \stackrel{<}{}1`$. ## 6. Summary and Conclusions We have modeled the number count and redshift distribution of faint HDF galaxies, with the observational selection effects properly taken into account. As a consequence of the selection effects in the theoretical modeling, predicted counts from the PLE model are smaller than previously considered, and they are more than 10 times smaller than the observed HDF counts at the faintest magnitudes in the EdS universe. A strong number evolution with $`\eta \stackrel{>}{}`$ 3–4 under the assumption of conserved luminosity density is required to explain the faintest counts in this EdS universe, when the number evolution is parametrized as $`\varphi ^{}\left(1+z\right)^\eta `$ and $`L^{}\left(1+z\right)^\eta `$. However, such a strong number evolution is not consistent with the overall $`N`$-$`m`$ slope or the photometric redshift distribution. These discrepancies become even worse when one considers a more realistic merger process, i.e., enhanced star formation following by gas-rich mergers. In addition, such a strong evolution is rejected at least for average $`L^{}`$ galaxies at $`z<1`$ from the data of spectroscopic redshift surveys (Totani & Yoshii 1998; Shade et al. 1999; Le Févre et al. 2000). Therefore, we conclude that it is almost impossible to explain the HDF galaxies in the EdS universe, unless we invoke ultra-exotic galaxy populations such as galaxies forming only massive stars at high redshifts to escape from local galaxy surveys due to the complete lack of long-lived stars. The present work revitalizes the practice of using faint number counts as an important cosmological test, which gives one of the arguments against the EdS universe by its outstanding statistics compared with other cosmological tests. An open universe with $`\mathrm{\Omega }_0>0.2`$ does not fit to the HDF data either, for the similar reasons for rejecting the EdS universe. An open universe with $`\mathrm{\Omega }_00.1`$ might be consistent with the HDF data, but such a low value of $`\mathrm{\Omega }_00.1`$ would not be reconciled with other constraints on $`\mathrm{\Omega }_0`$, such as the baryon-gas to dark-matter mass ratio in poor clusters of galaxies combined with the standard big-bang nucleosynthesis prediction of baryon mass density in the universe (e.g., Pedersen, Yoshii, & Sommer-Larsen 1997). We have extensively checked systematic uncertainties in our theoretical modeling of galaxy formation and evolution, and found that they are unlikely to resolve the above discrepancies emerged in the EdS universe and also in an open universe. On the other hand, such discrepancies are naturally resolved if we invoke a $`\mathrm{\Lambda }`$-dominated flat universe. This suggests that the existence of the cosmological constant or an exotic form of the vacuum energy density of the universe which is now accelerating the expansion of the universe. The PLE model in a $`\mathrm{\Lambda }`$-dominated flat universe with $`\mathrm{\Omega }_00.2`$ gives a reasonable fit to the HDF data, and a modest number evolution with $`\eta \stackrel{<}{}1`$ is also suggested by the HDF counts at the faintest magnitudes. It should be noted that this number evolution does not necessarily mean mergers of galaxies, but may suggest strongly clumpy star-forming regions within an individual galaxy system becoming visible at high redshifts (Colley et al. 1996; Bunker, Spinrad, & Thompson 1999). On the other hand, it is interesting to note that this indication of mild number evolution is consistent with the merger rate evolution of $`L^{}`$ galaxies at $`z<1`$ recently inferred from a high-resolution image study for galaxies in the CFRS survey (Le Févre et al. 2000). This result is consistent with some models of galaxy formation based on the hierarchical structure formation in the CDM universe (Le Févre et al. 2000), although there are considerable uncertainties in the theoretical calculations for the merging history of baryonic component. A stronger number evolution with $`\eta \stackrel{>}{}1`$ is, however, strongly disfavored by the observed HDF galaxy counts. This will give an important constraint when galaxy formation is modeled in the framework of the structure formation in a cold dark matter universe. Inclusion of the selection effects in this paper leads to a considerably different result from previous studies on galaxy number count and redshift distribution. This means that any cosmological interpretations will be seriously misled if the selection effects are ignored. All future studies related to the detection and statistics of high-redshift galaxies should take into account these effects. The selection effects give a bias against high-redshift galaxies, reducing a problem of overprediction of such galaxies by the PLE model, which has been claimed by several studies ignoring the selection effects (e.g., Ellis 1997). In fact, we have shown that the PLE model is in overall agreement with the HDF galaxies, even if a modest number evolution of galaxies ($`\eta \stackrel{<}{}1`$) may be required. A strong number evolution, however, predicts too small a number of high-redshift galaxies to be consistent with the photometric redshift distribution of the HDF galaxies. The authors would like to thank K. Shimasaku for providing numerical data for the filter functions of HST photometry bands, and T. Tsujimoto and C. Kobayashi for providing their models of galaxy evolution in a tabular form. We also thank an anonymous referee for many useful comments which have considerably improved this manuscript. This work has been supported in part by a Grand-in-Aid for Conter-of-Excellence Research (07CE2002) of the Ministry of Education, Science, and Culture in Japan.
warning/0004/cond-mat0004144.html
ar5iv
text
# Collective Spin Fluctuation Mode and Raman Scattering in Superconducting Cuprates Electronic Raman scattering has proven to be a useful tool in exploring the superconducting state in the cuprate materials. The possibility of probing selectively electronic excitations in different regions of the Brillouin zone (BZ) by the choice of polarization geometries has allowed to explore the superconducting gap anisotropy. The successful explanation of the Raman data in $`B_{1g}`$ and $`B_{2g}`$ scattering geometries has provided one piece of evidence for the by now widely accepted $`d_{x^2y^2}`$ pairing symmetry in hole-doped cuprate superconductors . In the context of impurity effects as a testing ground for unconventional superconductivity, the observed $`\omega ^3`$ to $`\omega `$ crossover in the low frequency $`B_{1g}`$ Raman response fits consistently with the power law crossovers at low temperatures in the NMR relaxation rate and in the magnetic penetration depth. An even quantitatively consistent picture of electronic Raman scattering and infrared conductivity was achieved when the T-matrix approximation in the “dirty” d-wave scenario is extended to include a spatial extension of the impurity potential . However up to now the discrepancy between Raman data in $`A_{1g}`$ and $`B_{1g},B_{2g}`$ geometries has remained unresolved . Previous results for the A<sub>1g</sub> scattering geometry were found to be very sensitive to changes in the Raman vertex function $`\gamma (𝐤)`$ making a comprehensive explanation difficult for the experimental data in different cuprate materials. In this paper we calculate the Raman response of a $`d_{x^2y^2}`$ superconductor including the contribution from a collective spin fluctuation (SF) mode which is identified with the $`(\pi ,\pi ,\pi )`$-resonance (in short $`\pi `$-resonance) near $`\omega _R41`$ meV observed by inelastic neutron scattering (INS) on bilayer cuprates . Our results suggest that the $`A_{1g}`$ peak position is largely controlled by the strength and frequency of the $`\pi `$-resonance mode which on the other hand does not affect the Raman response in the $`B_{1g}`$ and $`B_{2g}`$ channels. The inclusion of the collective SF mode allows for a simultaneous fit of the Raman data in all channels in optimally doped materials. Furthermore we find that the inclusion of the SF term significantly reduces the sensitivity to the special choice of the underlying tight-binding band structure, i.e. the sensitivity to the choice of the Raman vertex in $`A_{1g}`$ symmetry, resolving the previously encountered problems in the symmetry analysis of the light scattering amplitude . On the basis of the observation of the collective SF mode in Y-123 and Bi-2212, we consider a bilayer model represented by a tight-binding band structure $$ϵ_𝐤=2t(\mathrm{cos}(k_x)+\mathrm{cos}(k_y))+4t^{}\mathrm{cos}(k_x)\mathrm{cos}(k_y)t_{}(𝐤),$$ with an inter-plane hopping given by $$t_{}(𝐤)=2t_{}\mathrm{cos}(k_z)(\mathrm{cos}(k_x)\mathrm{cos}(k_y))^2$$ (1) where $`k_z`$ is $`0`$ or $`\pi `$ for the bonding or anti-bonding bands of the bilayer, respectively. The spin susceptibility ($`\chi _s`$) is modeled by extending the weak coupling form of a BCS superconductor in a $`d_{x^2y^2}`$ pairing state to include antiferromagnetic spin fluctuations by a RPA form with an effective interaction $`\overline{U}`$; i.e. $`\chi _s(𝐪,i\omega )=\chi ^0(𝐪,i\omega )/(1\overline{U}\chi ^0(𝐪,i\omega ))`$ where $$\chi ^0(𝐪,i\omega )=\frac{1}{\beta }\text{Tr}\underset{𝐤,i\omega ^{}}{}\widehat{G}(𝐤,i\omega ^{})\widehat{G}(𝐤+𝐪,i\omega ^{}+i\omega ).$$ (2) Tr denotes the trace and $`\beta =T^1`$. $`\widehat{G}(𝐤,i\omega )`$ is the BCS Green’s function in Nambu space $$\widehat{G}(𝐤,i\omega )=\frac{i\omega \widehat{\tau }_0+\xi _𝐤\widehat{\tau }_3+\mathrm{\Delta }_𝐤\widehat{\tau }_1}{(i\omega )^2\xi _𝐤^2\mathrm{\Delta }_𝐤^2}$$ (3) with $`\widehat{\tau }_i`$ (i=1,2,3) the Pauli matrices, $`\widehat{\tau }_0`$ the $`2\times 2`$ unit matrix, $`\xi _𝐤=ϵ_𝐤\mu `$ and $`\mathrm{\Delta }_𝐤=\mathrm{\Delta }_0[\mathrm{cos}(k_x)\mathrm{cos}(k_y)]/2`$. This form of the spin susceptibility contains a strong magnetic resonance peak at $`𝐪=𝐐(\pi ,\pi ,\pi )`$ which was proposed to explain the INS resonance at energies near 41 meV in Y-123 and Bi-2212 . Other forms for the spin susceptibility can be straightforwardly used within our model. However, the results are mainly determined by the collective mode at $`𝐐`$. Therefore we take the bilayer susceptibility for a representative calculation. The intensity of scattered light $`I(\omega )`$ in Raman experiments is proportional to the imaginary part of the response function for the effective density operator $$\stackrel{~}{\rho }_𝐪=\underset{𝐤,\sigma }{}\widehat{\gamma }(𝐤)c_{\sigma ,𝐤+𝐪}^{}c_{\sigma ,𝐤}$$ (4) in the long wavelength limit $`𝐪\mathrm{𝟎}`$. Specifically $`I(\mathrm{\Omega })`$ $``$ $`\left(1+n(\mathrm{\Omega })\right)\text{Im}\chi (\mathrm{\Omega }+i0^+)`$ (5) $`\chi (i\mathrm{\Omega })`$ $`=`$ $`{\displaystyle _0^{1/T}}d\tau e^{i\mathrm{\Omega }\tau }T_\tau [\stackrel{~}{\rho }(\tau ),\stackrel{~}{\rho }(0)],`$ (6) with the Bose function $`n(\omega )`$ and the time ordering operator $`T_\tau `$. The bare Raman vertex $`\widehat{\gamma }(𝐤)=\tau _3\gamma (𝐤)`$ in different scattering geometries are classified according to the elements of the $`D^{4h}`$ point group. For the limiting case of vanishingly small scattered ($`\omega _S`$) and incident ($`\omega _I`$) photon energies, it can be represented in the effective-mass approximation (EMA) $$\gamma (𝐤)=\underset{\alpha ,\beta }{}e_\alpha ^I\frac{^2ϵ_𝐤}{k_\alpha k_\beta }e_\beta ^S$$ (7) where $`𝐞^I`$ and $`𝐞^S`$ are the unit vectors for in-plane polarizations (i.e. $`\alpha ,\beta \{x,y\}`$) of the incoming and the scattered light, respectively. Using Eq. (7) and the bilayer tight-binding dispersion Eq. 1 we obtain $`\gamma _𝐤^{B_{1g}}`$ $`=`$ $`2t\gamma _𝐤^d\left(1+{\displaystyle \frac{4t_{}}{t}}\mathrm{cos}(k_z)[\mathrm{cos}(k_x)+\mathrm{cos}(k_y)]\right)`$ (8) $`\gamma _𝐤^{A_{1g}}`$ $`=`$ $`2t\gamma _𝐤^s2t_{}\mathrm{cos}(k_z)(\mathrm{cos}(2k_x)+\mathrm{cos}(2k_y))`$ (10) $`4\mathrm{cos}(k_x)\mathrm{cos}(k_y)[t^{}+2t_{}\mathrm{cos}(k_z)]`$ where $`\gamma _𝐤^{d,s}=(\mathrm{cos}(k_x)\mathrm{cos}(k_y))/2`$. However, the EMA has a questionable region of validity for all Raman measurements on the cuprates since the incoming photons have energy $`2`$eV, which is on the order of the bandwidth and the inter-band excitations according to local density calculations. EMA based arguments in previous works about relative Raman intensities for different channels are therefore questionable . We hence consider other forms for the vertices as well which obey the proper symmetry transformations. For the $`A_{1g}`$ geometry some symmetry compatible choices are $`\gamma (𝐤)=\mathrm{cos}(k_x)+\mathrm{cos}(k_y)`$ and $`\gamma (𝐤)=\mathrm{cos}(k_x)\mathrm{cos}(k_y)`$. These basis functions assign weight to different regions of the BZ and this is the reason why previous results for the $`A_{1g}`$ response were particularly sensitive to the specific choice of the bare Raman coupling vertex. The spin fluctuations lead to an additional contribution for the Raman response via a 2-magnon like process as shown diagrammatically in Fig. 1 . Here, the SF propagator is incorporated in its RPA form for the bilayer (as described above) by the ladder diagram series with an effective on-site Hubbard interaction $`\overline{U}`$. We therefore write the Raman response function at finite temperature as the sum of a pair-breaking (PB) and a SF contribution $$\chi _{\gamma \gamma }(𝐪,i\omega )=\chi _{\gamma \gamma }^{PB}(𝐪,i\omega )+\chi _{\gamma \gamma }^{SF}(𝐪,i\omega )$$ (11) with the Raman vertex specifying the scattering geometry. In the limit $`𝐪\mathrm{𝟎}`$ the diagram for the SF contribution translates into $`\chi _{\gamma \gamma }^{SF}(i\mathrm{\Omega })`$ $`=`$ $`{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{𝐪^{},i\omega }{}}V^\gamma (𝐪^{},i\mathrm{\Omega },i\omega )\chi _s(𝐪^{},i\omega )`$ (13) $`\times \chi _s(𝐪^{},i\omega +i\mathrm{\Omega })V^\gamma (𝐪^{},i\mathrm{\Omega },i\omega ).`$ The vertex function $`V^\gamma (𝐪^{},i\mathrm{\Omega },i\omega )`$ includes the bare Raman vertex and is evaluated as $`V^\gamma (𝐪^{},i\mathrm{\Omega },i\omega )=\text{Tr}\{{\displaystyle \frac{1}{\beta }}{\displaystyle \underset{𝐤,i\omega ^{}}{}}\widehat{\gamma }(𝐤)\widehat{G}(𝐤,i\omega ^{}+i\mathrm{\Omega })\widehat{\tau }_0\overline{U}`$ (14) $`\widehat{G}(𝐤+𝐪^{},i\omega ^{}+i\mathrm{\Omega }+i\omega )\tau _0\overline{U}\widehat{G}(𝐤,i\omega ^{})\}.`$ (15) $`i\mathrm{\Omega }`$, $`i\omega `$ denote bosonic and $`i\omega ^{}`$ fermionic Matsubara frequencies. Similarly, $`\chi _{\gamma \gamma }^{PB}`$ is evaluated as $$\chi _{\gamma \gamma }^{PB}(i\mathrm{\Omega })=\frac{\mathrm{Tr}}{\beta }\underset{𝐤,i\omega ^{}}{}\widehat{\gamma }(𝐤)\widehat{G}(𝐤,i\omega ^{})\widehat{\gamma }(𝐤)\widehat{G}(𝐤,i\omega ^{}+i\mathrm{\Omega }).$$ (16) The total Raman response is calculated in the gauge invariant form which results from taking into account the long wavelength fluctuations of the order parameter to guarantee local charge conservation . The total Raman susceptibility thus follows as $$\chi (i\mathrm{\Omega })=\chi _{\gamma \gamma }(i\mathrm{\Omega })\frac{\chi _{\gamma 1}^2(i\mathrm{\Omega })}{\chi _{11}(i\mathrm{\Omega })}$$ (17) where $`\chi _{1\gamma }`$ and $`\chi _{11}`$ are obtained by the replacement $`\gamma (𝐤)1`$ in one or both bare Raman vertices in the vertex function Eq. (15). The analytical continuation to the real axis is performed using Padé approximants. The band structure parameters are chosen for all the numerical calculations to be applicable to optimally doped systems: $`n=0.85`$, $`t^{}/t=0.45`$, $`t_{}/t=0.1`$ while the gap has been chosen as $`\mathrm{\Delta }_0/t=0.25`$. We have evaluated the Raman response at a temperature $`T/t=0.08`$. Let’s first consider $`\chi _{SF}`$ alone. In Fig. 2 we plot $`\chi _{SF}^{\prime \prime }`$ for the $`A_{1g}`$ channel versus frequency. $`\chi _{SF}`$ is a convolution of two spin susceptibilities (Eq. (13)) and is thus peaked at an energy near twice the magnetic resonance energy. An important point is that in the $`B_{1g}`$ and $`B_{2g}`$ geometries the SF term introduces vanishingly small corrections to the total response, rendering the presence of the SF term important only in the A<sub>1g</sub> geometry. This is due to the sharpness in momentum space of the resonance peak at $`𝐐`$ in the SF propagator. In fact, if the transfer is taken only at $`𝐐`$, both the $`B_{1g}`$ and $`B_{2g}`$ contributions to $`\chi _{SF}`$ vanish identically, as can been seen from Eq. 15. Therefore for these channels the response is given by the PB term alone. In Fig. 3 we illustrate the frequency dependence of $`\chi ^{\prime \prime }(\mathrm{\Omega })`$ in the A<sub>1g</sub> geometry for three different values of the effective interaction with both the PB and SF contributions included. The shape of the Raman response is modified varying $`\overline{U}/t`$ and in particular the position of the resonance is shifted towards higher energies for increasing $`\overline{U}`$. With the inclusion of the SF term we now obtain a peak slightly above $`\mathrm{\Delta }_0`$ as observed experimentally in Y-123 and Bi-2212. A comment is in order on the relative magnitude of the SF and the PB term. Comparing Figs. 2 and 3 it is clear that the SF term is much smaller than the PB term, but the effect of this new term is nevertheless visible since the backflow (the second term in Eq. (17)) mixes in a non trivial way the two contributions. Since the SF term varies as $`\overline{U}^4`$ in our model it starts to dominate for larger $`\overline{U}`$, leading to a shift in spectral weight out towards $`2\omega _R`$, which also changes with $`\overline{U}`$. The experimental position of the peak is not sample dependent in the $`A_{1g}`$ geometry as already mentioned, and is almost the same in different cuprates. On the other hand the theoretical description with the PB term alone is very sensitive to Raman vertex changes, which can produce variation of its position between $`\mathrm{\Delta }_0`$ and $`2\mathrm{\Delta }_0`$ , not allowing for a comprehensive modeling of different cuprates. In Fig. 4 we address the problem of the sensitivity of the result to changes in the bare Raman vertex $`\gamma (𝐤)`$. To investigate the effect of changes of the vertex function, we have calculated the final response using the three different forms for the vertex $`\gamma (𝐤)=\mathrm{cos}(k_x)+\mathrm{cos}(k_y)`$, $`\gamma (𝐤)=\mathrm{cos}(k_x)\mathrm{cos}(k_y)`$ and in the EMA which posses the correct transformation properties required by symmetry. In the first panel of Fig. 4, we have plotted the Raman response for $`\overline{U}=0`$, i.e. the PB contribution alone, and in the second panel for $`\overline{U}/t=1.3`$. All curves are renormalized to their peak height to allow for an easier comparison. Clearly the strong sensitivity to changes of the bare Raman vertex (first panel) is much reduced when the SF term is added (second panel). In Fig. 5 we compare the theoretical results with experimental data on optimally doped Bi-2212 . Adding the SF contribution leads to a shift of the peak position from near $`\mathrm{\Delta }_0`$ for $`\overline{U}=0`$ to higher frequencies, and thus to a better agreement with the experimental relative peak positions in $`A_{1g}`$ and $`B_{1g}`$ geometries. For the fit we have adjusted $`t`$ to achieve a good agreement with the $`B_{1g}`$ channel, and then adjusted $`\overline{U}`$ to match the $`A_{1g}`$ peak position. The value of $`t`$ obtained from the fit is $`t=130`$ meV. This value has to be compared with $`t105`$ meV, which results from the condition $`\omega _R40`$ meV. This slight discrepancy is most probably related to our simple modeling of the propagators which neglects strong renormalizations from interactions as well as impurities. From this work we conclude that including the SF contribution in the Raman response solves the previously unexplained sensitivity of the $`A_{1g}`$ response to small changes in the Raman vertex. Also, within our model it is now possible to obtain the correct relative peak positions of the $`A_{1g}`$ and the $`B_{1g}`$ scattering geometry. Whereas the SF (two-magnon) contribution controls the $`A_{1g}`$ peak, the $`B_{1g}`$ and $`B_{2g}`$ scattering geometries are essentially unaffected and determined by pair breaking processes alone. We would like to thank R. Hackl for numerous discussions. One of the authors (F.V.) would like to thank the Gottlieb Daimler and Karl Benz Foundation for financial support. This work was partially supported by the Deutsche Forschungsgemeinschaft through SFB 484.
warning/0004/astro-ph0004195.html
ar5iv
text
# Hybrid Thermal-Nonthermal Synchrotron Emission from Hot Accretion Flows ## 1 Introduction The mechanisms of particle heating and acceleration, and the emission spectra from the resulting particle energy distributions, are of great importance in the theory of collisionless hot accretion flows onto compact objects. Discussions in the literature have focused on the physics of electron heating and acceleration (Begelman & Chiueh 1988; Bisnovatyi-Kogan & Lovelace 1997; Quataert & Gruzinov 1999; Gruzinov & Quataert 1999, Medvedev 2000), the efficiency of particle thermalization (Ghisellini, Guilbert, & Svensson 1988; Ghisellini, Haardt, & Fabian 1993; Mahadevan & Quataert 1997; Ghisellini, Haardt, & Svensson 1998; Nayakshin & Melia 1998), and the generation of hybrid thermal-nonthermal electron energy distributions in these plasmas (see, e.g., Coppi 1999 and references therein). Despite a substantial amount of work, many issues remain unresolved, primarily because of our incomplete understanding of physical processes such as magnetic reconnection, MHD turbulence, and collective plasma modes. These questions are especially relevant for an optically-thin advection-dominated accretion flow (ADAF). An ADAF is an example of a hot, rarefied, magnetic plasmas with low radiative efficiency (Ichimaru 1977; Narayan & Yi 1994, 1995b; Abramowicz et al. 1995; see Narayan, Mahadevan, & Quataert 1998b and Kato, Fukue, & Mineshige 1998 for reviews). A basic property of the nearly collisionless plasma in an ADAF is that the Coulomb coupling between the electrons and ions is weak so that energy transfer from the ions to the electrons is inefficient. In addition, it is commonly assumed that the viscously generated energy primarily heats the heavier species, the ions, and that the electrons retain a thermal distribution throughout the flow (e.g., Narayan & Yi 1995b; Mahadevan 1997). There are, however, processes such as MHD turbulence, pair production (e.g., through pion decay), and electron-proton coupling which can both heat the electrons and generate non-thermal distributions. Quataert and Gruzinov (1999; see also Gruzinov & Quataert 1999) considered two processes specific to MHD turbulence that accelerate particles in magnetic collisionless plasmas: Landau damping by electric fields parallel to the local magnetic field and transit-time damping by time-varying magnetic fields. They found that the assumption of negligible electron heating/acceleration is valid only for weak magnetic fields, i.e., when the ratio of the gas pressure to total pressure $`\beta _{\mathrm{ADAF}}`$ is larger than a critical value $`\beta _{\mathrm{crit}}`$. The value of $`\beta _{\mathrm{crit}}`$ is very uncertain and is around 0.9. For $`\beta _{\mathrm{ADAF}}0.9`$, turbulence primarily accelerates the protons, while for stronger magnetic fields, the results are inconclusive. Shocks and pion decay can also lead to non-thermal electrons in the accretion flow. Several processes, such as Coulomb collisions and synchrotron self-absorption, can potentially lead to thermalization of particles in accretion flows (Mahadevan & Quataert 1997; Ghisellini et al. 1998). Mahadevan & Quataert (1997) showed that Coulomb collisions are ineffective in thermalizing the electrons in an ADAF. However, they argued that, for sufficiently high mass-accretion rates, the electrons in the plasma can be thermalized by synchrotron self-absorption. Nayakshin and Melia (1998) showed that considerable deviations from a Maxwell-Boltzmann distribution can be sustained in a plasma with low source compactness when Coulomb collisions, Comptonization, and pair processes are taken into account. In view of the difficulty of calculating the heating, cooling and thermalization of particles from first principles, many authors have restricted their models to either purely Maxwellian or purely non-thermal (extended power laws or monoenergetic) electron distributions. Only recently have there been attempts towards explaining spectra of accreting black holes with models including hybrid thermal/non-thermal distributions of electrons. Some models have physically motivated distributions, such as non-thermal electrons produced by decaying pions (Mahadevan 1998), whereas others invoke more ad hoc distributions to fit the data (e.g., Beckert & Duschl 1997; Falcke & Biermann 1999). The reverse process of trying to constrain the energy distributions of particles in accreting plasmas by comparing models to data, however, faces issues of uniqueness which can be addressed only by a more comprehensive study of hybrid models. In this paper, we consider generalized electron distributions consisting of a dominant Maxwellian plus a small non-thermal power-law component of varying slope and energy content, and study the synchrotron emission from the resulting hybrid plasmas. We identify the characteristic signatures of the non-thermal electrons on the emitted radio synchrotron spectrum of an accretion flow and on its image as observed with a radio telescope. We also discuss to what extent the observed effects could be used to determine the details of the underlying non-thermal electron distribution. This work is relevant for interpreting observations of low luminosity AGNs such as the Galactic Center source, Sagittarius A (Sgr A). In $`\mathrm{\S }2`$ we review the basic properties of ADAFs, hybrid plasmas, synchrotron radiation, and radiative transfer. In $`\mathrm{\S }3`$ we present a series of models of supermassive black holes with low accretion rates. We apply these models to Sgr A and derive constraints on the fraction of the electron energy that can be present in a non-thermal form. In $`\mathrm{\S }4`$, we study the correspondence between the energy distribution of the electrons and the resulting spectrum. In $`\mathrm{\S }5`$ we study the energetics of a hybrid flow and calculate the heating, cooling and thermalization rates of the non-thermal electrons. We summarize our conclusions in $`\mathrm{\S }6`$. We present in an Appendix approximate analytic expressions for the contribution of a non-thermal particle population to the synchrotron spectrum of an accretion flow. ## 2 Formalism ### 2.1 Advection-Dominated Flows We begin by reviewing some of the basic properties of the optically-thin branch of ADAFs. ADAFs are quasi-spherical, hot, magnetic accretion flows in which the accreting plasma is too rarefied to cool efficiently by radiative processes. The viscously dissipated energy is therefore advected into the black hole or other compact object at the center (see Narayan et al. 1998b and Kato et al. 1998 for reviews). In the limit where the fraction of the viscous energy advected inward is independent of radius, a self-similar analytic solution for the thermodynamic quantities of the accreting gas can be obtained (Narayan & Yi 1994, 1995b). We make use of this solution in Appendix A. For the numerical calculations presented in the rest of the paper, we use more accurate global solutions to obtain the run of electron temperature and density with radius. These solutions are calculated by the methods described in Narayan, Kato, & Honma (1997b), Chen, Abramowicz, & Lasota (1997), and Popham & Gammie (1998). The magnetic field strengths are obtained by assuming that the ratio of gas pressure to total pressure (sum of gas and magnetic pressure) is equal to a specified value $`\beta _{\mathrm{ADAF}}`$. Since the focus of this paper is on massive black holes in galactic nuclei with low mass-accretion rates, with specific applications to the black hole in our own Galactic nucleus, Sgr A, we scale masses in units of $`10^6`$ solar mass, i.e., $`Mm_610^6M_{},`$ and radii in units of the Schwarzschild radius, i.e., $`RrR_{Sch}`$, where $$R_{Sch}=\frac{2GM}{c^2}=2.95\times 10^{11}m_6\mathrm{cm}.$$ (1) We scale the mass accretion rate in units of $`10^3\dot{M}_{\mathrm{Edd}}`$, i.e., $`\dot{M}\dot{m}_310^3\dot{M}_{\mathrm{Edd}},`$ where the Eddington mass accretion rate is $$\dot{M}_{\mathrm{Edd}}=\frac{L_{\mathrm{Edd}}}{\eta _{\mathrm{eff}}c^2}=1.39\times 10^{24}\left(\frac{\eta _{\mathrm{eff}}}{0.1}\right)^1m_6\mathrm{g}\mathrm{s}^1.$$ (2) In defining the Eddington rate, we assume a standard radiative efficiency of $`\eta _{\mathrm{eff}}=0.1`$. (This is purely for the purposes of the definition; the actual radiative efficiency can be very different from $`0.1`$). For the calculations presented here we use ADAF models in which the viscosity parameter is set to $`\alpha =0.1`$, the equipartition parameter to $`\beta _{\mathrm{ADAF}}=0.968`$, and the ratio of viscous electron heating to proton heating to $`\delta =10^3`$. Note that $`\beta _{\mathrm{ADAF}}`$ differs from the usual plasma parameter $`\beta `$, which is the ratio of the gas pressure to the magnetic pressure; the value of $`\beta _{\mathrm{ADAF}}=0.968`$ assumed in our ADAF models corresponds to a plasma $`\beta =10`$ (Quataert & Narayan 1999). ### 2.2 Hybrid Populations We assume that a large fraction of the electrons in the plasma are in a thermal distribution with temperature $`T`$ and that the rest of the electrons are in a non-thermal distribution, usually with a power-law form. We denote the number density of electrons in the thermal population by $`N_{\mathrm{th}}`$ and in the non-thermal population by $`N_{\mathrm{pl}}`$. We denote the emissivities of the thermal and power-law electron populations by $`j_{\mathrm{th}}`$ and $`j_{\mathrm{pl}}`$ respectively, and the corresponding absorption coefficients by $`\alpha _{\mathrm{th}}`$ and $`\alpha _{\mathrm{pl}}`$. For the thermal electron population, we use the relativistic Maxwell-Boltzmann distribution given by $$n_{\mathrm{th}}(\gamma )=N_{\mathrm{th}}\gamma ^2\beta \mathrm{exp}(\gamma /\theta _\mathrm{e})/[\theta _\mathrm{e}\mathrm{K}_2(1/\theta _\mathrm{e})],$$ (3) where $`\gamma `$ is the electron Lorentz factor, $`\beta `$ is the relativistic electron velocity, and $`\theta _ekT/m_ec^2`$ is the dimensionless electron temperature. The modified Bessel function of second order $`K_2(1/\theta _e)`$ arises from the normalization of the Maxwellian. Similarly, for the non-thermal electron population we use a power-law distribution extending from $`\gamma =1`$ to infinity, $$n_{\mathrm{pl}}(\gamma )=N_{\mathrm{pl}}(p1)\gamma ^p.$$ (4) The number density of thermal electrons $`N_{\mathrm{th}}`$ is a function of the flow radius and is determined by the global ADAF solutions. We determine $`N_{\mathrm{pl}}`$ at each radius by assuming that the steady-state energy in the non-thermal distribution is equal to a fraction $`\eta `$ of the energy in the thermal distribution, with $`\eta `$ constant in radius. Although the calculations presented in $`\mathrm{\S }3`$ are all carried out with this assumption, generalizations to radially-dependent $`\eta (r)`$ as well as a discussion of the energetics of such a flow will be presented in $`\mathrm{\S }5`$. Note that we implicitly assume that the non-thermal electron population does not affect the dynamics or the thermal properties of the flow; this will be justified in $`\mathrm{\S }5`$. The energy density of a Maxwell-Boltzmann distribution of electrons at temperature $`\theta _e`$ was derived by Chandrasekhar $`(1939,\mathrm{eq}.[236])`$ to be $$u=a\left(\theta _e\right)N_{\mathrm{th}}m_ec^2\theta _e,$$ (5) where $$a\left(\theta _e\right)\frac{1}{\theta _e}\left[\frac{3K_3(1/\theta _e)+K_1(1/\theta _e)}{4K_2(1/\theta _e)}1\right]$$ (6) is a coefficient that varies from 3/2 for a non-relativistic electron gas to 3 for fully relativistic electrons, and $`K_n`$ are the modified Bessel functions of the n$`\mathrm{𝑡ℎ}`$ order. For the present purposes, we use a simplified expression for $`a(\theta _e)`$ which has an error of less than $`2\%`$ at all temperatures (Gammie & Popham 1998): $$a(\theta _e)=\frac{6+15\theta _e}{4+5\theta _e}.$$ (7) The number density of the non-thermal distribution is then $$N_{\mathrm{pl}}=\eta a(\theta _e)\theta _e(p2)N_{\mathrm{th}}.$$ (8) This normalization of the power law population typically corresponds to $$\frac{N_{\mathrm{pl}}}{N_{\mathrm{th}}}(0.110)\eta ,$$ (9) depending on the electron temperature and the power law index. The distributions considered above correspond to the steady state that results from the competition between heating/acceleration and cooling by radiation. We discuss in some detail in $`\mathrm{\S }5`$ the energy equations for the thermal and non-thermal electron populations. Here we simply note that the synchrotron cooling timescale of electrons moving with a Lorentz factor $`\gamma `$ scales as $`\gamma ^2`$, and hence electrons in the high energy tail of a power law distribution cool most rapidly. As a result, if electrons are injected with a power-law distribution with index $`s`$ $`[n(\gamma )\gamma ^s]`$ and cool only by synchrotron emission, the synchrotron cooling causes the power law index of the steady state distribution to be $`p=s+1`$ above a certain $`\gamma _b`$, called the cooling break, thus causing the steady state distribution to fall off more steeply at higher electron energies. Mahadevan & Quataert (1997) calculated $`\gamma _b`$ in an ADAF by comparing the cooling timescale to the inflow timescale and found that at sufficiently high accretion rates, the break occurs at a very low Lorentz factor, $`\gamma _b1.5`$. The Lorentz factors of interest to us are invariably larger than $`\gamma _b`$. Therefore, the values of $`p`$ we consider below in $`\mathrm{\S }3`$ and $`\mathrm{\S }4`$ are always equal to $`s+1`$, so that the injected energy distribution $`\gamma ^s`$ is harder by one power of $`\gamma `$ than the steady state energy distribution, $`\gamma ^p`$. Corresponding to $`\eta `$, we can define another quantity $`\eta _{\mathrm{inj}}`$ that measures the fraction of electron energy $`\mathrm{𝑖𝑛𝑗𝑒𝑐𝑡𝑒𝑑}`$ into a power law distribution. If $`s<2`$, then $`\eta _{\mathrm{inj}}`$ can be significantly greater than $`\eta `$. If we assume a distribution from $`\gamma _{\mathrm{min}}=1`$ to some $`\gamma _{\mathrm{max}}`$, $`\eta _{\mathrm{inj}}`$ is greater than $`\eta `$ by a factor $`\gamma _{\mathrm{max}}^{3p}`$. However, it is thought that the acceleration mechanisms typically encountered in astrophysics, such as shock acceleration or acceleration via MHD turbulence, inject energy into particles with $`s>2`$ such that the steady state distribution has $`p>3`$. In this case there is little dependence on $`\gamma _{\mathrm{max}}`$, and $$\frac{\eta _{\mathrm{inj}}}{\eta }\frac{p2}{p3},$$ (10) which is not very different from unity. Although we expect $`s>2`$ and $`p>3`$ for most systems, for completeness we consider models in the range $`2<p<4`$. ### 2.3 Synchrotron Emissivity The synchrotron emissivity of a relativistic electron moving with a Lorentz factor $`\gamma `$ in a magnetic field of strength $`B`$ is given by (Rybicki & Lightman 1979) $$j_\nu (\gamma ,\theta )=\frac{\sqrt{3}e^2}{2c}\nu _b\mathrm{sin}\theta F(x).$$ (11) Here, $`\nu _beB/2\pi m_ec`$ is the non-relativistic cyclotron frequency, $`\theta `$ is the angle between the direction of the magnetic field and the velocity of the electron, and $$F(x)x\underset{x}{\overset{\mathrm{}}{}}K_{5/3}(t)𝑑t,$$ (12) with $`K_{5/3}`$ the modified Bessel function of order 5/3, $`x\nu /\nu _c`$, and $`\nu _c\frac{3}{2}\gamma ^2\nu _b\mathrm{sin}\theta `$. For a thermal distribution of electrons, the total emissivity for a given angle $`\theta `$ is obtained by integrating equation (11) over the Maxwellian distribution (Pacholczyk 1970), $$j_{\nu ,\mathrm{th}}(\theta )=\frac{N_{\mathrm{th}}e^2}{\sqrt{3}cK_2(1/\theta _e)}\nu I\left(\frac{x_M}{\mathrm{sin}\theta }\right),$$ (13) where $$x_M\frac{2\nu }{3\nu _b\theta _e^2}$$ (14) and $$I(x_M)\frac{1}{x_M}\underset{0}{\overset{\mathrm{}}{}}z^2\mathrm{exp}(z)F(x_M/z^2)𝑑z.$$ (15) The limiting behaviour of $`I(x_M)`$ for small and large $`x_M`$ was derived by Pacholczyk $`(1970)`$ and Petrosian $`(1981)`$, respectively. Mahadevan, Narayan, & Yi (1996, hereafter MNY96) integrated equation (13) over the angle $`\theta `$ for an isotropic distribution of electron velocities and provided a fitting function for the emissivity in the ultrarelativistic to mildly relativistic regimes: $$j_{\nu ,\mathrm{th}}=\frac{N_{\mathrm{th}}e^2}{\sqrt{3}cK_2(1/\theta _e)}\nu M(x_M),$$ (16) with $`M(x_M)`$ given by $$M(x_M)=\frac{4.0505a}{x_M^{1/6}}\left(1+\frac{0.40b}{x_M^{1/4}}+\frac{0.5316c}{x_M^{1/2}}\right)\mathrm{exp}(1.8896x_M^{1/3}).$$ (17) The best fit values of the coefficients $`a,b`$, and $`c`$ for different temperatures are given in MNY96. The coefficients tend to unity in the ultrarelativistic limit. Finally, the synchrotron absorption coefficient $`\alpha _{\mathrm{th}}`$ is related to the emissivity via Kirchoff’s law, $$\alpha _{\nu ,\mathrm{th}}=j_{\nu ,\mathrm{th}}/B_\nu (T),$$ (18) where $`B_\nu (T)`$ is the black body source function. For the total emissivity of electrons in a power law distribution, we use the expression given in Rybicki & Lightman (1979) and average over angles, $$j_{\nu ,\mathrm{pl}}=C_{\mathrm{pl}}^j\eta \frac{e^2N_{\mathrm{th}}}{c}a(\theta _e)\theta _e\nu _b\left(\frac{\nu }{\nu _b}\right)^{(1p)/2},$$ (19) where $`N_{\mathrm{pl}}`$ is defined in terms of $`N_{\mathrm{th}}`$ as above and $$C_{\mathrm{pl}}^j=\frac{\sqrt{\pi }3^{p/2}}{4}\frac{(p1)(p2)}{(p+1)}\frac{\mathrm{\Gamma }(\frac{p}{4}+\frac{19}{12})\mathrm{\Gamma }(\frac{p}{4}\frac{1}{12})\mathrm{\Gamma }(\frac{p}{4}+\frac{5}{4})}{\mathrm{\Gamma }(\frac{p}{4}+\frac{7}{4})}.$$ (20) The corresponding absorption coefficient is $$\alpha _{\nu ,\mathrm{pl}}=C_{\mathrm{pl}}^\alpha \eta \frac{e^2N_{\mathrm{th}}}{c}a(\theta _e)\theta _e\left(\frac{\nu _b}{\nu }\right)^{(p+3)/2}\nu ^1,$$ (21) with $$C_{\mathrm{pl}}^\alpha =\frac{\sqrt{3\pi }3^{p/2}}{8}\frac{(p1)(p2)}{m_e}\frac{\mathrm{\Gamma }(\frac{3p+2}{12})\mathrm{\Gamma }(\frac{3p+22}{12})\mathrm{\Gamma }(\frac{6+p}{4})}{\mathrm{\Gamma }(\frac{8+p}{4})}.$$ (22) ### 2.4 Radiative Transfer and Numerical Methods The equation of radiative transfer for a time-independent, spherically symmetric flow is (e.g., Mihalas 1978) $$\mu (/r)+r^1(1\mu ^2)(/\mu )]I(r,\mu ,\nu )=j(r,\nu )\alpha (r,\nu )I(r,\mu ,\nu ),$$ (23) where $`\mu \mathrm{cos}\theta =(dz/ds)`$ is the cosine of the angle between the ray and the radial direction, $`r`$ is the radial coordinate, $`\nu `$ is the frequency, and $`j`$ and $`\alpha `$ are the emission and absorption coefficients defined above. One can simplify this equation by taking plane parallel rays of varying impact parameters (perpendicular distances of rays to the central line of sight) through the flow and solving the equation along these rays (Mihalas 1978). The equation then becomes $$\pm \frac{I_\nu ^\pm }{s}=j_\nu \alpha _\nu I_\nu ^\pm ,$$ (24) where now $`s`$ is the line element along the ray and the coefficients $`+1`$ and $`1`$ correspond to radiation coming towards and going away from an external observer, respectively. In our problem, $`j_\nu =j_{\mathrm{th}}+j_{\mathrm{pl}}`$ and $`\alpha _\nu =\alpha _{\mathrm{th}}+\alpha _{\mathrm{pl}}`$. Rewriting the equation in terms of the source function $`S_\nu =j_\nu /\alpha _\nu `$ and optical depth $`\tau (s)=_s^{s_{out}}\alpha 𝑑s^{}`$, where $`s_{out}`$ is the point of intersection of the ray with the outer boundary of the flow, equation (24) becomes $$\pm \frac{I_\nu }{\tau }=I_\nu S_\nu ,$$ (25) where the combined source function is $$S_\nu =\frac{j_\nu }{\alpha _\nu }=\frac{j_{\mathrm{th}}+j_{\mathrm{pl}}}{\alpha _{\mathrm{th}}+\alpha _{\mathrm{pl}}}=\frac{S_{\mathrm{th}}}{1+\alpha _{\mathrm{pl}}/\alpha _{\mathrm{th}}}+\frac{S_{\mathrm{pl}}}{1+\alpha _{\mathrm{th}}/\alpha _{\mathrm{pl}}}.$$ (26) In the numerical calculations reported below, we integrate equation (25) using the formal solution and the appropriate boundary conditions for a non-illuminated atmosphere (see Mihalas 1978). We carry out the integral along rays with impact parameters up to $`2000`$ Schwarzschild radii, beyond which the temperature of the electrons becomes too low for significant synchrotron emission. Because of the very steep dependence of the synchrotron emissivity on photon frequency, magnetic field strength, and electron temperature and density, we use an adaptive step size for the radiative transfer integral. The total flux is obtained by integrating over all impact parameters. We validated the implementation of our numerical algorithm by comparing its output to analytic solutions of the radiative transfer equation in uniform media. The discrepancy between the numerical and analytic solutions was $`0.5\%`$ for the cases considered. In all model ADAF spectra published so far, the transport of radiation has been calculated using an approximate method based on concentric shells (Narayan, Barret, & McClintock 1997a). In Figure 1, we compare, for a typical flow, the exact spectrum obtained by formally solving the above transfer equation to that obtained with the previous approximate method. We see that there is a fairly good agreement between the two methods, but with some differences. The most prominent difference is a shifting of the peak toward higher frequencies in the exact calculation, as well as some broadening. This is probably due to the poorer resolution of the concentric shell method which becomes a limiting factor close to the black hole. There is also a slight offset at lower frequencies, probably again due to poor resolution. Note that there is a second peak at high frequencies ($`10^{14}`$ Hz) in the spectrum calculated with the approximate method. This is caused by the inverse Compton scattering of soft photons, a process which is not included in the radiative transfer code described here. The remaining features of the two spectra are quantitatively consistent. ## 3 Numerical Results ### 3.1 Parameter Study In this section, we study the effects of an extended power-law electron distribution on the synchrotron emission spectra of ADAFs. Figure 2 shows the various components of the spectrum of a typical hybrid model, with $`p=3.5`$, $`\eta =1\%`$, $`m_6=2.5`$, and $`\dot{m}_3=1`$. Compared to the spectrum of a pure thermal model (dashed line), we see two primary effects due to the power-law electrons: (i) there is a prominent shoulder of optically thick emission at low frequencies and (ii) there is an extended power-law tail of optically thin emission at high frequencies. In between these two features there is a region of the spectrum where the thermal peak dominates. These features were first identified by Mahadevan (1998) for a specific model. We find that they are universal for hybrid models. To understand the results, we note that the spectra of hybrid populations are determined by a competition between $`j_{\mathrm{pl}}`$ and $`j_{\mathrm{th}}`$ and between $`\alpha _{\mathrm{pl}}`$ and $`\alpha _{\mathrm{th}}`$, each having different radial and frequency dependences. Since the emission at different frequencies arises from different radii in the flow, the relative importance of $`\alpha _{\mathrm{pl}}`$ to $`\alpha _{\mathrm{th}}`$ and $`j_{\mathrm{pl}}`$ to $`j_{\mathrm{th}}`$ at that particular frequency and radius determines the local behaviour of the spectrum. In the less steep segment of the low-frequency shoulder, the emission from the power-law population (which is more efficient at low frequencies) exceeds that of the thermal electrons, while absorption is still mostly dominated by the more numerous thermal electrons. As a result, this segment of the spectrum assumes a shape roughly described by the rather unusual source function $`S=j_{\mathrm{pl}}/\alpha _{\mathrm{th}}`$. At still lower frequencies, where the emission comes from larger radii in the flow, the contribution of the non-thermal electrons to the absorption becomes non-negligible and the spectrum falls more steeply with decreasing $`\nu `$. In the region of the thermal peak, we have both $`j_{\mathrm{th}}j_{\mathrm{pl}}`$ and $`\alpha _{\mathrm{th}}\alpha _{\mathrm{pl}}`$, with $`\alpha _{\mathrm{th}}R1`$ for frequencies smaller than the peak frequency and $`\alpha _{\mathrm{th}}R1`$ for frequencies higher than the peak frequency. The spectrum is essentially the same as for a purely thermal model (dashed line). (But, note that for sufficiently high $`\eta `$, the synchrotron spectrum assumes an entirely non-thermal character and this region too can be dominated by the power-law electrons, with the peak luminosity becoming higher and the peak flattened). Beyond the thermal peak, there is no self-absorption either by thermal or power-law electrons $`(\alpha _{\mathrm{th}}R,\alpha _{\mathrm{pl}}R1)`$ and the emission is optically thin. Here we find an extended power-law tail dominated entirely by the non-thermal population. This segment of the spectrum has the familiar form $`\nu L_\nu \nu ^{(p3)/2}`$ which gives a rising spectrum with increasing frequency for $`p<3`$ and a falling spectrum for $`p>3`$. Figures 3a-d show how the spectrum of a hybrid model depends on the various parameters. In Figure 3a, the energy content of the power-law population $`\eta `$ is varied from $`0\%`$ (purely thermal) to $`10\%`$; we see that the range and normalization of the low-frequency shoulder increase linearly with increasing $`\eta `$. The normalization of the high-frequency tail also depends linearly on the number of electrons in the power-law distribution and thus varies in an obvious way with $`\eta `$. In Figure 3b, we fix $`\eta `$ at $`1\%`$ and vary the power-law index $`p`$. Note that the overall luminosity of both the low-frequency shoulder and the high-frequency tail show considerable dependence on $`p`$. This is not unexpected. When $`p`$ is small, there are more particles both at intermediate Lorentz factors $`(\mathrm{log}\gamma =12)`$, which give rise to the low-frequency shoulder, as well as at large $`\gamma `$, which emit the high-frequency tail (see $`\mathrm{\S }4`$). Therefore, small values of $`p`$ give more emission in both these segments of the spectrum. The high-frequency tail has a spectral slope equal to $`(p3)/2`$ and therefore has a strong dependence on $`p`$. Two striking qualitative results in Figures 3a and 3b need to be highlighted. First, even very small values of the non-thermal energy content $`\eta `$ give rise to significant excess super-thermal emission at low frequencies. For instance, even if only $`1\%`$ of the total electron energy is in non-thermal electrons, there can be several orders of magnitude higher luminosity at frequencies below the thermal peak. Even more striking is the fact that the spectral shape of the low frequency shoulder is nearly independent of $`\eta `$ and $`p`$, when both these parameters are independent of radius, while the normalization depends on both (see Appendix A for an analytic expression for the spectral slope of this segment). This leads to a degeneracy between $`\eta `$ and $`p`$ in models of the low-frequency emission, so that it is not possible to distinguish between different combinations of $`\eta `$ and $`p`$, by studying spectral data below the thermal peak only. The degeneracy can be lifted with data on the high-frequency tail whose slope has a strong dependence on $`p`$. We finally consider the effect of changing the black hole mass and the mass-accretion rate. We find that the relative excess emission due to non-thermal electrons increases rather weakly with increasing black hole mass (Figure 3c). This is because the thermal peak moves to lower frequencies for higher $`m_6`$ (Mahadevan 1997) and at these frequencies the effect of non-thermal electrons is more prominent. The non-thermal contribution to both the low and high frequency emission increases only weakly with increasing $`\dot{m}_3`$ (Figure 3d). ### 3.2 Application to Sgr A Over the last decade, many models have been developed to explain the radio spectrum of Sgr A. This source is believed to be an accreting supermassive black hole at the center of our Galaxy. Nearly all the published models invoke synchrotron radiation from relativistic or quasi-relativistic electrons. Melia (1992) considered emission by thermal electrons in a spherical accretion flow and showed that the resulting cyclo-synchrotron emission is consistent with the broad features of the observed spectrum. Narayan, Yi, & Mahadevan (1995) and Narayan et al. (1998a) developed ADAF models of Sgr A which included rotation, viscosity and a two-temperature plasma, and obtained similar results, again with a purely thermal distribution of electrons. Beckert & Duschl (1997) and Falcke & Biermann (1999) considered non-thermal models, while Mahadevan (1998) analyzed a specific hybrid model in which the non-thermal electrons are produced by pion decay. This section is a generalization of Mahadevan’s work. In Figure 4 we apply our hybrid emission model to Sgr A. We take $`m_6=2.5`$ (Eckart & Genzel 1997; Ghez et al. 1998) and adjust $`\dot{m}_3`$ in order to fit approximately the thermal peak. The data shown are the same as in Narayan et al. (1998a). Figure 4 shows four models with $`p=2.5`$, $`\eta =0.05\%`$; $`p=3.0`$, $`\eta =0.2\%`$; $`p=3.5`$, $`\eta =0.5\%`$; and $`p=4.0`$, $`\eta =1\%`$. We first note that the agreement with data at low frequencies is significantly better with these hybrid models (dashed lines) than with a purely thermal model (solid line), as was first shown by Mahadevan (1998). The two major results pointed out in $`\mathrm{\S }3.1`$ are evident in this figure. First, there is no unique solution for the parameters $`p`$ and $`\eta `$. The four hybrid models shown in Figure 4 give indistinguishable spectra below the thermal peak, rendering it impossible to determine $`p`$ and $`\eta `$ from low-frequency spectral data alone. Second, $`\eta `$ is extremely small in all models. A very small fraction of the energy in non-thermal electrons, with $`\eta `$ at most $`1\%`$, is sufficient to produce all the observed emission at low frequencies. This means that the non-thermal electrons are a minor perturbation on the electron population, which is itself a minor perturbation on the more dominant ion population. We are therefore consistent when we compute the gas dynamics with a purely thermal model and ignore the non-thermal electrons for the dynamics. Further justification of this assumption as well as implications of this constraint are discussed in $`\mathrm{\S }5`$. We can obtain additional constraints on the parameters $`\eta `$ and $`p`$ in Sgr A by studying the infrared data. Sgr A is quiet at infrared wavelengths, with a current upper limit of $`10^{35}`$ erg $`\mathrm{s}^1`$ on the luminosity at 2.2 $`\mu `$m (Eckart & Genzel 1997; Ghez et al. 1998). Since $`\nu L_\nu `$ decreases with increasing frequency for $`p>3`$, we cannot use IR data to constrain $`\eta `$ very strongly in these cases. On the other hand, if $`p<3`$, electrons in an extended power-law distribution produce significant emission in the infrared. For example, if $`p=2.5`$, we find that the maximum allowed fraction of energy in non-thermal electrons is $`\eta =0.05\%`$. Tighter bounds on the infrared flux will constrain the parameters $`\eta `$ and $`p`$ even more strongly. We note that imposing a maximum Lorentz factor $`\gamma _{\mathrm{max}}`$ on the power-law electron distribution also has the effect of suppressing the high-frequency emission. Therefore, for $`p<3`$, we could alternatively use the IR data to constrain $`\gamma _{\mathrm{max}}`$ rather than $`\eta `$. The $`2.2\mu `$m emission is produced predominantly by electrons with $`\mathrm{log}\gamma 3`$ (see $`\mathrm{\S }4`$) placing a maximum Lorentz factor at $`\gamma _{\mathrm{max}}10^3`$, if $`p<3`$ and $`\eta >0.05\%`$. Finally, we have considered ADAF models with strong outflows following the ideas described in Blandford & Begelman (1999), Di Matteo et al. (1999), and Quataert & Narayan (1999). We find that the constraints on $`\eta `$ and $`\gamma _{\mathrm{max}}`$ obtained above do not depend strongly on the presence of winds. ### 3.3 Image sizes and shapes We now investigate the effect of a power-law electron population on the size and shape of the radio image of an ADAF. As a typical example, we consider an ADAF model of Sgr A with $`m_6=2.5`$ and $`\dot{m}_3=0.1`$ and take a hybrid electron distribution with $`p=3.5`$, $`\eta =0.5\%`$ which agrees well with the observed spectrum as shown in Figure 4. Preliminary size measurements of Sgr A are available at two frequencies (Lo et al. 1998, Krichbaum et al. 1998) and shape measurements may be possible in the near future. We note that although we use the parameters for a specific source, the qualitative results apply equally well to other massive black holes with different accretion rates. We define the image size as twice the radial distance from the center of the image to the point at which the specific intensity falls to half the value of the central intensity, i.e., the FWHM of the radio map. Figure 5 shows the predicted FWHM of the image as a function of frequency. The dashed curve corresponds to the case when all the electrons are thermal and the solid line to the case of a hybrid energy distribution. The two data points for Sgr A are from Lo et al. (1998) at 7 mm and Krichbaum et al. (1998) at 1.4 mm. Note that our radiative transfer code is one dimensional and can only handle spherical models, though ADAFs in general are oblate and are likely to appear elliptical in projection (Narayan & Yi 1995a). In the comparison with observations, we use only the measured long axes of the images. The effect of power-law electrons is to enlarge the image size at long wavelengths $`(\nu <10^{11}\mathrm{Hz})`$, where the excess non-thermal emission, which extends outward of the surface of unit optical depth, is most prominent. This is a general result which holds true for all model parameters we have studied. The other related result is the steepening of the dependence of the image size on the frequency at long wavelengths. While for thermal electrons we find $$\mathrm{FWHM}/\mathrm{R}_{\mathrm{Sch}}\lambda ^{0.7},$$ (27) when we include power law electrons, we find $$\mathrm{FWHM}/\mathrm{R}_{\mathrm{Sch}}\lambda ^{0.9}.$$ (28) This is in agreement with the preliminary result obtained by Lo et al. (1998) when they combined their measurement of the intrinsic size at 7 mm with the results of Krichbaum et al. (1998) at 1.4 mm, although the thermal and hybrid slopes are not distinguishable with the current data. We also calculate the effective brightness temperature $`T_b`$ at each frequency as $`T_b=L_\nu c^2/(8\pi ^2k_B\nu ^2\mathrm{FWHM}^2)`$. Figure 6 shows the result. For the thermal model, $`T_b`$ measures the electron temperature at the photosphere. Since the photosphere moves out with decreasing frequency, the brightness temperature falls. For the hybrid model, at lower frequencies, the contribution of the non-thermal electrons is large, the spectrum is non-thermal, and the resulting brightness temperature increases. This is another clear signature of a synchrotron-emitting hybrid electron population. We finally study the effect of the non-thermal electrons on the shape of the radio image. Figure 7 shows the variation of the brightness temperature $`T_b`$ across the source, for the thermal and hybrid models of Sgr A described above. We consider two wavelengths, 3.6 cm and 7 mm. The most striking feature of the hybrid case is the limb brightening seen at long wavelengths, a feature which is absent in the thermal models. Thus, if the accretion flow contains non-thermal electrons, its image would look like a shell rather than a disk. This is again due to the different radial dependences of absorption (which is predominantly thermal) and emission (which is mostly non-thermal) at the frequencies where the non-thermal shoulder appears in the energy spectrum. Because the total absorption falls off more steeply away from the center than the total emissivity, the image appears brighter for a range of impact parameters away from the center than at the center. In addition, due to the overall increase in the intensity of the emerging radiation in this same frequency range, the image looks brighter overall, with the brightness temperature $`T_b`$ of the hybrid case increasing to twice the thermal value in the center of the image at 3.6 cm. The limb brightening, along with the enhanced overall brightness at long wavelengths are signatures of a non-thermal population which may be accessible to observations. Unfortunately, in the case of Sgr A, precisely at the wavelengths where the effects are strongest, interstellar scattering blurs the observed image (e.g., Lo et al. 1998). It may be worthwhile to look for these effects in other sources where the scattering is less severe. ## 4 Electron Energy Distributions and the Shape of Synchrotron Spectra The numerical studies presented in $`\mathrm{\S }3`$ show that the presence of a small population of power-law electrons causes universal modifications to the spectrum: it introduces a shoulder at low frequencies and a power-law tail at high frequencies. We now attempt to associate each of these features of the hybrid spectrum with a specific range of Lorentz factors of the non-thermal electrons. In doing so, we address three important issues. The first is the question of degeneracy: why is the low frequency shoulder in the spectrum degenerate to a combination of the power-law index $`p`$ and the energy content $`\eta `$ of the non-thermal electrons? Understanding the source of non-uniqueness is especially important in trying to extract the underlying electron distributions from spectral data and determining what we can conclude $`\mathrm{𝑢𝑛𝑖𝑞𝑢𝑒𝑙𝑦}`$ about these distributions. Second, since introducing even a very small fraction of power-law electrons results in significant enhancement of the luminosity at low frequencies, existing radio data can be used to constrain the non-thermal energy fraction in these accretion flows. If only a part of the electron energy distribution is responsible for the excess emission, data can further constrain these specific parts of the particle distribution. The third question we address here is related to the absence of significant IR emission from the accretion flow around Sgr A (Eckart & Genzel 1997), in the frequency range where the extended optically-thin non-thermal emission is expected. Absence of emission at these wavelengths provides information about the underlying electron distributions by constraining either the power-law index $`p`$ or the maximum Lorentz factor $`\gamma _{max}`$ electrons can attain in these accretion flows. Determining $`\gamma _{max}`$ in a particle distribution may provide a better understanding of particle acceleration and cooling processes and timescales in hot accretion flows. ### 4.1 Correspondence between electron distributions and photon spectra We first investigate the relationship between the Lorentz factor $`\gamma `$ of an electron and the dominant frequency of the synchrotron emission it produces in the presence of a dominant thermal electron population. At very low frequencies, synchrotron emission comes from large radii in the flow where the temperatures are low, the non-thermal emission and absorption are efficient, and $`\alpha _{\mathrm{pl}}\alpha _{\mathrm{th}}`$. However, at higher frequencies near and above the low-frequency shoulder in the spectrum, $`\alpha _{\mathrm{th}}\alpha _{\mathrm{pl}}`$, as discussed earlier. For these frequencies, we can thus neglect $`\alpha _{\mathrm{pl}}`$. We make this key simplification only in this section. It simplifies the radiative transfer equation and allows us to compute separately the contribution to the intensity from each value of $`\gamma `$ in the non-thermal distribution. We proceed by writing all quantities as integrals over the power-law electron distribution, $$j_{\mathrm{pl}}=n_e^{\mathrm{pl}}(r,\gamma )j(\gamma )𝑑\gamma ,$$ (29) $$I(\nu )n_e^{\mathrm{pl}}(\gamma )I_\gamma (\gamma )𝑑\gamma ,$$ (30) and $$j_{\mathrm{th}}=\frac{j_{\mathrm{th}}n_e^{\mathrm{pl}}(\gamma )𝑑\gamma }{n_e^{\mathrm{pl}}(\gamma )𝑑\gamma },$$ (31) where $`n_e^{\mathrm{pl}}(r,\gamma )`$ is assumed to be a separable function of $`r`$ and $`\gamma `$ given by $$n_e^{\mathrm{pl}}(r,\gamma )=n_r^{\mathrm{pl}}(r)n_\gamma ^{\mathrm{pl}}(\gamma )=N_{\mathrm{pl}}(r)(p1)\gamma ^p.$$ (32) (Note that, for this analysis, the non-thermal distribution does not have to be a power-law, though we have assumed this in order to compare this analysis directly to our numerical results). Substituting equations (29)-(32) into the radiative transfer equation and rearranging terms, we obtain $$[\mu \frac{I_\gamma (\gamma )}{x}+\alpha _{\mathrm{th}}(r)I_\gamma (\gamma )n_r^{\mathrm{pl}}(r)j_\gamma (\gamma )j_{\mathrm{th}}]n_\gamma ^{\mathrm{pl}}(\gamma )𝑑\gamma =0.$$ (33) For this integral to vanish for any non-thermal electron distribution, the integrand must be identically zero, thus giving $$\mu \frac{I_\gamma (\gamma )}{x}=\alpha _{\mathrm{th}}(r)I_\gamma (\gamma )+n_r^{\mathrm{pl}}(r)j_\gamma (\gamma )+j_{\mathrm{th}}.$$ (34) We solve this equation for a wide range of Lorentz factors, $`\mathrm{log}\gamma =0.13.6`$, by choosing a specific radial profile of the non-thermal electron density $`N_{\mathrm{pl}}(r)`$ corresponding to $`p=3.5,\eta =0.5\%`$. We study the contribution of each Lorentz factor to the different parts of the spectra. The total intensity for the power-law distribution takes the form $$I(\nu )(p1)\gamma ^{p+1}I_\gamma (\gamma )d(\mathrm{log}\gamma ),$$ (35) and therefore, in order to study the true contribution of each $`\gamma `$ to the total spectrum, we plot the integrand $`\gamma ^{p+1}I_\gamma (\gamma )`$ as a function of frequency for each Lorentz factor. The result for $`\gamma =5,10,30,100`$, and 1000 (for $`p=3.5`$) are shown in Figure 8a. The curves give a good estimate of the individual contribution of each Lorentz factor to the total spectrum. Most of the contribution to the low-frequency shoulder comes from electrons with $`\gamma 3050`$. Emission from electrons with $`\gamma 100`$ is already lower by an order of magnitude at those frequencies, and the emission completely dies off beyond $`\gamma =100`$. Contribution to the high-frequency tail, on the other hand, starts around $`\gamma 100`$ and increases with increasing electron energy. To get a more quantitative idea of which range of Lorentz factors contributes to the three distinct regions of the spectrum, we show in Figure 8b the above integrand as a function of $`\gamma `$ for 3 frequencies, namely $`10^9`$ Hz (in the low-frequency shoulder), $`10^{12}`$ Hz (in the thermal peak), and $`10^{15}`$ Hz (in the high-frequency tail. We see that the emission at $`10^9`$ Hz is primarily from electrons with $`\gamma 10^{1.5}`$; the emission at $`10^{12}`$ Hz is mostly from $`\gamma 1`$ (thermal electrons, while the emission at $`10^{15}`$ Hz comes mostly from $`\gamma 10^2`$. To quantify this effect further, we plot in Figure 8c the minimum and maximum Lorentz factors for which the value of the integrand drops to half its maximum value for each frequency. This gives us exactly the contributing range of Lorentz factors to the emission at each frequency in the low-frequency shoulder and the high-frequency tail. We may summarize the results as follows. The low-frequency shoulder is caused by a narrow range of electron Lorentz factors, $`\mathrm{log}\gamma 12.`$ The narrowness of the range explains why the spectra are degenerate to different combinations of $`p`$ and $`\eta `$: a power-law distribution with small values of $`p`$ and small $`\eta `$ has nearly the same number of electrons in this narrow range of Lorentz factors as one with a larger $`p`$ and a larger $`\eta `$, thus producing the same emission in the shoulder. The high-frequency tail, however, is produced by electrons with a wide range of Lorentz factors $`(\mathrm{log}\gamma 2)`$ with higher frequencies coming from higher $`\gamma `$. Intermediate Lorentz factors $`\mathrm{log}\gamma 2`$ emit predominantly at frequencies around the thermal peak and are overpowered by the thermal emission, and therefore do not have an observable feature in the hybrid spectrum. ## 5 Energetics of the thermal and non-thermal populations In this section, we study the energy flow through the thermal and non-thermal electron populations as well as the energy exchange between these two populations through synchrotron self-absorption and Coulomb collisions. Starting with the energy equations for the two populations and a parametrization of the energy input into each, we calculate the steady-state energy content of both as a function of radius assuming that the shapes of the distributions are known a priori. Note that if the acceleration mechanism is known, the particle distributions can be calculated exactly (see, e.g., Nayakshin & Melia 1998). Here we simply assess the feasibility of a steady-state thermal/power-law hybrid distribution. We then extend our discussion of hybrid synchrotron spectra to cases where the parameter $`\eta `$ is not constant but is allowed to vary with radius. ### 5.1 Energy Equations We start by writing the energy equations for the two populations. Neglecting advection and diffusion which we estimate to be minor corrections, the energy balance for the power-law population reads $$\frac{E_{\mathrm{nt}}}{t}=0=\delta _{\mathrm{nt}}\dot{E}_{\mathrm{visc}}(r)j_{\mathrm{nt}}(\eta ,p,r)𝑑\nu 𝑑\mathrm{\Omega }\frac{E_{\mathrm{nt}}}{t_{ee}},$$ (36) while the thermal population obeys $$\frac{E_{\mathrm{th}}}{t}=0=\delta \dot{E}_{\mathrm{visc}}(r)\alpha _{\mathrm{th}}(B_\nu I_\nu )𝑑\nu 𝑑\mathrm{\Omega }+\frac{E_{\mathrm{nt}}}{t_{ee}},$$ (37) where $`E_{\mathrm{th}}`$ and $`E_{\mathrm{nt}}`$ are the steady state energy contents of the thermal and power-law populations respectively, $`\delta `$ is the fraction of viscous energy that heats the thermal electrons, and $`\delta _{\mathrm{nt}}`$ is the corresponding fraction injected into the power-law population ($`\delta _{\mathrm{nt}}=\eta _{\mathrm{inj}}\delta `$, where $`\eta _{\mathrm{inj}}`$ is defined in $`\mathrm{\S }2.2`$); the time derivatives are set to zero in steady-state. The energy exchange (thermalization) timescale due to Coulomb collisions of high energy electrons with the thermal bath is denoted by $`t_{ee}`$. The radiation term in equation (36) corresponds to the energy loss by the non-thermal electrons via optically thin synchrotron emission; because the non-thermal absorption coefficient is negligible throughout the flow, the non-thermal electrons do not gain energy by absorption of synchrotron radiation. The corresponding term in equation (37), on the other hand, describes the heating of the thermal electrons by the local radiation field that is in excess of the blackbody limit. When the radiation energy density is a blackbody, there is locally no heating or cooling, but as the radiation energy density is above this limit due to emission by the non-thermal electrons, the thermal electrons are heated by absorbing this total emission. Note that this is the same energy exchange mechanism invoked in the synchrotron boiler process (Ghisellini, Guilbert, & Svensson 1988; Ghisellini, Haardt, & Fabian 1993; Ghisellini, Haardt, & Svensson 1998). The transport of energy due to non-local synchrotron self-absorption is likely to be small and have little effect on the spectra as the photon mean free path becomes very short very rapidly along a radial path and thus the radiation emitted in the optically thick regions of the flow is reabsorbed locally. This is due to the very steep radial dependence of the synchrotron absorption coefficient. There are several issues we would like to address regarding the relative importance of thermal and non-thermal electrons in the flow and the energy exchange between the two populations. First, the quantity $`[\delta _{\mathrm{nt}}/\delta ](r)`$ determines as a function of radius the relative rate of heating of non-thermal electrons compared to thermal electrons. Second, the energy exchange between the populations (heating of the thermal electrons by non-thermal electrons) proceeds both via the Coulomb term and the absorption of the synchrotron photons emitted by the power-law populations (the second term of equation (37). Therefore the magnitude of these terms relative to the viscous heating of the thermal electrons needs to be assessed. Finally, the importance of the thermalization of the power-law distribution due to the Coulomb term also needs to be understood. We first estimate the relative magnitude of the terms in equation (36). The ratio of the energy loss of the power-law electrons due to Coulomb collisions (the last term) to the energy emitted in synchrotron photons (the second to last term) determines whether Coulomb collisions or synchrotron self-absorption is the dominant mechanism by which the non-thermal electrons cool. The hybrid spectra presented in this paper have been computed under the assumption that the power-law electrons lose energy primarily through synchrotron emission, and this needs to be checked for consistency. For a given electron velocity $`\beta `$ in the power-law distribution, the energy exchange rate with the thermal electrons is given by (Nayakshin & Melia 1998) $$\left(\frac{dE}{dt}\right)_{ee}\frac{3}{2}\frac{\mathrm{ln}\mathrm{\Lambda }}{t_T}\frac{1}{\beta \gamma _{\mathrm{av}}},$$ (38) where $`\mathrm{ln}\mathrm{\Lambda }20`$ is the Coulomb logarithm, $`\gamma _{\mathrm{av}}`$ is the average thermal Lorentz factor, $`t_T(n_ec\sigma _T)^1`$ is the Thomson mean-free time and $`n_e`$ the electron density. The rate of energy loss of an electron of Lorentz factor $`\gamma `$ emitting synchrotron radiation in a region of magnetic field B is (Rybicki & Lightman 1979) $$\left(\frac{E}{t_{1/2}}\right)_{\mathrm{syn}}\frac{2e^4B^2\gamma ^2}{3m^2c^3},$$ (39) where $`t_{1/2}`$ is the time for the electron to lose half its energy. Using the analytic expressions for the electron density and magnetic field strength given in $`\mathrm{\S }2.1`$ and substituting values for the coefficients and parameters appropriate for Sgr A, we find for an electron of a given $`\gamma `$: $$\frac{(dE/dt)_{ee}}{(E/t_{1/2})_{\mathrm{syn}}}0.1r\gamma ^2,$$ (40) where r is the radius in Scwarzschild units as usual. Most of the emission in the low-frequency shoulder in the spectrum originates from around $`r100`$ and the radiation is produced by electrons with $`\gamma 50`$ (cf. Fig. 8c and the Appendix). We thus see that the ratio in equation (40) is at most a few percent. This shows that Coulomb collisions play a negligible role in the cooling of non-thermal electrons. It also shows that that the heating of thermal electrons by Coulomb collisions is unimportant compared to energy exchange via synchrotron emission and absorption. We therefore neglect the Coulomb term in the following calculations. The power-law electrons then obey a simple energy balance between the injected energy through viscous dissipation and the energy loss through synchrotron emission. The latter simply is the total frequency- and angle-integrated synchrotron emissivity of the non-thermal electrons because this population is optically thin to synchrotron emission. Therefore, once $`\delta _{\mathrm{nt}}(r)`$ is specified, it is possible to calculate the parameter $`\eta `$, which describes the steady-state energy content of the power-law electrons, as a function of radius. Conversely, if we specify $`\eta (r)`$ as we did in the previous sections, it is possible to compute the energy $`\delta _{\mathrm{nt}}`$ that would need to be injected into the non-thermal electrons as a function of radius. We first study the latter case for $`\eta =`$ constant, as this is the assumption we have made in most of this paper. Figure 9 shows $`\delta _{\mathrm{nt}}/\delta `$ as a function of radius for one of the hybrid models of Sgr A\* with $`\eta =0.5\%`$, and $`p=3.5`$. The figure demonstrates that the energy input into the non-thermal electrons never exceeds $`20\%`$ of the heating of the thermal electrons and in fact does not exceed $`10\%`$ at those radii that contribute to the low-frequency shoulder in the spectrum $`(10R_s<R<1000R_s)`$. Thus, at no radius in the flow do the power-law electrons become energetically more important than, or even comparable to, the thermal electrons. This may seem like a surprising result considering that the non-thermal radiation clearly dominates over the thermal emission in the low-frequency shoulder of the spectrum and in the high-frequency tail. However, the synchrotron emissivities of the thermal and power-law populations are different by many orders of magnitude at these frequencies. (Optically thin synchrotron emission from a power-law population peaks at a much higher intensity and at a lower frequency than the thermal emission for the same energy density and magnetic field strength.) Therefore, even when the non-thermal population has less energy content than the thermal population, and even when a fraction of the non-thermal radiation is absorbed by the thermal electrons at frequencies where the flow is optically thick, the escaping radiation can still be dominated by the non-thermal emission and the intensity can be much above the blackbody limit. This is the situation in the low-frequency shoulder of the spectrum. In the high-frequency tail, the only particles that are energetic enough to produce the radiation are the non-thermal electrons, and since the flow is optically thin all the radiation escapes freely. Incidentally, the fact that the energy input into the non-thermal electrons does not exceed $`20\%`$ of the heating of the thermal electrons (and even this level is reached only over a small range of radius close to the black hole) shows that the dynamics of the flow is not affected by the presence of non-thermal electrons. We also consider the other mechanism of energy exchange between the two populations, namely synchrotron self-absorption by thermal electrons. The second term in equation (36) measures the maximum energy that can be radiated by the non-thermal electrons as a function of radius. This term also represents an upper limit on the energy that can be transferred to the thermal population via self-absorption. This term is bounded from above by $`\delta _{\mathrm{nt}}\dot{E}_{\mathrm{visc}}`$ according to equation (36). Since $`\delta _{\mathrm{nt}}/\delta `$ never exceeds $`20\%`$ as shown in figure 9, heating of the thermal electrons by absorbing synchrotron emission from non-thermal electrons can $`\mathrm{𝑎𝑡}\mathrm{𝑚𝑜𝑠𝑡}`$ introduce a $`20\%`$ correction to the thermal energy equation; the correction is only a few percent at large radius. This again justifies neglecting the contribution of the non-thermal electrons to the thermodynamic properties of the flow. Finally we note that the calculations in the previous sections considered a simple parametrization of the non-thermal population via the quantity $`\eta `$, which measures the fraction of the electron energy in steady state that is present in non-thermal electrons. Now, $`\eta `$ is a secondary quantity whose value depends on the balance between non-thermal heating and cooling. It would perhaps be more useful to parametrize the model by specifying the non-thermal heating parameter $`\delta _{\mathrm{nt}}`$. To this end, we consider models in which the energy injection rate varies as a power-law in radius, $`\delta _{\mathrm{nt}}r^q`$, and show results for $`q`$ between 0 and 1. We pick the normalization of $`\delta _{\mathrm{nt}}`$ at the inner edge of the flow such that the resulting $`\eta (r)`$ is comparable to the values used in $`\mathrm{\S }3`$ and is in accord with the results of Figure 9. In Figure 10a, we show $`\eta (r)`$ for $`p=3.5`$ and three different choices of q: 0, $`1/2`$, and $`3/4`$. For none of the three cases do we see very large variations in the resulting $`\eta (r)`$ versus radius; we find that $`\eta (r)`$ rises with radius for $`q1/2`$ and falls for larger values of $`q`$. Figure 10b shows the spectra corresponding to the three models of $`\delta _{\mathrm{nt}}`$. We see that the two universal features of a hybrid spectrum, the low-frequency shoulder and the high-energy tail, are present for all values of $`q`$. Thus, super-thermal emission is not an artifact of specific assumptions or parameter values of our models but is indeed a robust signature of a hybrid population. The slope of the low-frequency shoulder depends very mildly on the particular model of $`\delta (r)`$. ## 6 Discussion In this paper we considered hot accretion flows around supermassive black holes, using the ADAF model as a typical example. Such hot flows are expected to occur at low mass-accretion rates. We assume that a fraction of the viscous dissipation energy in the accretion flow goes into accelerating electrons to a non-thermal power-law distribution. We find that a power-law tail of high energy electrons can be sustained in such flows; neither Coulomb collisions nor synchrotron self-absorption is able to thermalize the power-law electrons. Assuming that there are no other thermalizing mechanisms (e.g., collective plasma modes), we calculate the resulting hybrid thermal/non-thermal spectrum from such a plasma. The presence of even a population of non-thermal electrons gives rise to two universal and prominent features in the synchrotron spectrum: a low-frequency shoulder and a high-frequency tail (Fig. 2). These features were identified by Mahadevan (1998) who considered a specific mechanism (via pion decay) for the production of power-law electrons. Even if only a small fraction of the total steady-state electron energy is in the non-thermal power-law component, we find that there is significant super-thermal emission in the low-frequency shoulder and the high frequency tail of the spectrum. Furthermore, each of these universal features can be associated with a specific range of Lorentz factors of the emitting electrons. The low-frequency shoulder is emitted by electrons with Lorentz factors $`\mathrm{log}\gamma =12`$, while the high-energy tail is emitted by electrons with $`\mathrm{log}\gamma 2`$ (Fig.8). Since the power-law electrons cause significant emission at low frequencies, comparing this low-frequency shoulder to data can provide stringent constraints on the fraction of the electron energy that is present in a non-thermal population. In the case of Sgr A (Fig. 4), we conclude that, in steady-state, electrons even with intermediate Lorentz factors ($`\gamma `$ as low as 10) make up only a very small fraction of the electron population. Since Coulomb collisions and synchrotron self-absorption are ineffective in thermalizing these flows, this means that either the energy injected into a non-thermal population, $`\delta _{\mathrm{nt}}\dot{E}_{\mathrm{visc}}`$, is small as in the calculations described in $`\mathrm{\S }5`$ or there is some other efficient thermalization mechanism which plays a role in these flows. The study on Sgr A needs to be extended to other sources before we can judge how universal the result is. Here it is important to emphasize two points regarding the constraints on non-thermal electrons obtained from the spectrum. First, for the part of the spectrum below the thermal peak, there is a mapping between each frequency in the spectrum and the thin shell of radii which predominantly emits at that frequency. This is due to the steep radial dependence of the synchrotron absorption that causes a sharp transition between optically thick and optically thin regions in the flow. Therefore, emission at each frequency comes primarily from a specific radius in the flow and can be used to probe the non-thermal content in a shell that extends from that radius out to a somewhat larger radius. (Mathematically, radiation emerging at each frequency comes from the entire range from the radius where the optical depth is unity out to infinity, but because the total emissivity also drops rapidly with radius, the biggest contribution to the intensity integral comes from a relatively narrow shell). As a result, the constraints on the non-thermal distribution obtained from each frequency of the low-frequency shoulder apply only to electrons within this shell. Inside the radius corresponding to optical depth unity, the emission is heavily self-absorbed and the electron distribution there is inaccessible to observations. Second, the constraints on the non-thermal energy content which we have obtained for Sgr A would not be invalidated if the observed low-frequency shoulder originated in a separate, extended region such as a jet or an outflow exterior to the accretion flow. In that case, the observed fluxes may be treated as upper limits to the emission from the inner accretion flow. This would imply that the non-thermal emission from the accretion flow is even smaller than in our models and the resulting constraints on the fraction of non-thermal electrons would become only tighter. When the non-thermal electron distribution is a power law, three parameters of the model and three corresponding pieces of physics can be extracted by comparing hybrid emission models to spectral data. The first parameter is $`\eta `$, or equivalently $`\delta _{\mathrm{nt}}`$ (cf.§5), which measures the fraction of the dissipated energy that goes into the non-thermal electrons. The second parameter is $`p`$, the slope of the power-law energy distribution (eq. 4). This can be uniquely determined by measuring the spectral slope of the high-frequency tail in the spectrum (provided there is no competing emission via Compton scattering at these wavelengths, cf. Fig. 1). The value of $`p`$ may provide an indication of the mechanism by which the electrons in a hot accretion flow are accelerated into a power-law tail. The third parameter, which is more significant for distributions with $`p<3`$, is $`\gamma _{\mathrm{max}}`$, the maximum Lorentz factor to which electrons are accelerated. If this parameter can be determined by an IR cutoff in the spectrum, it would provide information on the ratio of the acceleration timescale to the synchrotron cooling timescale. Due to the degeneracy of the low-frequency shoulder to the parameters $`\eta `$ and $`p`$ (Fig. 4), this segment of the spectrum by itself is not sufficient to constrain the two parameters. However, $`\eta `$ and $`p`$ may be determined uniquely if we have information on both the low-frequency shoulder and the high-frequency tail. For this to work, there should be no contamination from a jet or outflow to the observed radiation in the low-frequency shoulder or from Compton scattering in the high frequency tail. We showed in this paper that the presence of a non-thermal population of electrons also has measurable effects on the shape of the image of the source, and the size of the accretion flow, as a function of frequency (Figs. 5-7). Because the frequency dependence of the image size is strongly affected by the fraction of non-thermal electrons in the flow, measuring the size of accretion flows at multiple wavelengths can provide a quantitative constraint on the non-thermal electron energy content. Moreover, the brightness temperature of the source as a whole, as well as the variation of the brightness temperature across the source, at long wavelengths behave differently for a hybrid plasma compared to a purely thermal population of electrons. Finally, at the frequencies corresponding to the low-frequency shoulder, we show that there would be limb brightening of the image. All these image-related signatures may help to identify the presence of a non-thermal population of electrons, and the parameters of the corresponding energy distribution. We thank Eliot Quataert, Anthony Aguirre, Tiziana Di Matteo, and James Cho for many useful discussions and comments on the manuscript. We also thank S. Nayakshin, the referee, for many useful comments and suggestions that improved the presentation of the paper. D. P. was supported by a post-doctoral fellowship of the Smithsonian Institution. This work was supported in part by NSF Grant AST 9820686. ## Appendix A Analytic Approximations In this appendix, we provide analytic formulae for the slope of the spectrum below and above the thermal peak and the relative importance of non-thermal electrons as a function of mass, accretion rate, power law index $`p`$, and non-thermal energy content $`\eta `$. For convenience, we use the self-similar solution of ADAFs to obtain these analytic estimates. ### A.1 Self-Similar Advection-Dominated Flows We begin by presenting the self-similar ADAF solution developed by Narayan $`\&`$ Yi (1994, 1995b). The self-similar solution describes the local properties of the accretion flow as a function of the black hole mass $`M`$, the mass accretion rate $`\dot{M}`$, the radius $`R`$, the viscosity parameter $`\alpha `$, the ratio of gas pressure to the total pressure $`\beta `$, and the fraction of viscously dissipated energy that is advected inwards $`f`$. In terms of the scalings introduced in $`\mathrm{\S }2.1`$, the height-averaged electron number density $`n_e`$, the magnetic field strength B, and the dimensionless proton temperature $`\theta _pkT_p/m_pc^2`$ of the accretion flow are: $$n_e=n_1m_6^1\dot{m}_3r^{3/2}\mathrm{g}\mathrm{cm}^3,$$ (A1) $$B=b_1m_6^{1/2}\dot{m}_3^{1/2}r^{5/4}\mathrm{G},$$ (A2) and $$\theta _p=0.18\beta r^1,$$ (A3) where $$n_1=2.0\times 10^{10}\alpha ^1c_1^1c_3^{1/2},$$ (A4) and $$b_1=2.07\times 10^4\alpha ^{1/2}(1\beta )^{1/2}c_1^{1/2}c_3^{1/4}.$$ (A5) The coefficients $`c_1`$ and $`c_3`$ are defined in Narayan $`\&`$ Yi (1995b) and are related to the adiabatic index of the gas and the fraction of advected energy $`f`$. For the cases of interest here, $`c_10.5`$ and $`c_30.3.`$ The remaining parameters were specified in $`\mathrm{\S }2.1`$. ### A.2 Fundamental synchrotron quantities in self-similar ADAFs We now use the self-similar solution to obtain expressions for the cyclotron frequency and the harmonic number for synchrotron emission. Since we focus our applications to galactic nuclei and very large mass black holes where the synchrotron frequencies of interest are in the radio regime, we scale the frequency as $`\nu _{10}=\nu /10^{10}.`$ The magnetic field strength is then $$B=4.85\times 10^4m_6^{1/2}\dot{m}_3^{1/2}r^{5/4}$$ (A6) and thus fundamental cyclotron frequency defined in $`\mathrm{\S }2.3`$ becomes: $$\nu _b=1.4\times 10^{11}m_6^{1/2}\dot{m}_3^{1/2}r^{5/4}$$ (A7) The ion temperature retains its virial value throughout the flow and is given, in dimensionless units $`\theta _i=\frac{kT_i}{m_uc^2}`$, by: $$\theta _i=0.61\beta c_3r^1=0.2r^1$$ (A8) when we set the parameter values to those used in the simulations. The behaviour of the electron temperature on the other hand is more complicated and depends on the detailed balance of heating and cooling at each radius. Qualitatively, electron temperature starts out virial and equal to the ion temperature at large radii $`(r=10^5)`$, but increases less steeply than the ion temperature at smaller radii due to cooling by mainly synchrotron emission. Because it is not possible to derive the exact electron temperature as a function of radius analytically, we provide instead numerical fits to the average radial dependence of electron temperature $$\theta _e=1.65\times r^{0.6}.$$ (A9) Since synchrotron absorption coefficient is a steep function of radius, we can assume that most of the optically thick emission comes from a narrow range of radii around the radius with optical depth equal to unity. Due to the complicated nature of the expression for the thermal synchrotron absorption which does not allow analytic integration, we numerically determine the dependence of this $`\tau =1`$ radius, $`r_{t1}`$ on mass, mass accretion rate and frequency. The best fit gives $$r_{t1}=2.5\times 10^2m_6^{1/4}\dot{m}_3^{1/3}\nu _{10}^{0.6}$$ (A10) Substituting $`r_{t1}`$ into $`\nu _b`$ and $`\theta _e`$ we get $$\nu _b=1.4\times 10^8m_6^{3/16}\dot{m}_3^{1/12}\nu _{10}^{3/4}$$ (A11) and $$\theta _e=0.06\times m_6^{0.15}\dot{m}_3^{0.2}\nu _{10}^{0.36}$$ (A12) Finally, we compute the harmonic number defined in $`\mathrm{\S }2.3`$: $$x_M=\frac{2\nu }{3\nu _b\theta _e^2}$$ (A13) $$x_M=1.32\times 10^4m_6^{0.1}\dot{m}_3^{0.3}\nu _{10}^{0.47}$$ (A14) which provides all the necessary synchrotron expressions as a function of black hole mass, mass accretion rate and frequency. ### A.3 Where does the non-thermal emission dominate: Normalization We can now derive analytic estimates for the contribution of the non-thermal emission to the low-frequency shoulder and the high-frequency tail. In the optically thick shoulder, self-absorption of the synchrotron emission is done predominantly by the thermal electrons, and we find that absorption due to non-thermal electrons are negligible down to a frequency of $`\nu 10^9`$ as discussed above. Thus the source function will take the form $$S_{\mathrm{tot}}=\frac{j_{\mathrm{th}}}{\alpha _{\mathrm{th}}}+\frac{j_{\mathrm{pl}}}{\alpha _{\mathrm{th}}}$$ (A15) where $`j_{\mathrm{th}}`$ is synchrotron emissivity per unit volume given in $`\mathrm{\S }2.3`$. For the analytic approximations, we will use a simplified form of the thermal emissivity. We first take the fully relativistic limit derived first by Pacholczyk, which is equivalent to setting $`a,b`$ and $`c`$ equal to 1 in $`M(x_M)`$. We further neglect the second and third terms of the sum altogether by setting $`b`$ and $`c`$ equal to zero as these terms provide only a very small correction unimportant for the present purposes. Thus, in simplified form, the ratio of the two emissivities is given by $$\frac{j_{\mathrm{pl}}}{j_{\mathrm{th}}}=\eta C_{\mathrm{pl}}^{}a(\theta )K_2(1/\theta )\theta (\frac{\nu }{\nu _b})^{(p+1)/2}x_M^{1/6}\mathrm{exp}(1.9x_M^{1/3})$$ (A16) where $$C_{\mathrm{pl}}^{}=\frac{\sqrt{3}C_{\mathrm{pl}}}{4}$$ (A17) and all other quantities are as defined as before. The non-thermal emission dominates over the thermal emission in the spectrum when $`j_{\mathrm{pl}}/j_{\mathrm{th}}>1`$. This ratio can be easily computed using the expressions for $`x_M,\theta _e`$, and $`\nu _b`$ evaluated at the $`\tau =1`$ radius given above. The power law emission beyond the thermal peak is optically thin and hence the excess in this region depends sensitively on the power-law index as expected and is simply given by the optically thin synchrotron emission from the power-law electrons. The amount of emission is determined simply by $`j_{\mathrm{pl}}`$. We finally note that, for the small values of $`\eta `$ considered here, thermal emission dominates at the thermal peak. The strong dependence of the position and the normalization of the thermal peak on $`m`$ and $`\dot{m}`$ is discussed by Mahadevan (1997). ### A.4 Spectral Slope The shape of the low-frequency shoulder is dominated by the emission of the power-law population and absorption by the thermal electrons. In order to determine the spectral slope of this segment, we therefore first calculate the hybrid source function $$S_\nu =j_{\mathrm{pl}}/\alpha _{\mathrm{th}}$$ (A18) at the surface of unit optical depth. To convert this into a luminosity, we multiply the source function by the area of the $`\tau =1`$ surface. If we write the shoulder of the spectrum as $`\nu L_\nu \nu ^s`$, the spectral slope $`s`$ is then given by $$s=2.5p/87.2\dot{m}_3^{0.1}\nu _{10}^{0.16}+6m_6^{0.15}\dot{m}_3^{0.2}\nu _{10}^{0.36}$$ (A19) which gives the canonical value of $`4/3`$ for the mass and mass-accretion rate of Sgr A and scales (weakly) according to the expression above for other masses and accretion rates. One can immediately see that the spectral slope depends very weakly on p, only as $`p/8`$, thus demonstrating again the independence of the spectral slope of the low-frequency shoulder on the shape of the electron energy distribution. The slope of the optically thin power law emission is a well-known result, $$s=(p3)/2$$ (A20) which in $`\nu L_\nu `$ for $`p<3`$, flat for p=3 and falls off for $`p>3`$. It is independent of the black hole mass and the accretion rate.
warning/0004/nucl-th0004029.html
ar5iv
text
# In-Plane Elliptic Flow of Resonance Particles in Relativistic Heavy-Ion Collisions ## Abstract We analyze the second Fourier coefficient $`v_2`$ of the pion azimuthal distribution in non-central heavy-ion collisions in a relativistic hydrodynamic model. The exact treatment of the decay kinematics of resonances leads to almost vanishing azimuthal anisotropy of pions near the midrapidity, while the matter elliptic flow is in-plane at freeze-out. In addition, we reproduce the rapidity dependence of $`v_2`$ for pions measured in non-central Pb + Pb collisions at 158$`A`$ GeV. This suggests that resonance particles as well as stable particles constitute the in-plane flow and are important ingredients for the understanding of the observed pion flow. The main goals of relativistic nuclear collisions are to determine the nuclear equation of state (EOS) under extreme conditions and to understand a new phase of deconfined nuclear matter, the quark-gluon plasma (QGP) . Since it is the pressure gradient perpendicular to the collision axis that causes various transverse collective flows, such as radial flow, directed flow, and elliptic flow, in relativistic nuclear collisions, these flows observed in the final state are expected to carry the information about the EOS. If the QGP phase is created in nuclear collisions, the quark matter expands, cools down, and goes through the “softest point” where the ratio of the pressure to the energy density takes its minimum as a function of the energy density. Therefore, it is expected that the suppression of the collective flows is not only a signal for the existence of the QGP but also a useful tool to determine the EOS near the phase transition region. Some experimental groups reported that the inverse slope parameters of the transverse mass spectra for the non-multistrange hadrons $`\pi `$, $`K`$, $`p`$, and $`d`$ in central Pb + Pb collisions at 158$`A`$ GeV at the CERN Super Proton Synchrotron (SPS) are parametrized by two common values, the freeze-out temperature and the transverse flow velocity . This implies that these particles constitute the radial flow and the local thermalization at least among those particles is achieved in central collisions. It has been, however, an open question whether equilibration is achieved even in non-central collisions. In this Letter, we study the elliptic flow that is made of stable particles and resonances with a hydrodynamic model, and show that the final pion distribution is well described in this approach, if decay kinematics is appropriately taken into account. The rapidity dependence of azimuthal anisotropy for particle $`i`$ is characterized by the coefficients $`v_n^i`$ $`(n=1,2,3\mathrm{})`$ in the Fourier expansion of the azimuthal distribution of the particle : $`{\displaystyle \frac{dN^i}{p_Tdp_TdYd\varphi }}={\displaystyle \frac{1}{2\pi }}{\displaystyle \frac{dN^i}{p_Tdp_TdY}}[1+2v_1^i(Y,p_T)\mathrm{cos}\varphi +2v_2^i(Y,p_T)\mathrm{cos}2\varphi +\mathrm{}],`$ (1) where $`\varphi `$ is the azimuthal angle measured from the reaction plane and $`Y`$ is the rapidity. Non-vanishing $`v_1`$ and $`v_2`$ imply the formation of directed and elliptic flows, respectively. Experimentally, $`v_2(Y)`$ is of the order of four percent around the midrapidity ($`Y=2.92`$) for low transverse momentum ($`50<p_T<`$ 350 MeV/$`c`$) charged pions in non-central Pb + Pb collisions that correspond to the impact parameter range $`6.5<b<8.0`$ fm . The value of the observed $`v_2`$ has been claimed to be smaller than that predicted for direct pions in hydrodynamic models. In the following, we consider not only direct pions but also indirect pions that are from resonance decays. Thus, we need to strictly distinguish the matter flow before freeze-out and the observed particle flow. In this Letter we use a term “in-plane” when the hydrodynamic flow before freeze-out is directed preferentially to the positive and the negative $`x`$-axes on the transverse plane and a term “positive elliptic flow” when $`v_2`$ of observed particles is positive. Here the $`x`$-axis is defined as the direction of impact parameter in non-central collisions. When one only considers particles directly emitted from freeze-out hyper-surface, “positive elliptic flow” means that the hydrodynamic flow is “in-plane”. However, once feeding from resonance decays is included, the above two are no longer equivalent. First, let us suppose that a resonance particle with mass $`m_R`$ is emitted from freeze-out hyper-surface and decays into two identical daughter particles with mass $`m`$ in the vacuum. In the rest frame of the resonance particle the decay is isotropic and the daughter particles are uniformly distributed with respect to the azimuthal angle when one averages the spin of the resonance particles. This is, however, not the case anymore if the resonance is moving in a reference frame. To see this more clearly, let us assume that the resonance particle moves with velocity $`(V_{Rx},V_{Ry},V_{Rz})=(V,0,0)`$ in a reference frame. If $`V`$ is larger than the critical value $`V^{}=p^{}/E^{}`$, where $`p^{}=\sqrt{m_R^2/4m^2}`$ and $`E^{}=m_R/2`$, the probability that the daughter particle is emitted with an angle $`\varphi `$ from the $`x`$-axis peaks at $`\varphi _\pm =\pm \mathrm{sin}^1[p^{}(1V^2)^{1/2}/(mV)]`$ and $`\varphi `$ is limited in a range, $`\varphi _{}\varphi \varphi _+`$ . The two peaks are due to the Jacobian singularity in the Lorentz transformation from the resonance rest frame to the reference frame. As a result, the opening angle between the daughter particles tends to remain finite even if the resonance particle moves at a large velocity. This is the reason why the equivalence between “in-plane flow” and “positive elliptic flow” breaks down when one includes the decay of resonances in the final state. The multiplicity of pions through two body decays of resonances is given by $`N_{R\pi X}`$ $`=`$ $`{\displaystyle \frac{J(p_L,\varphi ;𝐕_R)dp_Ld\varphi }{4\pi p^{}}B_{R\pi X}\frac{d^3𝐩_R}{E_R}𝑑sW_R(s)\frac{d_R}{(2\pi )^3}\frac{p_R^\mu d\sigma _\mu }{\mathrm{exp}[(p_R^\nu u_\nu \mu )/T_f]1}},`$ (2) where $`u^\nu `$, $`\mu `$, $`T_f`$, and $`\sigma _\mu `$ are, respectively, the four-dimensional fluid velocity, the chemical potential, the freeze-out temperature, and the freeze-out hyper-surface. $`p_L`$ is the longitudinal momentum of pion. $`p_R^\nu `$ and $`d_R`$ are, respectively, the resonance four momentum in the reference frame and the degeneracy, and $`(+)`$ is for boson (fermion) resonances. $`B_{R\pi X}`$ is the branching ratio of the decay process and $`W_R`$ is the Breit-Wigner type function, which takes account of the finiteness of the resonance width. For $`W_R`$, we adopt the form used in Ref. . The Jacobian of the Lorentz transformation from the resonance rest frame to an arbitrary reference frame $`J(p_L,\varphi ;𝐕_R)`$ is defined by $`dp_L^{}d\varphi ^{}=J(p_L,\varphi ;𝐕_R)dp_Ld\varphi `$, where the quantities with (without) $``$ are the ones in the resonance rest (reference) frame. Two typical shapes of the Jacobian as functions of the azimuthal angle in $`\rho \pi \pi `$ are shown in Fig. 1 (a). If $`𝐕_\rho =(V_\rho ,0,0)`$ and $`V_\rho <V^{}(0.93c\mathrm{in}\mathrm{this}\mathrm{process})`$, $`J(p_L=0,\varphi ;𝐕_\rho )`$ has a broad peak around $`\varphi =0`$ at the same azimuthal angle of the resonance particle as expected. However, if $`V_\rho >V^{}`$, $`J(p_L=0,\varphi ;𝐕_\rho )`$ has a finite value only in the range $`\varphi _{}<\varphi <\varphi _+`$ and two sharp peaks appear at $`\varphi _\pm `$ in addition to the original broad one as explained above. In order to get an idea about the effect of thermal smearing, we estimate the azimuthal distribution of pions through the above process, taking the following simple model: There are only two fluid elements with local fluid velocities $`(v_x,v_y,v_z)=(\pm 0.5c,0,0)`$. Both elements are assumed to be thermalized at the temperature $`T_f=120`$ MeV. The pion distribution from $`\rho `$-meson decays in each fluid element is shown in Fig. 1 (b). Here we sum up the pion distribution over $`50<p_T<350`$ MeV and $`0.5<Y<0.5`$. The sharp peaks at $`\varphi \pm 1.21`$ ($`\pm 1.93`$) in the Jacobian are smeared by thermal motion of $`\rho `$-mesons, but they are still visible in the azimuthal distribution of pions from $`\rho `$-mesons in the fluid element with $`v_x=0.5c`$ ($`0.5c`$), while the peaks at $`\varphi =0`$ ($`\pi `$) are completely washed out. The superposition of the two distributions leads to an azimuthal distribution with broad peaks at $`\pm \pi /2`$, as shown in Fig. 1 (b). In this simple model, pions from $`\rho `$-meson decay have negative $`v_2`$, i.e., the elliptic flow in the final state is negative, while the motion of the two fluid elements is in-plane. We next carry out realistic hydrodynamic simulations for the space-time evolution of Pb + Pb collisions at 158$`A`$ GeV to see how large the effect of the Jacobian singularity is for the final state pion distribution. We first assume that the hot and dense nuclear matter produced in heavy-ion collisions is in local thermal equilibrium after $`t_0=1.44`$ fm/$`c`$ since the two nuclei touched . Then, we describe the space-time evolution of nuclear matter after this time by using a relativistic hydrodynamic model without assuming the Bjorken’s scaling solution or the cylindrical symmetry along the collision axis . Thus, it is possible to discuss the rapidity dependence of elliptic flow $`v_2(Y)`$ for charged pions through resonance decays as well as for charged pions directly emitted from the freeze-out hyper-surface in this model. We use a model EOS with a first order phase transition between the QGP phase and the hadron phase . The QGP phase is assumed to be free gas composed of quarks with $`N_f=3`$ and gluons. For the hadron phase we adopt a resonance gas model, which includes all baryons and mesons up to the mass of 2 GeV , together with an excluded volume correction . We use the critical temperature at zero baryon density, $`T_c(n_B=0)=160`$ MeV. The two model EOS’s are matched by imposing the Gibbs’ condition for phase equilibrium on the phase boundary. The numerical results of the hydrodynamic simulation give us the momentum distribution of hadrons through the Cooper-Frye formula with a freeze-out energy density $`E_\mathrm{f}=60`$ MeV/fm<sup>3</sup>. We have used this formula for the direct emission of $`\pi ^{}`$, $`K^{}`$, and $`p`$. In addition, we have taken into account negative pions from the decays of $`\rho `$, $`\omega `$, $`K^{}`$, and $`\mathrm{\Delta }`$ in the final state. In the numerical simulation we have fixed the impact parameter at 7.2 fm. We have chosen the initial parameters in the hydrodynamic simulation to reproduce not only the rapidity and transverse mass distribution of negative hadrons in central collisions but also the preliminary data of rapidity distribution of negative pion in non-central collisions by the NA49 collaboration. The corresponding central energy density $`E_0`$ and baryon number density $`n_{B0}`$ are, respectively, 3.9 GeV/fm<sup>3</sup> and 0.46 fm<sup>-3</sup>. The energy density and the baryon number density on the transverse plane are assumed to be in proportion to the number of wounded nucleons with the standard Woods-Saxon distribution for the nuclear density. We have assumed the initial transverse flow to vanish. For details on our hydrodynamic model, see Ref. . We first discuss the effect of the Jacobian singularity for the pions through $`\rho `$-meson decays in the realistic hydrodynamic calculation. The momentum distribution of $`\rho `$-mesons is free from the effects of the Jacobi function and is given by the second integral in Eq. (2), i.e., the integral with respect to $`s`$ and $`\sigma _\mu `$. In Fig. 2, we show $`v_2`$ for the $`\rho `$-mesons directly emitted from freeze-out hyper-surface and for the pions through $`\rho \pi \pi `$. The elliptic flow of $`\rho `$-mesons is positive and in-plane. Nevertheless, the $`v_2`$ for the pions through $`\rho \pi \pi `$ almost vanishes near the midrapidity due to the decay kinematics. This implies that the effect of the Jacobian singularity, which we discussed above with a simple model, survives even in this realistic calculation. The behavior of the $`v_2`$’s of pions from $`K^{}`$ or $`\mathrm{\Delta }`$ is similar to this. Finally, we discuss the rapidity dependence of the observed pion elliptic flow obtained from our fully three-dimensional hydrodynamic calculation. In Fig. 3 we compare our results with the experimental data by the NA49 Collaboration . The solid line represents $`v_2`$ for the total charged pions. For comparison, $`v_2`$ for pions directly emitted from freeze-out hyper-surface (dashed line) and that for pions through resonance decays (dotted line) are shown separately. Our results were obtained by summing up pion distribution over a $`p_T`$ range, $`50<p_T<350`$ MeV/$`c`$. The experimental data corresponds to an impact parameter range $`6.5<b<8.0`$ fm. The solid line is in good agreement with the experimental data near midrapidity . The $`v_2`$ for indirect pions from resonance decays vanishes. This reduces $`v_2`$ for the total pions by about 26 % at midrapidity. This figure tells us that hydrodynamic description, which assumes the local thermal equilibrium, works well also for the expansion stage of non-central collisions at the SPS energy . A few remarks on other calculations are in order here. At midrapidity, our result for the total pions is consistent with the previous result obtained by a two-dimensional hydrodynamic model . The authors of Ref. assumed that the longitudinal expansion can be described by the Bjorken’s scaling solution and numerically simulated the evolution of nuclear matter only in the transverse directions. Hence they could not obtain the rapidity dependence of elliptic flow. They also took into account resonance particles up to the mass of the $`\mathrm{\Delta }`$(1232), but concluded that resonance decays reduce the momentum anisotropy for pions by only 10-15 %. Their reduction factor corresponds to full $`p_T`$ range . The effect of Jacobian singularity is important in low $`p_T`$ region. When we integrate the distribution over full $`p_T`$ range, we obtain the result reduced by 11 % . Liu et al. compared their results from a transport model, the Relativistic Quantum Molecular Dynamics (RQMD), with the experimental data and concluded that the model calculations are in reasonable agreement with experimental data. The experimental data of elliptic flow was, however, later updated , and the agreement is not as good as before anymore. Soff et al. also obtained $`v_2(Y)`$ for pions from a microscopic transport model, the Ultrarelativistic Quantum Molecular Dynamics (UrQMD), but the situation is similar to the RQMD model. Recently, it was argued that the suppression of the elliptic flow is due to partial thermalization . However, as we have shown, this is not needed to explain the data. According to our result, the suppression rather implies the full thermalization of the system, including resonances. In summary, we have investigated the elliptic flow of pions in Pb + Pb non-central collisions at 158$`A`$ GeV in a relativistic hydrodynamic model. As sources of pions, we have considered not only direct pion emission from the freeze-out hyper-surface but also decays of resonance particles. Pions from resonance decays suppress the azimuthal anisotropy in the midrapidity region as much as 26 %. Taking this effect into account, we were able to reproduce the experimental data of the rapidity dependence of $`v_2`$. These results lead to our conclusion that the pions and the resonance particles constitute thermalized in-plane elliptic flow and that the hydrodynamic picture is applicable to the expansion stage in the non-central collisions at the SPS energy. The Jacobian singularity effect should exist also in cascade calculations such as RQMD and UrQMD if the decay kinematics is appropriately taken into account. It will be interesting to see how important this mechanism is for the suppression in those calculations. The author is much indebted to Professor I. Ohba and Professor H. Nakazato for their helpful comments. He wishes to acknowledge valuable discussions with M. Asakawa, T. Hatsuda, P. Huovinen, P. F. Kolb, T. Matsui, Y. Miake, S. Muroya, S. Nishimura, C. Nonaka, A. Ohnishi, J.-Y. Ollitrault, and A. M. Poskanzer. He also thanks M. Asakawa for a careful reading of the manuscript and C. Nonaka for giving him a numerical table of EOS.
warning/0004/cond-mat0004239.html
ar5iv
text
# The electronic structure of CuSiO3 – a possible candidate for a new inorganic spin-Peierls compound ? \[ ## Abstract Electronic structure calculations are presented for the well-known CuGeO<sub>3</sub> and the recently discovered isostructural CuSiO<sub>3</sub> compounds. The magnitude of the dispersion in chain direction is considerably smaller for CuSiO<sub>3</sub>, whereas the main interchain couplings are rather similar in both compounds. Starting from extended one-band tight-binding models fitted to the bandstructures, the exchange integrals were estimated for both compounds in terms of a spatially anisotropic Heisenberg model. Remarkable frustrating second neighbor couplings are found both for intra- and inter-chain interactions. A magnetic moment of about 0.35 $`\mu _B`$ is predicted for CuSiO<sub>3</sub> in the Néel state. \] Low-dimensional spin systems such as chains or ladders are of fundamental interest for contemporary solid state physics due to their peculiar electronic and magnetic properties. During the last years, many related materials have been found within the cuprate family, famous for the high temperature superconductivity. All cuprates contain CuO<sub>4</sub> plaquettes. In most cases it is energetically favorable to connect these plaquettes by the formation of chains or planes. According to the number ($`n=1,2`$) of oxygen atoms shared by adjacent plaquettes, these compounds can be classified as so-called edge-shared ($`n=2`$) or corner-shared ($`n=1`$) compounds. Obviously, the type of sharing affects strongly the physical properties of the compounds under consideration. For example, corner sharing leads to strong antiferromagnetic coupling between neighboring plaquettes compared with the weak inter-chain interactions. As a result, the straight CuO<sub>3</sub> chain in Sr<sub>2</sub>CuO<sub>3</sub> is the best known realization of the one-dimensional spin-1/2 Heisenberg model, with an in-chain exchange coupling of about 2200 K, but with a Néel temperature of only 5 K and with an extremely small ordered magnetic moment of about 0.06 $`\mu _\text{B}`$, both due to a small residual interchain exchange coupling. Spin-charge separation in the excitation spectra could be observed for Sr<sub>2</sub>CuO<sub>3</sub> and for the double chain compound SrCuO<sub>2</sub>. Somewhat surprisingly, in contrast to the similarity between different corner-shared chain compounds, the magnetic properties in the edge-shared chain family exhibit a remarkable variance. Thus, the edge-shared CuO<sub>2</sub> plaquettes in Li<sub>2</sub>CuO<sub>2</sub> order antiferromagnetically with a ferromagnetic arrangement along those chains and with a large ordered moment of 0.9 $`\mu _\text{B}`$, whereas the same chain in CuGeO<sub>3</sub> shows a spin-Peierls transition at low temperatures. Antiferromagnetically ordered chains were observed in Cu<sub>1-x</sub>Zn<sub>x</sub>GeO<sub>3</sub> for small concentrations of Zn impurities. It is noteworthy that, even for the intensively studied CuGeO<sub>3</sub>, a consensus with respect to the quantitative description of competing or complementary interactions such as the inter-chain coupling, frustration and spin-phonon coupling has not been reached so FIG. 1.: The orthorhombic unit cell of the CuSiO<sub>3</sub>-crystal, perspective view (top), front view (down left) and top view (down right). The edge-shared cuprate-chains run along the c direction and are canted against each other. far, despite the achieved qualitative understanding of their influence on different magnetically ordered states. Naturally, the magnetic properties depend very sensitively on the electronic interactions in these systems. Therefore, a comparative study of the electronic properties of closely related systems can shed light on the interactions responsible for the magnetically ordered states mentioned above. In this context, the recent discovery and first investigations of the long searched for compound CuSiO<sub>3</sub>, which is isostructural to the prototypical inorganic spin-Peierls system CuGeO<sub>3</sub> is of great scientific interest. The crystal structure of CuSiO<sub>3</sub> is shown in Fig. 1. The most important feature for the magnetic properties are the planar edge-shared CuO<sub>2</sub> chains running along c-direction. These chains are very similar to those of CuGeO<sub>3</sub>. The Cu-O(2) bond length in CuSiO<sub>3</sub> (CuGeO<sub>3</sub>) is 1.941 Å (1.942 Å), the Cu-O(2)-Cu bonding angle is 94 (99). Thus, the question arises, whether the very recently observed phase transition near 8 K does point to a new inorganic spin-Peierls system or to another ordered state realized at low temperature. To get theoretical insight into possible scenarios, we present here comparative band-structure calculations and tight-binding examinations for CuSiO<sub>3</sub> and CuGeO<sub>3</sub>. In this context we note that for the latter compound several (non full-potential) bandstructure calculation have been reported (e.g. in Ref. ), but to our knowledge the inter-chain interaction has not been analyzed in detail. The relevant electronic structure of these materials is very sensitive to details of hybridization and charge balance. In order to obtain a realistic and reliable hopping part of a tight binding Hamiltonian, band-structure calculations were performed using the full-potential nonorthogonal local-orbital minimum-basis scheme within the local density approximation (LDA). In the scalar relativistic calculations we used the exchange and correlation potential of Perdew and Zunger. Cu($`4s`$, $`4p`$, 3$`d`$), O(2$`s`$, 2$`p`$, 3$`d`$), Ge(3$`d`$, 4$`s`$, 4$`p`$, 4$`d`$) and Si(2$`p`$, 3$`s`$, 3$`p`$, 3$`d`$) states, respectively, were chosen as minimum basis set. All lower lying states were treated as core states. The inclusion of Ge 3$`d`$ and Si 2$`p`$ states in the valence states was necessary to account for non-negligible core-core overlaps. The O and Si 3$`d`$ as well as the Ge 4$`d`$ states were taken into account to increase the completeness of the basis set. The spatial extension of the basis orbitals, controlled by a confining potential $`(r/r_0)^4`$, was optimized to minimize the total energy. The results of the paramagnetic calculation for CuSiO<sub>3</sub> (see Fig. 2 (a)) and CuGeO<sub>3</sub> (see Fig. 2 (b); we find similar results as the non full-potential calculation of Ref. ) show a valence band complex of about 10 eV width with two bands crossing the Fermi level in both cases. These two bands are well separated from the rest of the valence band complex and show mainly Cu 3$`d`$ and O(2) 2$`p`$ character in the analysis of the corresponding partial densities of states (not shown). We note that the occupancy of the two O(2) 2$`p`$ orbitals along and perpendicular to the chain (lying in the plaquette-planes) is rather different, but it is almost identical for the corresponding orbitals in both compounds. Therein, we found only a small admixture of O(1) 2$`p`$ and Ge 4$`s`$ and 4$`p`$ states, respectively, with a total amount of few percent. The examination of the eigenstates of the latter bands at high symmetry points yields an antibonding character typical for cuprates. Here these relatively narrow antibonding bands are half-filled. Therefore, strong correlation effects can be expected which explain the experimentally observed insulating groundstate. Despite almost perfect qualitative one to one correspondence of all valence bands and main peak structures in the densities of states (DOS) (compare right panels in Fig. 2), the most important differences between both compounds FIG. 2.: Band structure and total density of states for CuSiO<sub>3</sub> (a), CuGeO<sub>3</sub> (b), and the zoomed antibonding bands (c) (CuSiO<sub>3</sub> full lines, CuGeO<sub>3</sub> dashed lines). The Fermi level is at zero energy. The notation of the symmetry points is as follows: Y = (010), T = (011), Z = (001), X = (100), S = (110), A = (111). The chain direction corresponds to Y–T, Z–$`\mathrm{\Gamma }`$ and S–A. occur for the antibonding bands (shown in detail in Fig. 2(c)). Therefore, we restrict ourselves to the extended tight-binding analysis and the discussion of these antibonding bands. The dispersion of these bands has been analyzed in terms of nearest neighbor transfer (NN), next nearest neighbor transfer (NNN) and higher neighbor terms in chain direction, but only NN hopping and a diagonal transition term between the CuO<sub>2</sub>-chains have been considered (see Fig. 3). Then, the corresponding dispersion relation takes the form FIG. 3.: Schematical chain and stack arrangement of CuO<sub>2</sub>-plaquettes, respectively, and considered transfer processes within the bc-plane (left panel) and in the ab-plane (right panel). $`E(\stackrel{}{k})`$ $`=`$ $`2({\displaystyle \underset{m=1,4}{}}t_{mz}\mathrm{cos}(mz)+\mathrm{cos}(x)[t_x+2t_{xz}\mathrm{cos}(z)]`$ (2) $`+\mathrm{cos}(y/2)[t_y+2t_{yz}\mathrm{cos}(z)+2t_{xy}\mathrm{cos}(x)]),`$ where $`x=k_za`$, $`y=k_yb`$, $`z=k_zc`$. Notice that in our effective one-band description the upper band (see Fig. 2(c)) e.g. along $`\mathrm{\Gamma }`$–X corresponds to $`k_y`$ = 0, whereas the lower one corresponds to $`k_y`$ = $`2\pi /b`$. The assignment of the parameters has been achieved by two numerically independent procedures: By straightforward least square fitting of the whole antibonding band in all directions and by using the bandwidths, the slopes and the curvatures at special selected high symmetry points. The latter procedure has the advantage to be less affected by hybridization effects from lower lying bands near the bottom of the antibonding band (being of some relevance near the Z-point in Fig. 2). The results are shown in Tab. I. The errors can be estimated between 1% for the large and 10% for the small parameters from the difference of both mentioned above fitting procedures. The analyzed antibonding bands of both compounds exhibit a rather similar shape except near the Z-points, where the hybridization with lower lying bands produces an additional band-crossing for CuGeO<sub>3</sub> (see Fig. 2(c)). Recall that the main difference to the corner-shared chains as e.g. in Sr<sub>2</sub>CuO<sub>3</sub> is a much smaller in-chain NN transfer due to the different geometry. In spite of the qualitative similarity, the calculated values for the transfer integrals are quite different. The in-chain dispersion is nearly twice as large for CuGeO<sub>3</sub> in comparison to CuSiO<sub>3</sub>. This can be attributed mainly to the larger Cu-O-Cu bond angle in CuGeO<sub>3</sub> (99 and 94, respectively). However, this geometrical effect is somewhat reduced by the different on-site energies of the oxygen orbitals along and perpendicular to the chain (lying in the plaquettes planes). The latter difference is reflected by the larger separation of the corresponding bands at the Z-point in CuSiO<sub>3</sub> (see Fig. 2). The inter-chain dispersions in $`b`$ direction are comparable. For both compounds, we find also rather significant diagonal hopping terms $`t_{yz}`$ which are reflected by different dispersions along the X–S and the T–Z directions. Somewhat surprisingly, we found a sizeable dispersion in $`x`$-direction for CuGeO<sub>3</sub> but only a very weak one for the CuSiO<sub>3</sub> counterpart. From the transfer integrals discussed above, we conclude that both compounds are not so well-defined quasi one-dimensional systems as compared to the corner-shared CuO<sub>3</sub> chain compounds . The inter-chain coupling is rather significant for CuGeO<sub>3</sub>, and CuSiO<sub>3</sub> can even be regarded as an anisotropic two-dimensional system. Since increasing inter-chain coupling tends to destabilize the spin-Peierls state, a Néel ordered antiferromagnetic ground state might be expected for CuSiO<sub>3</sub> in contrast to the spin-Peierls state realized in CuGeO<sub>3</sub>. The obtained transfer integrals enables us to estimate the relevant exchange integrals $`J`$. This knowledge is crucial for the derivation and examination of magnetic model Hamiltonians of the spin-1/2 Heisenberg type frequently used in the literature: $$H_{spin}=\underset{ij}{}^{}J_{ij}\stackrel{}{S_i}\stackrel{}{S_j}.$$ (3) In general, the total exchange $`J`$ can be divided into an antiferromagnetic and a ferromagnetic contribution $`J`$ = $`J^{AFM}+J^{FM}`$. In the strongly correlated limit, valid for typical cuprates, the former can be calculated in terms of the one-band extended Hubbard model $`J_{ij}^{AFM}`$ = $`4t_{ij}^2/(UV_{ij})`$. The indices $`i`$ and $`j`$ correspond to nearest and next nearest neighbors, $`U`$ is the on-site Coulomb repulsion and $`V_{ij}`$ is the inter-site Coulomb interaction. From experimental data mapped from the standard $`pd`$-model onto the one-band description, one estimates $`UV`$ $``$ 4.2 eV. For the sake of simplicity, we neglect the difference in the quantity $`UV`$ in the compounds. The calculated values for the exchange integrals are given in Tab. II. The value of the NN exchange integral $`J_1^{AFM}`$ $``$ 30 meV in CuGeO<sub>3</sub> exceeds the experimental values of about 11 meV from inelastic neutron scattering data , about 14 meV from magnetic susceptibility and about 22 meV from Raman scattering . This points to a significant ferromagnetic contribution due to the Goodenough-Kanamori-Anderson-type interaction. In the following, we shall adopt 15 meV for the resulting total exchange coupling $`J_1`$ as a representative value, suggested by the average of the above mentioned experimental data. Owing to the lack of experimental data we assume the same ratio $`J_1/J_1^{AFM}`$ in CuSiO<sub>3</sub> as in CuGeO<sub>3</sub>, suggested by the quite similar O(2) 2$`p`$ orbital occupancies mentioned above. For the latter compound, we note the reasonable agreement with the available experimental data and most of our calculated antiferromagnetic values for the remaining exchange parameters. Hence, further possible ferromagnetic contributions seem to be less relevant and are neglected in the following considerations. Further simplification can be obtained mapping J<sub>1</sub> and the frustrated NN term J<sub>2</sub> onto an effective intra-chain coupling $`J_{}=J_11.12J_2`$. The calculated values for $`J_{}`$ are 12.2 meV for CuGeO<sub>3</sub> and 2.8 meV for CuSiO<sub>3</sub>, respectively. The latter value is close to the value of 2 meV reported by Baenitz et al. from a one-dimensional fit of magnetic susceptibility data. We find also a considerable inter-chain frustration $`J_{yz}=\beta J_y`$ with $`\beta `$=0.36 (0.34) for the Ge- (Si-) compound. This is in good agreement with the suggestions of Uhrig $`\beta 0.5`$ for CuGeO<sub>3</sub>. Transfering the above mentioned idea to map frustrating terms onto one effective coupling, we adopt $`J_{}=J_y2J_{yz}`$ for the effective inter-chain exchange parameters in $`b`$-direction. The factor of two is introduced to account approximately the twice as large number of second neighbors. The effective anisotropy ratio $`R=J_{}/J_{}`$ measures approximately the magnitude of quantum fluctuations. In the crossover region between one and two dimensions, quantum fluctuations do strongly affect the magnitude of the staggered magnetization $`m`$ and the local Cu moment $`\mu =g_Ln_dm`$ at $`T=0`$ for a Néel ground state, where $`g_L`$=2.06 to 2.26 denotes the (anisotropic) Landé-factor (tensor) for Cu<sup>2+</sup> in CuGeO<sub>3</sub> and $`n_d`$ 0.8 is the hole occupation number of the related Cu 3$`d`$ plaquette orbital. Using the expression $$m=0.39\sqrt{R}(1+0.095R)\mathrm{ln}^{1/3}(1.3/R),$$ (4) taken from Ref. , we arrive at 0.17$`\mu _B`$ in reasonable agreement with the neutron data 0.22$`\pm 0.02`$ and 0.2 for the disorder induced Néel state achieved below 4.5K in Zn-doped CuGeO<sub>3</sub>. The same approach predicts a significantly larger value of about 0.35$`\mu _B`$ for CuSiO<sub>3</sub> realized in a possible Néel state. To summarize, our LDA-FPLO calculation reveals valuable insight into the relevant couplings of CuGeO<sub>3</sub> and CuSiO<sub>3</sub>. We can classify CuGeO<sub>3</sub> as a quasi one-dimensional compound with significant inter-chain interaction, whereas CuSiO<sub>3</sub> is closer to an anisotropic two-dimensional compound. The significantly reduced energy scale of the in-chain exchange interactions and the large inter-chain interaction in CuSiO<sub>3</sub> are less favorable for a spin-Peierls state than for a Néel order. However, due to the large frustrations other states such as a spin-Peierls state cannot be excluded. Further investigations are required to elucidate the unknown ground state. We acknowledge fruitful discussions with M. Baenitz, C. Geibel, W. Pickett and G. Uhrig. This work was supported by individual grants of the DAAD (H.R.) and the DFG (S.D.).
warning/0004/hep-ph0004177.html
ar5iv
text
# Weyl Invariant Formulation of Flux-Tube Solution in the Dual Ginzburg-Landau Theory ## I Introduction Recent studies in lattice QCD in the maximally Abelian gauge suggest remarkable properties of the QCD vacuum, such as Abelian dominance and monopole condensation, which provide the dual superconductor picture of the QCD vacuum as is described by the dual Ginzburg-Landau (DGL) theory . The DGL theory is obtained by using the Abelian projection. In this scheme, QCD is reduced into the $`[U(1)]^2`$ gauge theory including color-magnetic monopoles. Based on the dual superconductor picture of the QCD vacuum, we get an intuitive picture of hadrons as the vortex excitation of the color-electric flux, which we call the color-electric flux-tube, or simply the flux-tube. In this vacuum, the color-electric flux is squeezed into an almost one dimensional object like a string due to the dual Meissner effect caused by monopole condensation. This situation seems to be the same with the appearance of the Abrikosov vortex in the ordinary superconductor system, which is caused by the Cooper pair condensation. We know that Abrikosov-Nielsen-Olesen (ANO) vortex in the ordinary superconductor can be described by using the Abelian Higgs theory, where the keyword is the breaking of $`U(1)_e`$ gauge symmetry through the Higgs mechanism. Moreover, there exists an analytic solution of the ANO vortex in the border of the type-I and the type-II vacuum, so called the Bogomol’nyi limit. The analytical solution exhibits interesting features of the superconductivity and is useful to understand the properties of the vortex dynamics. Hence, it is considered quite interesting to investigate the flux-tube solution in the dual superconductor QCD vacuum corresponding to the ANO vortex in the Bogomol’nyi limit. However, the symmetry in the QCD vacuum is not so simple compared with the ordinary superconductor system, since now we have to take into account the $`[U(1)]_m^2`$ dual gauge symmetry corresponding to the $`U(1)_e`$ gauge symmetry in the ordinary superconductor. Note that the symmetry $`[U(1)]^{N1}`$ is originated from the maximal torus subgroup of $`SU(N)`$. Furthermore, we also have the Weyl symmetry, which is the permutation invariance of the labels among the Abelian color charges. Therefore, the flux-tube in the QCD vacuum is expected to have some characteristic aspects beyond the analogue of the ANO vortex in the ordinary superconductor system. In this paper, we investigate the flux-tube solution in the DGL theory in the Bogomol’nyi limit. This study seems to be similar as is given in Ref. . In fact, our result will be shown identical. However, we would like to present an usuful method to find the Bogomol’nyi limit, and this can be achieved by taking into account the Weyl symmetry in the DGL theory. This idea can be extended straightforwardly to the $`[U(1)]^{N1}`$ dual Abelian Higgs theory which would be reduced from the $`SU(N)`$ gluodynamics. We first write the DGL Lagrangian in a manifestly Weyl invariant form. At the same time, we pay attention to the singular structure in the DGL theory, since it plays a significant role to obtain the string-like flux-tube solution. Note that the boundary condition of the dual gauge field depends crucially on this singular structure. Second, we consider the Bogomol’nyi limit, the border between the type-I and the type-II vacuum. The string tension in this limit is computed analytically. Finally, we discuss the properties of the flux-tube solution in the DGL theory. ## II Manifestly Weyl invariant form of the DGL Lagrangian The DGL Lagrangian is given by Throughout this paper, we use the following notations: Latin indices i,j express the labels 1,2,3, which is not to be summed over unless explicitly stated. Boldface letter denotes three-vector. $`_{\mathrm{DGL}}`$ $`=`$ $`{\displaystyle \frac{1}{4}}\left(_\mu \stackrel{}{B}_\nu _\nu \stackrel{}{B}_\mu {\displaystyle \frac{1}{n}}\epsilon _{\mu \nu \alpha \beta }n^\alpha \stackrel{}{j}^\beta \right)^2`$ (2) $`+{\displaystyle \underset{i=1}{\overset{3}{}}}\left[\left|\left(_\mu +ig\stackrel{}{ϵ}_i\stackrel{}{B}_\mu \right)\chi _i\right|^2\lambda \left(\left|\chi _i\right|^2v^2\right)^2\right],`$ where $`\stackrel{}{B}_\mu `$ and $`\chi _i`$ denote the dual gauge field with two components $`(B_\mu ^3,B_\mu ^8)`$ and the complex scalar monopole field, respectively. The quark field is included in the current $`\stackrel{}{j}_\mu =e\overline{q}\gamma _\mu \stackrel{}{H}q`$, $`\stackrel{}{H}=(T_3,T_8)`$. Here, $`\stackrel{}{ϵ}_i`$ is the root vector of $`SU(3)`$ algebra, $`\stackrel{}{ϵ}_1=(1/2,\sqrt{3}/2),\stackrel{}{ϵ}_2=(1/2,\sqrt{3}/2),\stackrel{}{ϵ}_3=(1,0)`$, and $`n^\mu `$ denotes an arbitrary constant 4-vectorIf the dual gauge symmetry is broken through monopole condensation, $`n^\mu `$ can not be an arbitrary vector any more. Instead, this vector describes the dynamics of the string and gives the contribution to the energy of the system., which corresponds to the direction of the Dirac string. The gauge coupling $`e`$ and the dual gauge coupling $`g`$ hold the relation $`eg=4\pi `$. This relation guarantees the unobservability of the Dirac string when the dual gauge symmetry is not broken. Note that the DGL Lagrangian (2) is invariant under the $`[U(1)]_m^2`$ dual gauge transformation, $`\chi _i\chi _ie^{if_i},\chi _i^{}\chi _i^{}e^{if_i},`$ (3) $`\stackrel{}{B}_\mu =(B_\mu ^3,B_\mu ^8)(B_\mu ^3{\displaystyle \frac{1}{g}}_\mu f_3,B_\mu ^8{\displaystyle \frac{1}{\sqrt{3}g}}(_\mu f_1_\mu f_2)),(i=1,2,3),`$ (4) where the phase $`f_i`$ has the constraint $`_{i=1}^3f_i=0`$. The non-local term, in the kinetic term of the dual gauge field, is concretely written as $$\frac{1}{n}\epsilon _{\mu \nu \alpha \beta }n^\alpha \stackrel{}{j}^\beta =d^4x^{}x|\frac{1}{n}|x^{}\epsilon _{\mu \nu \alpha \beta }n^\alpha \stackrel{}{j}^\beta (x^{}),$$ (5) where $`x|{\displaystyle \frac{1}{n}}|x^{}`$ $`=`$ $`\left[p\theta ((xx^{})n)(1p)\theta ((x^{}x)n)\right]\delta ^{(3)}(\stackrel{}{x}_{}\stackrel{}{x}_{}^{}).`$ (6) Here $`p`$ is an arbitrary real number and $`\delta ^{(3)}(x)`$ is the $`\delta `$-function defined on a three dimensional hyper-surface which has the normal vector $`n_\mu `$, so that $`\stackrel{}{x}_{}`$ and $`\stackrel{}{x}_{}^{}`$ are 3-vectors (not necessarily spatial) which are perpendicular to the $`n_\mu `$. It is noted that in order to define the color-electric charge of the quark in terms of the dual gauge field, we need such a non-local term, which is a result of the choice of one potential approach. Now, we define an extended dual gauge field to take into account the Weyl invariance in the DGL theory as $`B_{i\mu }\sqrt{{\displaystyle \frac{2}{3}}}\stackrel{}{ϵ}_i\stackrel{}{B}_\mu ,(i=1,2,3)`$ (7) Here, the constraint $`_{i=1}^3B_{i\mu }=0`$ appears, which has the same structure with the constraint $`_{i=1}^3f_i=0`$. Furthermore, we divide the dual gauge field into two parts, the regular part and the singular part, $`\stackrel{}{B}_\mu \stackrel{}{B}_\mu ^{\mathrm{reg}}+\stackrel{}{B}_\mu ^{\mathrm{sing}}.`$ (8) The factor $`\sqrt{2/3}`$ in (7) is a simple normalization to get the factor $`1/4`$ in front of the kinetic term of the dual gauge field (See (10)). The singular dual gauge field $`\stackrel{}{B}_\mu ^{\mathrm{sing}}`$ is determined so as to cancel the Dirac string in the non-local term as $`_\mu \stackrel{}{B}_\nu ^{\mathrm{sing}}_\nu \stackrel{}{B}_\mu ^{\mathrm{sing}}{\displaystyle \frac{1}{n}}\epsilon _{\mu \nu \alpha \beta }n^\alpha \stackrel{}{j}^\beta \stackrel{}{C}_{\mu \nu }.`$ (9) In the static $`q`$-$`\overline{q}`$ system, $`\stackrel{}{C}_{\mu \nu }`$ is nothing but the color-electric field originated from the color-electric charge like the electric field induced by an electric charge, where an explicit form of $`\stackrel{}{B}_\mu ^{\mathrm{sing}}`$ is given in Sec. III. It is noted that the cross term of the regular dual field tensor $`{}_{}{}^{}\stackrel{}{F}_{\mu \nu }^{\mathrm{reg}}_\mu \stackrel{}{B}_\nu ^{\mathrm{reg}}_\nu \stackrel{}{B}_\mu ^{\mathrm{reg}}`$ and $`\stackrel{}{C}_{\mu \nu }`$ can be integrated out, and the square of $`\stackrel{}{C}_{\mu \nu }`$ and its integration gives the Coulomb energy including the self-energy of the color-electric charge. However, we drop it hearafter in order to concentrate on the flux-tube itself. Correspondingly, we pay attention to the string tension for an ideal flux-tube system which has terminals at infinity<sup>§</sup><sup>§</sup>§In order to classify the types of the flux-tube, we use the word such as the $`q`$-$`\overline{q}`$ system.. Then, we obtain $`_{\mathrm{DGL}}={\displaystyle \underset{i=1}{\overset{3}{}}}\left[{\displaystyle \frac{1}{4}}{}_{}{}^{}F_{i\mu \nu }^{\mathrm{reg}}{}_{}{}^{2}+\left|\left(_\mu +ig^{}\left(B_{i\mu }^{\mathrm{reg}}+B_{i\mu }^{\mathrm{sing}}\right)\right)\chi _i\right|^2\lambda \left(\left|\chi _i\right|^2v^2\right)^2\right],`$ (10) $`{}_{}{}^{}F_{i\mu \nu }^{\mathrm{reg}}_\mu B_{i\nu }^{\mathrm{reg}}_\nu B_{i\mu }^{\mathrm{reg}},`$ (11) where the dual gauge coupling $`g`$ is scaled as $$g^{}\sqrt{\frac{3}{2}}g.$$ (12) One finds that the dual gauge symmetry becomes very easy to observe, since the dual gauge transformation is defined by $`\chi _i\chi _ie^{if_i},\chi _i^{}\chi _i^{}e^{if_i},`$ (13) $`B_{i\mu }^{\mathrm{reg}}B_{i\mu }^{\mathrm{reg}}{\displaystyle \frac{1}{g^{}}}_\mu f_i,(i=1,2,3)`$ (14) and accordingly the Lagrangian (10) has the extended local symmetry $`[U(1)]_m^3`$. However, it does not mean an increase of the gauge degrees of freedom because we have the constraint $`_{i=1}^3B_{i\mu }=0`$. The field equations are given by $`\left(_\mu +ig^{}\left(B_{i\mu }^{\mathrm{reg}}+B_{i\mu }^{\mathrm{sing}}\right)\right)^2\chi _i=2\lambda \chi _i(\chi _i^{}\chi _iv^2),`$ (15) $`^\nu {}_{}{}^{}F_{i\mu \nu }^{\mathrm{reg}}k_{i\mu }=ig^{}\left(\chi _i^{}_\mu \chi _i\chi _i_\mu \chi _i^{}\right)+2g^2\left(B_{i\mu }^{\mathrm{reg}}+B_{i\mu }^{\mathrm{sing}}\right)\chi _i^{}\chi _i,`$ (16) These field equations are to be solved with the proper boundary conditions that quantize the color-electric flux. The flux is given by $$\mathrm{\Phi }_i{}_{}{}^{}F_{i\mu \nu }^{\mathrm{reg}}𝑑\sigma ^{\mu \nu }=B_{i\mu }^{\mathrm{reg}}𝑑x^\mu ,$$ (17) where $`\sigma ^{\mu \nu }`$ is a two-dimensional surface element in the Minkowski space. By using the polar decomposition of the monopole field as $`\chi _i=\varphi _ie^{i\eta _i}`$ $`(\varphi _i,\eta _i\mathrm{})`$, we get, from the field equation (16), $`B_{i\mu }^{\mathrm{reg}}={\displaystyle \frac{k_{i\mu }}{2g_{}^{}{}_{}{}^{2}\varphi _i^2}}B_{i\mu }^{\mathrm{sing}}{\displaystyle \frac{1}{g^{}}}_\mu \eta _i.`$ (18) We substitute this expression into (17) and integrate out over a large closed loop where the monopole current $`k_{i\mu }`$ is vanished. Thus we get $$\mathrm{\Phi }_i=\left(B_{i\mu }^{\mathrm{sing}}+\frac{1}{g^{}}_\mu \eta _i\right)𝑑x^\mu .$$ (19) It is suggested from this expression that there are two possibilities to obtain the flux-tube configuration. One is originated from the singularity in $`B_{i\mu }^{\mathrm{sing}}`$ and the other is from the singularity in $`_\mu \eta _i`$. We find that the former case, as can be seen from the relation (9), corresponds to the flux-tube which has the quark source. On the other hands, the latter case, it does not contain any information of the quark, which means no terminal, hence, it cannot provide the physical state like a $`q`$-$`\overline{q}`$ system. If one assumes the existence of the external color-electric source or the glueball state as the flux-tube ring, it should be taken into account. However, since this is not the case which we discuss in this paper, we assume that there is no singularity in $`_\mu \eta _i`$. Then, this term can be absorbed into the regular dual gauge field $`B_{i\mu }^{\mathrm{reg}}`$ by the replacement $`B_{i\mu }^{\mathrm{reg}}+_\mu \eta _i/g^{}B_{i\mu }^{\mathrm{reg}}`$. In this case, the flux (19) just has the meaning of the boundary condition of the regular dual gauge field which should behave as $`B_{i\mu }^{\mathrm{reg}}B_{i\mu }^{\mathrm{sing}}`$ at infinity, where monopoles are condensed. ## III The static $`q`$-$`\overline{q}`$ system In this section, we consider the static $`q`$-$`\overline{q}`$ system. The quark source is given by the $`c`$-number current, which is typical in the heavy quark system, $$\stackrel{}{j}^\mu \stackrel{}{j}_{}^{\mu }{}_{j}{}^{}(x)=\stackrel{}{Q}_jg^{\mu 0}\left[\delta ^{(3)}\left(𝒙𝒂\right)\delta ^{(3)}\left(𝒙𝒃\right)\right],$$ (20) where $`\stackrel{}{Q}_je\stackrel{}{w}_j`$ is the Abelian color-electric charge of the quark. Here, $`𝒂`$ and $`𝒃`$ are position vectors of the quark and the antiquark, respectively, and $`\stackrel{}{w}_j`$ is the weight vector of $`SU(3)`$ algebra, $`\stackrel{}{w}_1=(1/2,\sqrt{3}/6),\stackrel{}{w}_2=(1/2,\sqrt{3}/6),\stackrel{}{w}_3=(0,1/\sqrt{3})`$. This vector is nothing but the diagonal component of $`\stackrel{}{H}=(T_3,T_8)`$. The label $`j=1,2,3`$ can be assigned to the charge red($`R`$), blue($`B`$) and green($`G`$). We assume the cylindrical geometry of the system by taking $`𝒂=(r/2)𝒆_z`$, $`𝒃=(r/2)𝒆_z`$, and $`n_\mu =𝒆_z`$, where the distance between the quark and the anti-quark is defined by $`r`$. In this system, we get an explicit form of the singular dual gauge field from the relation (9) as $`𝑩_i^{\mathrm{sing}}=\sqrt{{\displaystyle \frac{2}{3}}}\stackrel{}{ϵ}_i\left[{\displaystyle \frac{\stackrel{}{Q}_j}{4\pi \rho }}\left({\displaystyle \frac{z+r/2}{\sqrt{\rho ^2+(z+r/2)^2}}}{\displaystyle \frac{zr/2}{\sqrt{\rho ^2+(zr/2)^2}}}\right)𝒆_\phi \right],`$ (21) where $`\phi `$ is the azimuthal angle around the $`z`$-axis and $`\rho `$ denotes the radial coordinate. Since the color-electric charges are defined on the weight vector of $`SU(3)`$ algebra, there arises the relation $$\stackrel{}{ϵ}_i\stackrel{}{w}_j=\frac{1}{2}\left(\begin{array}{ccc}0& 1& 1\\ 1& 0& 1\\ 1& 1& 0\end{array}\right)=\frac{1}{2}\underset{k=1}{\overset{3}{}}ϵ_{ijk}\frac{1}{2}m_{ij},$$ (22) where $`m_{ij}`$ takes 0 or $`\pm `$ 1. The zero of the diagonal component means that one of the monopole field is decoupled from the system and it does not contribute to the energy when we pay attention to the one of the color-electric charge, since the color-magnetic charge of the monopole field is defined on the root vector of $`SU(3)`$ algebra, as $`g\stackrel{}{ϵ}_i`$. Here, we investigate the ideal system for the limit $`r\mathrm{}`$. That is, $`\underset{r\mathrm{}}{lim}𝑩_i^{\mathrm{sing}}=\sqrt{{\displaystyle \frac{2}{3}}}{\displaystyle \frac{em_{ij}}{4\pi \rho }}𝒆_\phi ={\displaystyle \frac{m_{ij}}{g^{}\rho }}𝒆_\phi ,`$ (23) where we have used $`eg=4\pi `$ and $`g^{}=\sqrt{3/2}g`$. Then, the fields depend only on the radial coordinate, $$\varphi _i=\varphi _i(\rho ),𝑩_i^{\mathrm{reg}}=B_i^{\mathrm{reg}}(\rho )𝒆_\phi \frac{\stackrel{~}{B}_i^{\mathrm{reg}}(\rho )}{\rho }𝒆_\phi ,$$ (24) and the field equations (15) and (16) are reduced to $`{\displaystyle \frac{d^2\varphi _i}{d\rho ^2}}+{\displaystyle \frac{1}{\rho }}{\displaystyle \frac{d\varphi _i}{d\rho }}\left({\displaystyle \frac{g^{}\stackrel{~}{B}_i^{\mathrm{reg}}+m_{ij}}{\rho }}\right)^2\varphi _i2\lambda \varphi _i(\varphi _i^2v^2)=0,`$ (25) (26) $`{\displaystyle \frac{d^2\stackrel{~}{B}_i^{\mathrm{reg}}}{d\rho ^2}}{\displaystyle \frac{1}{\rho }}{\displaystyle \frac{d\stackrel{~}{B}_i^{\mathrm{reg}}}{d\rho }}2g^{}\left(g^{}\stackrel{~}{B}_i^{\mathrm{reg}}+m_{ij}\right)\varphi _i^2=0,`$ (27) The string tension can be defined by the energy per unit length of the flux-tube, $`\sigma =2\pi {\displaystyle \underset{i=1}{\overset{3}{}}}{\displaystyle _0^{\mathrm{}}}\rho 𝑑\rho \left[{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{1}{\rho }}{\displaystyle \frac{d\stackrel{~}{B}_i^{\mathrm{reg}}}{d\rho }}\right)^2+\left({\displaystyle \frac{d\varphi _i}{d\rho }}\right)^2+\left({\displaystyle \frac{g^{}\stackrel{~}{B}_i^{\mathrm{reg}}+m_{ij}}{\rho }}\right)^2\varphi _i^2+\lambda (\varphi _i^2v^2)^2\right],`$ (28) and we obtain the flux quantization condition, $$\mathrm{\Phi }_i=\frac{2\pi m_{ij}}{g^{}}.$$ (29) The boundary conditions are given by $`\stackrel{~}{B}_i^{\mathrm{reg}}=0,\varphi _i=\{\begin{array}{cc}0& (ij)\\ v& (i=j)\end{array}\mathrm{as}\rho 0,`$ (32) $`\stackrel{~}{B}_i^{\mathrm{reg}}={\displaystyle \frac{m_{ij}}{g^{}}},\varphi _i=v\mathrm{as}\rho \mathrm{}.`$ (33) Here, we shall confirm the relation (9). In this cylindrical system, the non-local term can be computed explicitly, $`\sqrt{{\displaystyle \frac{2}{3}}}\stackrel{}{ϵ}_i{\displaystyle \frac{1}{n}}\epsilon _{\mu \nu \alpha \beta }n^\alpha \stackrel{}{j}^\beta `$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{3}}}\stackrel{}{ϵ}_i\stackrel{}{Q}_j\delta (x)\delta (y)𝒆_z(\stackrel{}{Q}_je\stackrel{}{w}_j)`$ (34) $`=`$ $`\mathbf{}\times \left({\displaystyle \frac{m_{ij}}{g^{}\rho }}𝒆_\phi \right).`$ (35) As can be seen from this expression, one finds that this term exactly cancels with the color-electric field which is originated from the singular dual gauge field $`𝑩_i^{\mathrm{sing}}`$ in (23). It shows that the kinetic term of the dual gauge field in the Lagrangian (10) can be written with no singular field. ## IV Bogomol’nyi limit In this section, we discuss the properties of the flux-tube in the Bogomol’nyi limit. Since now we have the same Lagrangian with $`U(1)`$ gauge symmetry except only the labels of $`i`$ and $`j`$ which classify the kinds of the monopole and the quark corresponding to $`[U(1)]_m^3`$ dual gauge symmetry, we can use the same strategy to find the Bogomol’nyi limit as given in Ref. . Thus, we can write the string tension (28) exactly in the form, $`\sigma `$ $`=`$ $`2\pi {\displaystyle \underset{i=1}{\overset{3}{}}}|m_{ij}|v^2+2\pi {\displaystyle \underset{i=1}{\overset{3}{}}}{\displaystyle _0^{\mathrm{}}}\rho d\rho [{\displaystyle \frac{1}{2}}({\displaystyle \frac{1}{\rho }}{\displaystyle \frac{d\stackrel{~}{B}_i^{\mathrm{reg}}}{d\rho }}\pm g^{}(\varphi _i^2v^2))^2`$ (37) $`+({\displaystyle \frac{d\varphi _i}{d\rho }}\pm (g^{}\stackrel{~}{B}_i^{\mathrm{reg}}+m_{ij}){\displaystyle \frac{\varphi _i}{\rho }})^2+{\displaystyle \frac{1}{2}}(2\lambda g^2)(\varphi _i^2v^2)^2].`$ From this expression, we find the Bogomol’nyi limit, $`g_{}^{}{}_{}{}^{2}=2\lambda ,\mathrm{or}3g^2=4\lambda .`$ (38) In this limit, one find that the string tension is reduced to $`\sigma `$ $`=`$ $`2\pi {\displaystyle \underset{i=1}{\overset{3}{}}}|m_{ij}|v^2=4\pi v^2,`$ (39) and the profiles of the dual gauge field and the monopole field is determined by the first order differential equations, $`{\displaystyle \frac{d\varphi _i}{d\rho }}\pm \left(g^{}\stackrel{~}{B}_i^{\mathrm{reg}}+m_{ij}\right){\displaystyle \frac{\varphi _i}{\rho }}=0,`$ (40) $`{\displaystyle \frac{1}{\rho }}{\displaystyle \frac{d\stackrel{~}{B}_i^{\mathrm{reg}}}{d\rho }}\pm g^{}(\varphi _i^2v^2)=0.`$ (41) These field equations of cource reproduce the second order differential equations (25) and (27) when the relation (38) is satisfied. Here, to obtain the string tension of the form (37) and the saturated string tension (39), we have paid attention to the boundary conditions of the fields (33) by taking into account the relation (22). For instance, let us consider the $`R`$-$`\overline{R}`$ flux-tube, which is given by the label $`j=1`$. In this system, the monopole field $`\varphi _1`$ which has the magnetic charge $`g\stackrel{}{ϵ}_1`$ is decoupled from the system, since $`\varphi _1`$ does not feel any singularity of the flux-tube core, and accordingly, the regular dual gauge field $`B_1^{\mathrm{reg}}`$ is also decoupled. The behavior of the other fields is interesting, $`\varphi _2`$ and $`\varphi _3`$ behaves as the same monopole field, and $`B_2^{\mathrm{reg}}`$ and $`B_3^{\mathrm{reg}}`$ provides the $`U(1)_{i=2}`$ flux-tube and $`U(1)_{i=3}`$ anti flux-tube due to the sign of the $`m_{ij}`$, which takes $`1`$ and $`1`$, respectively. Here, both dual gauge fields are related with each other through the constraint $`_{i=1}^3B_i^{\mathrm{reg}}=0`$, and $`U(1)_{i=3}`$ anti flux-tube can be regarded as the $`U(1)_{i=2}`$ flux-tube, or vise versa. As a result, these flux-tubes provide the same string tension $`2\pi v^2`$, and finally, we get two times of this string tension, $`2\times 2\pi v^2`$. This is caused by the $`[U(1)]_m^2`$ dual gauge symmetry. We note that this discussion is the Weyl symmetric, and thus, the final expression for the string tension (39) does not depend on kind of the color-electric charges $`\stackrel{}{Q}_j`$. The profiles of the color-electric field can be obtained by solving the first order equations (40) and (41) by taking into account the above discussion as is discussed in Ref. . Let us consider the meaning of (38). Here, we can define two characteristic scales using three parameters in the DGL theory, $`g`$, $`\lambda `$ and $`v`$. One is the mass of the dual gauge field $`m_B=\sqrt{2}g^{}v=\sqrt{3}gv`$ and the other is the mass of the monopole field $`m_\chi =2\sqrt{\lambda }v`$. These masses are extracted from the Lagrangian (10) by taking into account the dual Higgs mechanism. Thus, one finds that the Bogomol’nyi limit in the DGL theroy (38) is the supersymmetry between the dual gauge field and the monopole field. Since these inverse masses $`m_B^1`$ and $`m_\chi ^1`$ corresponds to the penetration depth of the color-electric field and the coherent length of the monopole field, respectively, the Ginzburg-Landau (GL) parameter is defined: $`\stackrel{~}{\kappa }{\displaystyle \frac{m_B^1}{m_\chi ^1}}={\displaystyle \frac{\sqrt{2\lambda }}{g^{}}}={\displaystyle \frac{2\sqrt{\lambda }}{\sqrt{3}g}}.`$ (42) Therefore, $`\stackrel{~}{\kappa }=1`$ is regarded as the Bogomol’nyi limit, and the vacuum is classified into two types in terms of the Bogomol’nyi limit: $`\stackrel{~}{\kappa }<1`$ belongs to the type-I vacuum and $`\stackrel{~}{\kappa }>1`$ is the type-II vacuum. Now, we would like to discuss the interaction between two parallel flux-tubes of the same type, such as the system $`R`$-$`\overline{R}`$ and $`R`$-$`\overline{R}`$. In general, the flux-tubes would interact with each other. However, in the Bogomol’nyi limit, there is no interaction between them. This can be understood through an investigtion of the generalized string tension for an exotica that the color-electric charges are given by $`n\stackrel{}{Q}_j`$ and $`n\stackrel{}{Q}_j`$ for an integer $`n`$. In this system, we get the generalized string tension, $`\sigma _n=4\pi nv^2,`$ (43) where $`m_{ij}`$ is simply replaced to $`nm_{ij}`$. One finds that the string tension (43) is proportional to $`n`$, which implies that the interaction energy is zero. It is considered that this comes from the balance of propagation range of the dual gauge field and the monopole field since $`m_Bm_\chi `$. In the type-I or in the type-II vacuum, which is away from the Bogomol’nyi limit, the interaction range of these fields lose its balance, and the flux-tube interaction manifestly appears. The string tension is not proportional to $`n`$ any more. While the attractive force is worked between two parallel flux-tubes in the type-I vacuum, the flux-tubes repel each other in the type-II vacuum. Numerical investigations of the interaction between two or more parallel flux-tubes of the same type in the DGL theory are given in Ref. . It is interesting to investigate what happens if two parallel flux-tubes of different types are placed at a certain distance. Here, according to the $`[U(1)]_m^3`$ dual gauge symmetry, there appear three different types of the flux-tube, such as given by $`R`$-$`\overline{R}`$, $`B`$-$`\overline{B}`$, and $`G`$-$`\overline{G}`$, so that these interactions seem to be very complicated. However, now the system has remarkable aspects owing to the Weyl symmetry. For instance, let us consider the interaction between $`R`$-$`\overline{R}`$ and $`B`$-$`\overline{B}`$. We find that the interaction between them is attractive, since if we suppose that these flux-tubes are unified into one flux-tube, it becomes $`\overline{G}`$-$`G`$ (See the relation (22)). It means that the energy of the system after unification is reduced into a half of the initial one. The same interaction property would be observed in the process, $`B`$-$`\overline{B}`$ \+ $`G`$-$`\overline{G}`$ $``$ $`\overline{R}`$-$`R`$ and $`G`$-$`\overline{G}`$ \+ $`R`$-$`\overline{R}`$ $``$ $`\overline{B}`$-$`B`$. These investigations show that if we pay attention to the Weyl symmetry, we can easily obtain qualitative information about the flux-tube interaction. ## V Conclusion We have studied the flux-tube solution in the DGL theory in the Bogomol’nyi limit by using the manifestly Weyl invariant form of the DGL Lagrangian. Here, the original dual gauge symmetry $`[U(1)]_m^2`$ is extended to $`[U(1)]_m^3`$. This replacement makes the further manipulation of the Lagrangian analogous to the $`U(1)`$ case. We have found that the Bogomol’nyi limit is given by $`3g^2=4\lambda `$, and the string tension is calculated as $`\sigma _n=4\pi nv^2`$ for a $`q`$-$`\overline{q}`$ pair with the charge $`nQ_j`$ and $`nQ_j`$ in the both ends. In this limit, the mass of the dual gauge field and the monopole field becomes exactly the same. It should be noted that we could see the same relation with $`U(1)`$ Abelian Higgs theory except for three different types of the flux-tube. To summarize, the very similar properties with the ANO vortex in the Abelian Higgs theory is observed when we see the single flux-tube in the DGL theory, and the flux-tube solution can be easily obtained if we pay attention to the Weyl symmetry in the the DGL theory. Finally, we would like to mention about the relation between the work in Ref. and our study. If we replace the monopole field and the parameters that they have used as $`\chi \sqrt{2}\chi `$, $`\eta \sqrt{2}v`$, and $`\lambda \lambda /4`$, we get the same framework at the starting point, and the Bogomol’nyi limit is replaced as $`3g^2=16\lambda 3g^2=4\lambda `$. The idea of the extension of the dual gauge symmetry based on the Weyl symmetry in our case, however, seems to be simple to reach the final expression on the string tension, which can be applied to the $`[U(1)]^{N1}`$ dual Abelian Higgs theroy reduced from the $`SU(N)`$ gluodynamics, straightforwardly. ## Acknowledgment The authors are grateful to H. Suganuma for fruitful discussions and in particular stressing the importance of the Weyl symmetry in the DGL theory. We also acknowledge M. I. Polikarpov to inform us of the work of Ref. , which is closely related to our work.
warning/0004/nlin0004009.html
ar5iv
text
# References nlin.SI/0004009 SNBNCBS-2000 Geometrical Aspects of Integrability in Nonlinear Realization Scheme R.P.Malik <sup>*</sup><sup>*</sup>* E-mail address: malik@boson.bose.res.in Invited talk delivered in a workshop on “Dynamical Systems: Recent Developments” (held from November 4 to November 6, 1999) at School of Physics, University of Hyderabad (India). S. N. Bose National Centre for Basic Sciences, Block-JD, Sector-III, Salt Lake, Calcutta-700 091, India ABSTRACT We discuss the integrability properties of the Boussinesq equations in the language of geometrical quantities defined on an appropriately chosen coset manifold connected with the $`W_3`$ algebra of Zamolodchikov. We provide a geometrical interpretation to the commuting conserved quantities, Lax-pair formulation, zero-curvature representation, Miura maps, etc. in the framework of nonlinear realization method. Two ($`1+1`$) dimensional integrable nonlinear partial differential equations have played a notable role in the understanding of some of the physical phenomena of nature. The KdV and Boussinesq equations, their modified versions, their higher order hierarchies, etc., are the cardinal examples of such a class of equations which have found applications in as diverse areas of research as two-dimensional (2D) conformal field theories, ($`W`$)-string theories, fluid mechanics, plasma physics, 2D ($`W`$)-gravity theories, etc.\[1-4\]. The latter equation (i.e., the Boussinesq equation) can be realized on $`u(x,t)`$ and $`v(x,t)`$ fields as $$\begin{array}{ccc}\frac{u}{t}=\frac{160}{3}v^{},\frac{v}{t}=\frac{1}{10}u^{\prime \prime \prime }\frac{24}{5}uu^{},\hfill & & \end{array}$$ $`(1)`$ which can be combined together to yield a nonlinear partial differential equation (NLPDE) realized on the single field $`u(x,t)`$ as $$\begin{array}{ccc}\frac{^2u}{t^2}=\frac{16}{3}u^{\prime \prime \prime \prime }+256uu^{\prime \prime }+256u^{}u^{},\hfill & & \end{array}$$ $`(2)`$ where the primes (i.e., $`u^{}=\frac{u}{x},v^{}=\frac{v}{x}`$) denote the partial derivatives on the fields with respect to the space variable $`x`$ and it has, as is physically evident, the dimension of length $`L`$ (i.e. $`xL`$). Taking into account this dimension, it can be readily seen that the evolution parameter $`t`$, fields $`u(x,t)`$ and $`v(x,t)`$ have the dimensions $`L^2,L^2`$ and $`L^3`$ respectively. In the more sophisticated language of conformal field theory, one says that the naive conformal dimensions of $`t,u,v`$ are $`2,+2,+3`$ respectively. In what follows, we shall be calling it the conformal spins as is the practice in the realm of research activities in the rational conformal field theories 2D conformal field theory can be written in terms of the complex variables $`z`$ and $`\overline{z}`$. There are holomorphic and antiholomorphic parts in the theoy which factorize due to conformal invariance. The conformal spin is actually defined as ($`h\overline{h}`$) where $`h`$ and $`\overline{h}`$ are the conformal weights of a primary field.. We shall see below how this dimensional analyses will be useful in our discussions for the nonlinear realization method (connected with the group realizations on homogeneous spaces). One of the key methods used in the discussions of the spontaneously broken gauge theories is the coset space construction $`(G/H)`$ where the total Lagrangian density of the theory is found to be invariant under the group $`G`$ and the vacuum of the theory is found to be invariant under the subgroup $`H`$. This coset space construction turns out to be useful in the determination of the number of Goldstone bosons, massless gauge fields, massive gauge bosons, etc., in the language of group theory. To understand the geometry behind a symmetry group $`G`$, however, the key concept is to consider it as a group of transformations acting on the coset space $`(G/H)`$ for the appropriately chosen stability subgroup $`H`$. This was the starting point for the nonlinear realization method applied to the spontaneously broken chiral gauge theories in the late 60’s and early 70’s where corresponding Lie algebras were found to be linear. In the recent past, however, this method was exploited for the discussion of geometry behind the 2D (super) NLPDE connected with (super) W-type algebras which are nonlinear to begin with. In fact, in these works, an infinite dimensional linear algebra was constructed from the nonlinear (super) $`W_3`$ algebra and all the techniques of the nonlinear realization method were exploited. In this presentation, we shall see how some of the key properties of integrability of the Boussinesq equation can be understood in the language of geometry on the coset manifold. Before we come over to the description of the coset space construction for the $`W_3`$ algebra of Zamolodchikov, we shall dwell a bit on the basic concepts associated with this method. A realization of a Lie group (or corresponding Lie algebra) is a mathematical concept and it corresponds to the validity of certain specific type of differential equations for a group valued function associated with the Lie group. As an example, it is a well known fact that the linear realization of a compact, connected (semi)simple Lie group (or corresponding Lie algebra) is nothing but the (matrix) representation of the Lie algebra. To elaborate and explain these statements, let us begin with a set of fields $`\psi _n`$ (where $`n`$ is the multiplicity) and consider the action of the group elements $`g`$ of the compact Lie group $`G`$ such that the field $`\psi _n`$ transforms as $$\begin{array}{ccc}g:\psi _nf_n(\psi ,g),\hfill & & \end{array}$$ $`(3)`$ where $`f_n(\psi ,g)`$ is an abstract form of the transformed field $`\psi _n`$. Now one can impose all the four properties of a group on transformation (3). For instance, the action of the identity $`\mathrm{𝟏}`$, the existence of the inverse $`(g^1g=gg^1=\mathrm{𝟏}`$) and the closure property can be explicitly explained in the language of transformations as $$\begin{array}{ccc}\mathrm{𝟏}:\psi _nf_n(\psi ,\mathrm{𝟏})=\psi _n,\hfill & & \end{array}$$ $`(4)`$ $$\begin{array}{ccc}gg^1:\psi _nf_n(\psi ,g^1)f_n(f(\psi ,g^1);g)\psi _n,\hfill & & \end{array}$$ $`(5)`$ $$\begin{array}{ccc}g^1g:\psi _nf_n(\psi ,g)f_n(f(\psi ,g);g^1)\psi _n,\hfill & & \end{array}$$ $`(6)`$ $$\begin{array}{ccc}g_1g_2:\psi _nf_n(\psi ,g_2)f_n(f(\psi ,g_2);g_1)f_n(\psi ,g_1g_2),\hfill & & \end{array}$$ $`(7)`$ where the last equation is just the closure property under binary operation ($``$). Here the binary operation is nothing but the transformation (3). One can exploit this property to establish the associativity: $$\begin{array}{ccc}f_n(\psi ,g_1(g_2g_3))=f_n(\psi ,(g_1g_2)g_3).\hfill & & \end{array}$$ $`(8)`$ So far, the group element $`g`$ was treated as an abstract object, now we can write the explicit form of it in terms of a set of infinitesimal transformation parameters $`ϵ_\alpha `$ as $$\begin{array}{ccc}g=\mathrm{𝟏}+iϵ_\alpha \mathrm{\Gamma }_\alpha ,\hfill & & \end{array}$$ $`(9)`$ where $`\mathrm{\Gamma }_\alpha `$ are the set of generators for the above transformations which obey a commutation relationship for the given Lie algebra as $$\begin{array}{ccc}[\mathrm{\Gamma }_\alpha ,\mathrm{\Gamma }_\beta ]=iC_{\alpha \beta \gamma }\mathrm{\Gamma }_\gamma .\hfill & & \end{array}$$ $`(10)`$ Here $`C_{\alpha \beta \gamma }`$ are the structure constants of the algebra. The explicit form of the transformed field (with $`f_n(\psi ,\mathrm{𝟏})=\psi _n,`$ and eqn. (9)) is $$\begin{array}{ccc}f_n(\psi ,g)\hfill & =& f_n(\psi ,\mathrm{𝟏}+iϵ_\alpha \mathrm{\Gamma }_\alpha )\hfill \\ & & \psi _n+iϵ_\alpha f_{n\alpha }(\psi )+O(ϵ^2),\hfill \end{array}$$ $`(11)`$ where $`f_{n\alpha }(\psi )`$ is a group valued function and it retains the information about the multiplicity as well as the group properties. For a given Lie algebra, the group valued functions $`f_{n\alpha }`$ can not be chosen in a perfectly arbitrary way. Rather, they obey certain specific kind of differential equation for the given Lie algebra. For instance, using the closure property of equation (7), it can be seen that $$\begin{array}{ccc}f_n(\psi ,g_1^1g_2g_1)\hfill & =& f_n(f(\psi ,g_2g_1);g_1^1)\hfill \\ & & f_n(f\{f(\psi ,g_1);g_2\};g_1^1).\hfill \end{array}$$ $`(12)`$ With the following inputs from the group properties (see, e.g., eqns. 4–8) $$\begin{array}{ccc}f_n(\psi _m,\mathrm{𝟏})=\delta _{nm}\psi _m,f_n(f(\psi ,g);g^1)=\psi _n,\hfill & & \end{array}$$ $`(13a)`$ $$\begin{array}{ccc}f_n(f(\psi ,g)=\psi _n+iϵ_\alpha (\frac{f_n(\psi )}{\psi _m})f_{m\alpha }+O(ϵ^2),\hfill & & \end{array}$$ $`(13b)`$ we obtain the following differential equation satisfied by $`f_{n\alpha }`$ $$\begin{array}{ccc}f_{n\beta }C_{\alpha \gamma \beta }=i\frac{f_{n\alpha }}{\psi _m}f_{m\gamma }+i\frac{f_{n\gamma }}{\psi _m}f_{m\alpha }.\hfill & & \end{array}$$ $`(14)`$ All set of functions $`f_{n\alpha }(\psi )`$ that satisfy the above differential equation is said to provide a realization of the given Lie algebra. For instance, it can be seen that the following linear choice of $`f_{n\alpha }(\psi )`$ $$\begin{array}{ccc}f_{n\alpha }(\psi )=(\tau _\alpha )_{nm}\psi _m,\hfill & & \end{array}$$ $`(15)`$ leads to $$\begin{array}{ccc}[\tau _\alpha ,\tau _\beta ]=iC_{\alpha \beta \gamma }\tau _\gamma ,\hfill & & \end{array}$$ $`(16)`$ which is nothing but the matrix representation of the Lie algebra in (10). For the spontaneously broken (chiral) gauge theories, the generators $`\mathrm{\Gamma }_\alpha `$ of the full group $`G`$ have two parts. Some of the generators $`T_i`$ ($`i:`$ dimensionality of $`H`$) belong to the subgroup $`H`$ and the rest of the generators $`X_a`$ belong to the coset space $`G/H`$. Thus, one can parametrize the coset space by coset fields and the generators $`X_a`$. It was shown in Refs. that one can parametrize the coset space in terms of exponentials and it also provides a realization for the factorized Lie algebra under consideration. Due to the presence of exponentials, this special realization is known as the nonlinear realization. It can be seen that the linear term of this realization (exponentials) is nothing but the representaion (cf. eqn. (16)) of the Lie algebra if the stability subalgebra contains only identity element of the Lie algebra. Thus, we see that the linear realization is a special case of the nonlinear realization. The first thing we note, before the application of nonlinear realization method to the classical $`W_3`$ algebra with central extension (parametrized by the central charge $`c`$) $$\begin{array}{ccc}& & [L_n,L_m]=(nm)L_{n+m}+\frac{c}{12}(n^3n)\delta _{n+m,0},(\mathrm{}n,m+\mathrm{})\hfill \\ & & [L_n,W_m]=(2nm)W_{n+m},\hfill \\ & & [W_n,W_m]=16(nm)\mathrm{\Lambda }_{n+m}\frac{c}{9}(n^3n)(n^24)\delta _{n+m,0}\hfill \\ & & \frac{8}{3}(nm)(n^2+m^2\frac{1}{2}nm4)L_{n+m},\hfill \end{array}$$ $`(17)`$ is the fact that it is a nonlinear algebra because of the composite and nonlinear nature of $`\mathrm{\Lambda }_n=\frac{8}{3}_mL_{nm}L_m`$. It was thus essential to get a linear algebra out of it so that the whole arsenal of techniques of coset space construction can be applied here. To this goal, in Ref. , all the higher spin composite generators were treated as independent generators. Invoking this idea immediately entails upon the $`W_3`$ algebra to become an infinite dimensional linear algebra $`W_3^{\mathrm{}}`$ as $$\begin{array}{ccc}W_3^{\mathrm{}}=\{L_n,W_n,\mathrm{\Lambda }_n,\mathrm{\Phi }_n\mathrm{}\mathrm{}\mathrm{}J_n^h\mathrm{}\mathrm{}\mathrm{}..\},\hfill & & \end{array}$$ $`(18)`$ where $`\mathrm{\Phi }_n(WT)_n`$ is the conformal spin-5 composite generator and $`J_n^h`$ is a generic composite generator with conformal spin-$`h`$. It can be readily seen by taking a single contraction of the OPE’s that all the higher conformal spin ($`4`$) composite generators form a closed algebra among themselves and they form an ideal. One of the key subalgebras of this infinite dimensional algebra is the one in which the Laurent indices of the generators with conformal spin-$`h`$ (i.e., $`J(z)={\displaystyle \underset{n}{}}J_n^hz^{nh}`$) vary from $`(h1)`$ to $`\mathrm{}`$. For instance, for the conformal spin-2, we have the generators: $`L_1,L_0,L_{+1},L_{+2}\mathrm{}\mathrm{}.`$ and for the conformal spin-3, we have $`W_2,W_1,W_0,W_1\mathrm{}\mathrm{}\mathrm{}`$ and so on and so forth. The stability subalgebra $``$ of our interest, from this truncated version of $`W_3^{\mathrm{}}`$, is $$\begin{array}{ccc}=\{W_1+2L_1,W_0,W_1,W_2,L_1,L_0,\mathrm{\Lambda }_n(n3),J_n^h(h5,nh+1)\},\hfill & & \end{array}$$ $`(19)`$ where it can be easily seen that $`W_2`$ is not present and $`L_1`$ appears in a particular linear combination with $`W_1`$ for the closure of the algebra. Thus, $`W_2`$ and $`L_1`$ can be taken into the coset space. It will be also noticed that all the higher order conformal spin composite generators have been taken into the stability subalgebra as they form an ideal. Now the element $`g`$ in the coset space can be parametrized as $$\begin{array}{ccc}g\frac{𝒢}{}=e^{tW_2}e^{xL_1}e^{\psi _3L_3}\left(\mathrm{\Pi }_{n=4}e^{\psi _nL_n}e^{\xi _nW_n}\right)e^{uL_2}e^{vW_3}.\hfill & & \end{array}$$ $`(20)`$ It is an interesting point to note that out of all the generators in the coset space, only two generators (i.e., $`L_1,W_2`$) commute with each other. They have the dimensions of length as : $`W_2L^2,L_1L^1`$. To make the exponentials in equation (15) dimensionless, it is clear that $`t`$ and $`x`$ must have dimensions of length as: $`tL^2,xL`$. It is illuminating to see that these are exactly the dimensions of $`t`$ and $`x`$ for the Boussinesq equations which were discussed after eqn.(2). The commutativity of the generators $`W_2,L_1`$ ensures that the $`t`$ and $`x`$ directions are linearly independent on the coset manifold and, therefore, they can be treated as coordinates. This feature should be contrasted with other generators and associated tower of coset (Goldstone) fields $`u,v,\psi _3,\xi _4,\psi _4,\xi _5\mathrm{}\mathrm{}\mathrm{}.`$ which cannot be treated as coordinates. Now any point on the manifold can be parametrized by the coordinates $`x`$ and $`t`$ and all the fields can be treated as functions of these coordinates. The most important geometrical quantity in the framework of nonlinear realization method is the one-differential Cartan form $`\mathrm{\Omega }=g^1dg`$ (where $`g\frac{𝒢}{}`$) in terms of which the curvature tensor, torsion, complex structure, etc. of the coset manifold can be determined. Due to its very structure, it obeys the following Maurer-Cartan equation $$\begin{array}{ccc}d^{ext}\mathrm{\Omega }+\mathrm{\Omega }\mathrm{\Omega }=0,\hfill & & \end{array}$$ $`(21)`$ which is nothing but the zero-curvature representation for the non-Abelain gauge theory if we choose the 1-differential Cartan form in terms of the gauge connections $`(A_\mu =A_\mu ^aT^a)`$ as: $`\mathrm{\Omega }=A_\mu dx^\mu `$. Here $`T^a`$ are the generators of the Lie algebra under consideration. Now, it can be seen that the non-Abelain curvature tensor, emerging from (21), is zero, namely; $$\begin{array}{ccc}& & F_{\mu \nu }=[D_\mu ,D_\nu ]=_\mu A_\nu _\nu A_\mu +[A_\mu ,A_\nu ]=0,\hfill \\ & & D_\mu =_\mu +A_\mu ,A_\mu =A_\mu ^aT^adx^\mu .\hfill \end{array}$$ $`(22)`$ In fact, in the language of gauge theory, the choice $`\mathrm{\Omega }=g^1dg`$ is eaxctly like the pure gauge choice. Thus, the zero curvature representation is bound to be satisfied. For the truncated version of the algebra $`W_3^{\mathrm{}}`$, we obtain $$\begin{array}{ccc}\mathrm{\Omega }=g^1dg=\underset{n=1}{\overset{\mathrm{}}{}}\omega _nL_n+\underset{n=2}{\overset{\mathrm{}}{}}\theta _nW_n+\text{ higher spin contributions}.\hfill & & \end{array}$$ $`(23)`$ As higher spin composite generators form an ideal, it is essential to know only some of the lower order forms to obtain the dynamical equations of motion if we exploit the ideas of Inverse Higgs-Covariant Reduction (IH-CR) procedure . These lower order forms are $$\begin{array}{ccc}\omega _1\hfill & =& dx,\omega _0=0,\omega _1=3udx+160vdt,\omega _2=du4\psi _3dx+320\xi _4dt,\hfill \\ \omega _3\hfill & =& d\psi _3+(\frac{3}{2}u^25\psi _4)dx+(560\xi _5240uv)dt,\hfill \\ \omega _4\hfill & =& d\psi _46\psi _5dx+(896\xi _6192v\psi _3768u\xi _4)dt,\hfill \\ & & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..\hfill \\ \theta _2\hfill & =& dt,\theta _1=0,\theta _0=6udt,\theta _1=8\psi _3dt,\hfill \\ \theta _2\hfill & =& 5vdx+(12u^210\psi _4)dt,\theta _3=dv6\xi _4dx+(24u\psi _312\psi _5)dt\hfill \\ & & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}.\hfill \end{array}$$ $`(24)`$ According to the IH-CR procedure, one can set equal to zero all the components of the forms connected to the generators in the coset space. In fact, these forms transform homogeneously under the left action of the truncated version of $`W_3^{\mathrm{}}`$ symmetry and setting them equal to zero does not spoil the symmetry of the group. These constraints are just like the “gauge-fixing” conditions on the gauge-connections in the language of gauge theory. To make this statement more transparent, it can be seen that the following constraints: $$\begin{array}{ccc}\omega _n=0,n2,\theta _n=0,n3,\hfill & & \end{array}$$ $`(25)`$ lead to the following kinematical and dynamical equations $$\begin{array}{ccc}\psi _3\hfill & =& \frac{1}{4}u^{},\psi _4=\frac{1}{5}\psi _3^{}+\frac{3}{10}u^2,\psi _5=\frac{1}{6}\psi _4^{},\xi _4=\frac{1}{6}v^{},\hfill \\ \dot{u}\hfill & =& \frac{160}{3}v^{},\dot{v}=\frac{1}{10}u^{\prime \prime \prime }\frac{24}{5}uu^{},\dot{u}=\frac{u}{t},\dot{v}=\frac{v}{t}.\hfill \end{array}$$ $`(26)`$ Thus, we see that the Boussinesq equation of eqn. (1) emerges here by the IH-CR procedure applied on the coset manifold as all the tower of fields can be expressed in terms of the essential fields $`u`$ and $`v`$ and the derivatives on them . In the language of geometrical properties on the coset manifold, it can be seen that the Boussinesq equations are nothing but the embedding conditions on a two dimensional $`(x,t)`$ geodesic surface (parametrized by the basic fields $`u(x,t)`$ and $`v(x,t)`$ and the derivatives on them) when one singles out this hypersurface from the infinite dimensional coset manifold. Mathematically, the Boussinesq equation can be understood in the language of group motion when infinite dimensional algebra of the infinite dimensional coset manifold reduces covariantly to the ‘covariant reduced algebra’ generated by the elements of the $`sl(3,R)`$ algebra. In other words, the original Cartan form now reduces to a reduced Cartan form (due to IH-CR procedure) as $$\begin{array}{ccc}\mathrm{\Omega }=g^1dg\mathrm{\Omega }_{red}=g_{red}^1dg_{red}=\underset{n=2}{\overset{n=2}{}}\theta _nW_n+\underset{n=1}{\overset{n=+1}{}}\omega _nL_n,\hfill & & \end{array}$$ $`(27)`$ which satisfies the Maurer-Cartan equation: $`d^{ext}\mathrm{\Omega }_{red}+\mathrm{\Omega }_{red}\mathrm{\Omega }_{red}=0`$. The explicit form of this reduced Cartan form is $$\begin{array}{ccc}& & \mathrm{\Omega }_{red}=A_xdx+A_tdt,\hfill \\ & & A_x=L_13uL_15vW_2,\hfill \\ & & A_t=160vL_16uW_0+W_28\psi _3W_1+(12u^210\psi _4)W_2.\hfill \end{array}$$ $`(28)`$ It can be now readily seen that the following zero-curvature condition $$\begin{array}{ccc}F_{tx}=[_t+A_t,_x+A_x]=0,\hfill & & \end{array}$$ $`(29)`$ leads to the derivation of the Boussinesq equations. It will be noticed that $`A_x`$ and $`A_t`$ are nothing but the $`sl(3,R)`$ valued Drinfeld-Sokolov type Lax-pairs in eqn. (28). In the language of geometry, these Lax-pairs can be understood as the projections of the reduced Cartan form along $`x`$ and $`t`$ directions of the coset manifold (cf. eqn. (28)). It is straightforward to notice that if the generators $`W_2,W_1,L_1`$ are taken out from the stability subalgebra in (19), still the algebra will be closed. Thus, we obtain a new subalgebra $`_1`$ from $``$ as $$\begin{array}{ccc}_1=\{W_1+2L_1,W_0,L_0,\mathrm{\Lambda }_n(n3),J_n^h(h5,nh+1)\}.\hfill & & \end{array}$$ $`(30)`$ In this case, the coset space can be parametrized as $$\begin{array}{ccc}g_1\frac{𝒢}{_1}=e^{tW_2}e^{xL_1}e^{\psi _3L_3}\left(\mathrm{\Pi }_{n=4}e^{\psi _nL_n}e^{\xi _nW_n}\right)e^{uL_2}e^{vW_3}e^{u_1L_1}e^{v_1W_1}e^{v_2W_2}.\hfill & & \end{array}$$ $`(31)`$ Furthermore, one can take out $`W_0,L_0`$ from the stability subalgebra (30) and still algebra will be closed with only one $`U(1)`$ basic generator ($`W_1+2L_1`$). Now the coset element is $$\begin{array}{ccc}g_2\frac{𝒢}{_2}=g_1e^{u_0L_0}e^{v_0W_0},\hfill & & \end{array}$$ $`(32)`$ where $`_2`$ is given by $$\begin{array}{ccc}_2=\{W_1+2L_1,\mathrm{\Lambda }_n(n3),J_n^h(h5,nh+1)\}.\hfill & & \end{array}$$ $`(33)`$ In both the cases of $`g_1`$ and $`g_2`$, one can define a one-differential Cartan form $`\mathrm{\Omega }_1=g_1^1dg_1`$ and $`\mathrm{\Omega }_2=g_2^1dg_2`$ and apply the IH-CR procedure on it. This leads to a covariant relationship between essential fields $`u_1,v_1`$ of the coset manifold (31) and $`u,v`$ fields of coset the manifold (20). Similarly, one gets a relationship between essential fields $`u_0,v_0`$ of coset manifold (32) and $`u_1,v_1`$ fields of (30). These are nothing but the so-called Miura maps. The dynamical equations on $`u_0,v_0`$ and $`u_1,v_1`$ fields can also be obtained due to appropriate application of CR procedure on coset manifolds (30) and (32) as we obtained for the essential fields $`u,v`$ in eqn. (26). Thus, we see that the Miura maps are nothing but the kinematical relationships among essential fields when one goes covariantly from one coset manifold to another one. It is obvious that the kinematical and dynamical relationships can be obtained from the nonlinear realization method by application of the IH-CR procedure . For the first time, however, this procedure was extended one step further to derive commuting conserved quantities for the Boussinesq equations in Ref. . We shall briefly dwell a bit on it. For the derivation of the commuting conserved quantities, one has to compute more higher order forms than the ones required for the derivation of the dynamical equations. Some of these forms are $$\begin{array}{ccc}\omega _5\hfill & =& d\psi _5+ud\psi _3+\left(\frac{1}{2}u^35u\psi _4+2\psi _3^240v^27\psi _6\right)dx\hfill \\ & +& \left(192u^2v336u\xi _5704\psi _3\xi _4160v\psi _4+1344\xi _7\right)dt,\hfill \\ \omega _6\hfill & =& d\psi _6+2ud\psi _4+(8\psi _3\psi _412u\psi _58\psi _7)dx\hfill \\ & +& \left(1920\xi _8+768u\xi _6+768u^2\xi _4640\psi _4\xi _41664\psi _3\xi _5\right)dt,\hfill \\ & .& \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..\hfill \\ & .& \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..\hfill \\ \theta _4\hfill & =& d\xi _4+(3uv7\xi _5)dx+\left(20\psi _3^2+20u\psi _414\psi _68u^380v^2\right)dt,\hfill \\ \theta _5\hfill & =& d\xi _5+vdu+(6u\xi _44v\psi _38\xi _6)dx\hfill \\ & +& \left(56\psi _3\psi _4+320v\xi _4+12u\psi _512u^2\psi _316\psi _7\right)dt,\hfill \\ \theta _6\hfill & =& d\xi _6+3vd\psi _3+\left(\frac{9}{2}u^2v15v\psi _49\xi _7\right)dx\hfill \\ & +& bigl(72\psi _3\psi _5+1680v\xi _5+30\psi _4^2360v^2u18\psi _8)dt,\hfill \\ & .& \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..\hfill \\ & .& \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..\hfill \end{array}$$ $`(34)`$ It can be readily seen that, due to IH-CR procedure, we can set the $`dt`$ projection of the form $`\theta _5`$ equal to zero. This leads to the following equation $$\begin{array}{ccc}\frac{\xi _5}{t}=16\psi _7+12u^2\psi _312u\psi _556\psi _3\psi _4320v\xi _4.\hfill & & \end{array}$$ $`(35)`$ Now using the kinematical relationships, it can be seen that the r.h.s. of the above expression is a total space derivative $$\begin{array}{ccc}\frac{\xi _5}{t}=\frac{}{x}\left[2\psi _6+\frac{1}{5}u^32u\psi _4\frac{16}{3}\psi _3^2\right],\hfill & & \end{array}$$ $`(36)`$ which ultimately leads to the following conservation law $$\begin{array}{ccc}\frac{(uv)}{t}=\frac{}{x}\left[\frac{11}{5}u^3+2u\psi _4\frac{16}{3}\psi _3^2\frac{80}{3}v^2\right],\hfill & & \end{array}$$ $`(37)`$ where we have used $`\xi _5=\frac{\xi _4^{}}{7}+\frac{3}{7}uv,\psi _6=\frac{1}{7}\left(\psi _5^{}+2\psi _3^2u^340v^2\right)`$ from the kinematical relationships. Similarly other conserved quantities can be calculated (see, e.g., Ref. for details). Some of the conserved quantities are $$\begin{array}{ccc}H_1\hfill & =& \frac{c}{2}𝑑xu(x,t),H_2=\frac{40c}{3}𝑑xv(x,t),H_4=c𝑑x(uv)(x,t),\hfill \\ H_5\hfill & =& c𝑑x\left[\frac{(u^{})^2}{20}+\frac{4u^3}{5}+\frac{80v^2}{3}\right],\hfill \\ H_7\hfill & =& c𝑑x\left[\frac{(u^{\prime \prime })^2}{3200}+\frac{9u(u^{})^2}{400}+\frac{(v^{})^2}{6}+\frac{3u^4}{50}+4uv^2\right],\hfill \\ & & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..\hfill \\ & & \mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}\hfill \end{array}$$ $`(38)`$ Here the subscripts for the conserved quantities stand for the naive conformal dimensions (i.e. $`H_1L^1,H_2L^2`$ etc.). The commutativity of the conserved quantities (i.e., $`\{H_i,H_j\}=0,i,j=1,2,4,5,7\mathrm{}..`$) can be established if we exploit the following second Hamiltonian structure associated with $`u`$ and $`v`$ fields for the classical $`W_3`$ Poisson brackets $$\begin{array}{ccc}\{u(x,t),u(y,t)\}\hfill & =& \frac{2}{c}\left[\frac{1}{6}\frac{^3}{y^3}2u(y)\frac{}{y}\frac{u}{y}\right]\delta (xy),\hfill \\ \{u(x,t),v(x,t)\}\hfill & =& \frac{2}{c}\left[\mathrm{\hspace{0.33em}3}v(y)\frac{}{y}+\frac{v}{y}\right]\delta (xy),\hfill \\ \{v(x,t),v(y,t)\}\hfill & =& \frac{3}{100c}[\frac{1}{48}\frac{^5}{y^5}+\frac{5}{4}u(y)\frac{^3}{y^3}+\frac{15}{8}\frac{u}{y}\frac{^2}{y^2}\hfill \\ & +& (\frac{9}{8}\frac{^2u}{y^2}12u^2)\frac{}{y}+(\frac{1}{4}\frac{^3u}{y^3}12u\frac{u}{y})]\delta (xy).\hfill \end{array}$$ $`(39)`$ It will be noticed that there are no conserved quantities for the conformal dimensions 3, 6, 9, 12…..( n = 0 mod 3). This is in agreement with the famous Lenard recursion relations for the commuting conserved quantities for the Boussinesq equations. These conserved quantities can be understood in terms of the generators of the infinite dimensional algebra $`W_3^{\mathrm{}}`$. For instance, if we take the Laurent mode decomposition for the $`u`$ and $`v`$ fields and consider the holomorphic and antiholomorphic parts together, the contour integration in (38) will lead to the following set of generators modulo some constant factors: $$\begin{array}{ccc}\{L_1,W_2,\mathrm{\Phi }_4,S_5\mathrm{}\mathrm{}\mathrm{}\mathrm{}\mathrm{}..\}\hfill & & \end{array}$$ $`(40)`$ where $`\mathrm{\Phi }=\frac{48}{c}(TW),S=\frac{1}{c}\left(W^2\frac{128}{3c}T^3+\frac{4}{3}(T)^2\right)`$…….etc. These generators form a Cartan subalgebra in the infinite dimensional algebra $`W_3^{\mathrm{}}`$ as they commute among themselves. In the framework of nonlinear realization method, these generators correspond to the translation generators on the infinite dimensional coset manifold. For instance, the first conserved quantity $`H_1`$ corresponds to the generator $`L_1`$ which is nothing but the space translation (momentum) generator. The rest of the conserved quantities correspond to the time evolution generators on the coset manifold. Their commutativity corresponds to the linear independence of the directions of the space $`x`$ and all the other “time” evolution parameters with one another. To summarize, we have shown that: (i) the Boussinesq equations are the embedding conditions on a 2D geodesic hypersurface in the infinite dimensional coset manifold, (ii) Miura maps are the covariant kinematical relationships among the essential fields as one goes covariantly from one coset manifold to another, (iii) Lax-pairs are the components of projections of the reduced Cartan forms along $`x`$ and $`t`$ directions of the coset manifold, (iv) Commuting conserved quantities are the translation generators on the manifold and they form a Cartan subalgebra in the infinite dimensional algebra $`W_3^{\mathrm{}}`$, (v) Commutativity of the conserved quantities are reflected in the linear independence of all the evolution directions $`(x,t,t^{}\mathrm{}..)`$ on the coset manifold.
warning/0004/hep-th0004106.html
ar5iv
text
# Proposal for non-BPS D-brane action ## 1 Introduction In the paper Sen proposed a supersymmetric invariant Dirac-Born-Infeld (DBI) action for a non-BPS D-brane. Since the non-BPS branes break all supersymmetries, it seams to be strange to construct a supersymmetric action describing this brane. However, although there is no manifest supersymmetry of the world-volume theory, we still expect the world-volume theory to be supersymmetric, with the supersymmetry realised as a spontaneously broken symmetry. From these arguments Sen showed that the action has to contain the full number of fermionic zero modes (=32), because they are fermionic Goldstone modes of the completely broken supersymmetry, while a BPS D-brane contains 16 zero modes, because it breaks one half of the supersymmetry. Sen showed that the DBI action for the non-BPS D-brane (without the presence of the tachyon) is the same as the supersymmetric action describing the BPS D-brane. This action is manifestly invariant under all space-time supersymmetries. Sen argued that the ordinary action for the BPS D-brane contains the DBI term and the Wess-Zumino (WZ) term, which are invariant under supersymmetry. But only when they both are present in the action for the D-brane, the action is invariant under the local $`\kappa `$-symmetry, which is needed for gauging away one half of the fermionic degrees of freedom, so that only 16 physical fermionic fields remain on the BPS D-brane, as should be the case for an object breaking 16 bulk supersymmetries. Sen showed that the DBI term for a non-BPS D-brane is exactly the same as the DBI term in the action of a BPS D-brane (when we suppose that other massive fields are integrated out, including the tachyon) that is invariant under supersymmetry transformations, but has no $`\kappa `$-symmetry, so that the number of fermionic degrees of freedom is 32 which is the appropriate number of fermionic Goldstone modes for an object braking bulk supersymmetry completely. Sen also showed how we could include the tachyonic field into the action. Sen proposed a form of the term expressing the interaction between the tachyon and other light fields on the world-volume of a non-BPS D-brane on the grounds of invariance under the supersymmetry and general covariance. This term has the useful property that for a constant tachyon field it is zero, so that the action for the non-BPS D-brane in that case vanishes identically. There have been many attempts to generalise Sen’s proposal for the construction of non-BPS D-branes. In a previous paper we tried to construct the action for a non-BPS Dp-brane in Type IIA theory, which was manifestly supersymmetric and also had the property that tachyon condensation in the form of a kink solution leads to the BPS D-brane in Type IIA theory. Different forms of the action for a non-BPS D-brane was suggested in . These proposals reflects the remarkable symmetry between the tachyon and other massless degrees of freedom, which was anticipated in . The action presented in was T-duality covariant, while the action presented in was not T-duality covariant, which is the main problem of that proposal. In the present paper we would like to propose yet another form of the action for a single non-BPS D-brane which will combine the virtues of all previous attempts. We propose the form of the action, which will be manifestly supersymmetry invariant, T-duality covariant and in the linear approximation the equation of motion for the tachyon will naturally have a smooth tachyon kink solution as a solution. As a result, the action for a non-BPS Dp-brane will reduce to the Dirac-Born-Infeld action for D(p-1)-brane, which together with the tachyon condensation in the Wess-Zumino (WZ) term for a non-BPS D-brane will lead to the supersymmetric action for a BPS D(p-1)-brane. We will also discuss the tachyon kink solution for the action proposed in . We will find a remarkable fact that the tachyon kink solution in the form of a piecewise function is a natural solution of the linearised equation of motion obtained from this action regardless to the form of the tachyon potential. In the conclusions, we will suggest possible extensions of this work. ## 2 Proposal for non-BPS D-brane action We start this section with recapitulating the basic facts about non-BPS D-branes in Type IIA theory, following <sup>1</sup><sup>1</sup>1For non-BPS D-brane the situation is basically the same with difference in chirality of the Majorana-Weyl fermions. We refer to for more details.. Let $`\sigma ^\mu ,\mu =0,\mathrm{}p`$ are world-volume coordinates on a D-brane. Fields living on this D-brane arise as the lightest states from the spectrum of the open string ending on this D-brane. These open strings have two CP sectors : The first one, with unit $`2\times 2`$ matrix, which corresponds to the states of the open string with the usual GSO projection $`(1)^F|\psi =|\psi `$, where $`F`$ is the world-sheet fermion number and $`|\psi `$ is a state from the Hilbert space of the open string living on a Dp-brane. The second CP sector has CP matrix $`\sigma _1`$ and contains states having opposite GSO projection $`(1)^F|\psi =|\psi `$. The massless fields living on a Dp-brane are ten components of $`X^M(\sigma ),M=0,\mathrm{},9`$ ; a $`U(1)`$ gauge field $`A(\sigma )_\mu `$ and a fermionic field $`\theta `$ with $`32`$ real components transforming as a Majorana spinor under the Lorenz group $`SO(9,1)`$. We can write $`\theta `$ as a sum of a left-handed Majorana-Weyl spinor and a right-handed Majorana-Weyl spinor: $$\theta =\theta _L+\theta _R,\mathrm{\Gamma }_{11}\theta _L=\theta _L,\mathrm{\Gamma }_{11}\theta _R=\theta _R$$ (1) All fields except $`\theta _R`$ come from the CP sector with the identity matrix, while $`\theta _R`$ comes from the sector with the $`\sigma _1`$ matrix <sup>2</sup><sup>2</sup>2Our conventions are following. $`\mathrm{\Gamma }^M`$ are $`32\times 32`$ Dirac matrices appropriate to 10d with the relation $`\{\mathrm{\Gamma }^M,\mathrm{\Gamma }^N\}=2\eta ^{MN}`$, with $`\eta ^{MN}=(1,1,\mathrm{},1)`$. For this choice of gamma matrices the massive Dirac equation is $`(\mathrm{\Gamma }^M_MM)\mathrm{\Psi }=0`$. We also introduce $`\mathrm{\Gamma }_{11}=\mathrm{\Gamma }_0\mathrm{}\mathrm{\Gamma }_9,(\mathrm{\Gamma }_{11})^2=1`$.. As Sen argued, the action for a non-BPS D-brane (without tachyon) should go to the action for BPS D-brane, when we set $`\theta _R=0`$ (we have opposite convention that ). For this reason, the action for a non-BPS D-brane in was constructed as the supersymmetric DBI action but without $`\kappa `$-symmetry so that we are not able to gauge away half of the fermionic degrees of freedom. This action thus describes a non-BPS D-brane. The next thing is to include is the effect of the tachyon. In order to get some relation between tachyon condensation and the supersymmetric D-branes, we would like to have an effective action for the massless fields and the tachyon living on the world-volume of a non-BPS D-brane. Following , the effective action for a non-BPS Dp-brane with a tachyonic field on its world-volume should has a form: $$S=C_pd^{p+1}\sigma \sqrt{det(𝒢_{\mu \nu }+2\pi \alpha ^{}_{\mu \nu })}F(T,T,\theta _L,\theta _R,𝒢,..),$$ (2) $$\mathrm{\Pi }_\mu ^M=_\mu X^M\overline{\theta }\mathrm{\Gamma }^M_\mu \theta ,𝒢_{\mu \nu }=\eta _{MN}\mathrm{\Pi }_\mu ^M\mathrm{\Pi }_\nu ^N$$ (3) and $$_{\mu \nu }=F_{\mu \nu }[\overline{\theta }\mathrm{\Gamma }_{11}\mathrm{\Gamma }_M_\mu \theta (_\nu X^M\frac{1}{2}\overline{\theta }\mathrm{\Gamma }^M_\nu \theta )(\mu \nu )].$$ (4) The constant $`C_p=\sqrt{2}T_p=\frac{2\pi \sqrt{2}}{g(4\pi ^2\alpha ^{})^{\frac{p+1}{2}}}`$ is a tension for a non-BPS Dp-brane, where $`T_p`$ is a tension for a BPS Dp-brane and $`g`$ is a string coupling constant. The function $`F`$ contains the dependence of the tachyon and its derivatives and may also depend on other world-volume and background field. We have proposed an action for a non-BPS D-brane in Type IIA theory in the form $$S=C_pd^{p+1}\sigma \sqrt{det(𝒢_{\mu \nu }+2\pi \alpha ^{}_{\mu \nu })}F(T,T,\theta _L,\theta _R,𝒢,..),$$ (5) where the function $`F`$ takes the form <sup>3</sup><sup>3</sup>3The meaning of $`\stackrel{~}{𝒢}_S^{\mu \nu }`$ will be explained latter. $$F=\left(\stackrel{~}{𝒢}_S^{\mu \nu }_\mu T_\nu T+V(T)+I_{TF}\right),$$ (6) where $`I_{FT}`$ contains interaction terms between the tachyon and fermionic fields, which was determined on the base of the supersymmetric invariance. However, this term also contains the expression $`f(T)\stackrel{~}{𝒢}_S^{\mu \nu }_\mu \overline{\theta }_R_\nu \theta _L`$, which, as was shown in , is not T-duality covariant, so that this term should not be present in the action. On the other hand, the equation of motion for the tachyon obtained from (5) do lead to the tachyon kink solution and the non-BPS D-brane reduces to the BPS D-brane of codimension one and the presence of the term cited above leads to the elimination of one half of fermionic degrees of freedom, which suggests that the resulting kink solution is a BPS D-brane. Then we argued that through tachyon condensation we have restored the $`\kappa `$-symmetry on the world-volume of the resulting D-brane. It can seam that the elimination of the term $`f(T)\stackrel{~}{𝒢}_S^{\mu \nu }_\mu \overline{\theta }_R_\mu \theta _L`$ on the grounds of T-duality covariance will lead to conclusion that through tachyon condensation we are not able to obtain BPS D-brane. However, as we will show, this is not completely true. We must also say that the term $`I_{TF}`$ contain many interaction terms with a difficult structure, while the interaction between fermions and tachyon presented in emerges in a very natural and symmetric way. This seems to tell us to follow their approach in the construction of the action for a non-BPS D-brane. In this paper we would like to propose the Dirac-Born-Infeld action for a single non-BPS D-brane in the form $$S_{DBI}=C_pd^{p+1}\sigma V(T)\sqrt{det\left(𝒢_{\mu \nu }+2\pi \alpha ^{}_{\mu \nu }+\frac{2\pi \alpha ^{}_\mu T_\nu T}{V(T)}\right)},$$ (7) where $`V(T)`$ is a tachyonic potential, which in the zeroth order approximation is equal to $$V(T)=2\pi \alpha ^{}m^2T^2/2+\lambda T^4+\frac{(2\pi \alpha ^{}m^2)^2}{16\lambda }=\lambda (T^2T_0^2)^2,$$ (8) where $`m^2=\frac{1}{2\alpha ^{}},T_0^2=\frac{2\pi \alpha ^{}m^2}{4\lambda }`$. In the following we do not need to know the explicit value of the constant $`\lambda `$. We must stress that the form of the action (7) was mainly inspired with the recent proposals where the action for non-BPS D-brane was given as $$S=C_pd^{p+1}\sigma V(T)\sqrt{det(𝒢_{\mu \nu }+2\pi \alpha ^{}_{\mu \nu }+2\pi \alpha ^{}_\mu T_\nu T)}.$$ (9) We have modified the action given above to the action (7) in order to get smooth tachyon kink solution . Than we will show that the tachyon condensation in the action (7) leads naturally to the DBI action for BPS D-brane and together with the tachyon condensation in the Wess-Zumino term for a non-BPS D-brane proposed in and generalised to the supersymmetric invariant form in $$S_{WZ}=CdTe^{2\pi \alpha ^{}},$$ (10) gives a correct description of a non-BPS D-brane in the approximation of slowly varying fields. However we will see in the next section that the equation of motion for the tachyon obtained from the linearised form of the action (9) leads to solution which is a piecewise tachyon kink solution regardless the form of the tachyon potential. In this paragraph we will discuss the properties of the action (7). The action is manifestly supersymmetric invariant, since contains the supersymmetric invariant terms , together with a natural requirement that the tachyon field is invariant under supersymmetric transformations. The action is manifestly invariant under the world-volume reparametrisation as well. The action is also T-duality covariant . This can be easily seen from the fact that the potential $`V(T)`$ does not change under T-duality transformation and the T-duality covariance of the other terms in (7) was proven in . The action (7) is equal to zero for the tachyon equal to its vacuum value $`T_0`$ which can be seen from the fact that for $`T=const`$ its derivative is equal to zero while $`V(T)0`$ for $`TT_0`$, so that the action reduces to the action anticipated by Sen for the case of a constant tachyon field $$S=C_pd^{p+1}\sigma V(T)\sqrt{det(𝒢_{\mu \nu }+2\pi \alpha ^{}_{\mu \nu })}0,\mathrm{if}TT_0.$$ (11) We can also see that for $`T=0`$, which corresponds to the $`(1)^{F_L}`$ operation that takes a non-BPS D-brane in Type IIA (IIB) theory into a BPS D-brane in Type IIB(IIA) theory, the derivative of the tachyon is zero, so that the term $`T`$ in the action (7) is equal to zero, while $`V(0)`$ is nonzero and depends on the precise form of the tachyon potential. For example, for the zeroth order approximation of the tachyon potential, $`V(T=0)`$ is equal about $`0.60`$ of the tension of a non-BPS D-brane so that the DBI action for a non-BPS D-brane in Type IIA (IIB) theory goes to the DBI action for a BPS D-brane in Type IIB(IIA) theory (of course, with appropriate modification of fermionic terms, since in Type IIB theory we have spinors of the same chirality) with the tension $$T_p=0.6T_p^c,$$ (12) where $`T_p^c=\frac{2\pi }{g(4\pi ^2\alpha ^{})^{(p+1)/2}}`$ is the correct tension for a BPS Dp-brane. In the previous equation we have used the transformation rule for the tension of the non-BPS D-brane under $`(1)^{F_L}`$ operation : $`(1)^{F_L}:C_pT_p`$. We believe that with the inclusion of the higher order corrections to the tachyon potential we get the exact result. It is also easy to see that the action given in (7) reduces into the action (9), when we neglect the higher powers of the tachyon field in the expression $`2\pi \alpha ^{}_\mu T_\nu TV(T)^1`$, because the tachyon potential must contain the constant term ensuring that the potential is equal to zero for the tachyon equal to its vacuum value. Then we have $$2\pi \alpha ^{}_\mu T_\nu TV(T)^1A2\pi \alpha ^{}_\mu T_\nu T+O(T^4),$$ (13) where $`A`$ is some constant that depends on the precise form of the tachyon potential. As was argued in , the requirement of T-duality do not precisely fix the numerical constant in front of the term $`(T)^2`$, so that the presence of constant $`A`$ do not affect the similarity with the term given in (9). As a last check we will show that in the linear approximation the action (7) reduces to the action (5) without the interaction terms $`I_{TF}`$ between the fermions and the tachyon. In fact, the interaction between tachyon and fermions is included directly in the form of the DBI action, which can be easily seen from the rewriting the determinant in (7) in the form $$det\left[(𝒢+2\pi \alpha ^{})_{\mu \nu }\right]det\left[\delta _\nu ^\mu +2\pi \alpha ^{}\stackrel{~}{𝒢}_S^{\mu \kappa }_\kappa T_\nu TV(T)^1\right],$$ (14) where $`\stackrel{~}{𝒢}^{\mu \nu }`$ is a inverse of $`𝒢+`$ and $`(\mathrm{})_S`$ means the symmetric part of a given matrix. This result follows from the fact that $`_\mu T_\nu T`$ is symmetric in the world-volume indexes. When we expand the second determinant in (14) and when we restrict ourselves to the linear approximation, then we obtain from (7) $$S_{DBI}=C_pd^{p+1}\sigma \sqrt{det(𝒢_{\mu \nu }+2\pi \alpha ^{}_{\mu \nu })}\left(V(T)+\frac{2\pi \alpha ^{}}{2}\stackrel{~}{𝒢}_S^{\mu \nu }_\mu T_\nu T\right),$$ (15) which is the same action as (5) without the fermionic terms. We can show that the equation of motion for the tachyon obtained from (15) leads naturally to the solution, which has the behaviour of a kink solution. This solution has been given earlier in and we will review this calculation. We get the equation of motion for the tachyon from the variation of (15), which give (we consider dependence of the tachyon on only one of the coordinates, $`x`$ say): $$D\left(\frac{d}{dx}\left(\frac{\delta F}{\delta _xT}\right)\frac{dF}{dT}\right)+(2\pi \alpha ^{})_\mu D\stackrel{~}{𝒢}_S^{\mu x}_xT=0,$$ (16) where $`F`$ has a form: $$F=\left(\frac{2\pi \alpha ^{}}{2}\stackrel{~}{𝒢}_S^{\mu \nu }_\mu T_\nu T+V(T)\right)$$ (17) and we have defined $$D=\sqrt{det(𝒢_{\mu \nu }+(2\pi \alpha ^{})_{\mu \nu })}.$$ (18) Firstly we will consider the first bracket in (16). The first expression in (16) gives $$2\pi \alpha ^{}_\mu (\stackrel{~}{𝒢}_S^{\mu x}_xT(x))=2\pi \alpha ^{}_\mu (\stackrel{~}{𝒢}_S^{\mu x})_xT+2\pi \alpha ^{}\stackrel{~}{𝒢}_S^{xx}_x\left(_xT\right),$$ (19) where we have used the fact that the tachyon field is a function of $`x`$ only. Since for tachyon in the form of a kink solution the first derivative is nonzero, the first term in (19) leads to the result $$\stackrel{~}{𝒢}_S^{\mu x}=const.$$ (20) Since the constant in (20) has not any physical meaning we can take solution in the form $`\stackrel{~}{𝒢}_S^{\mu x}=0,x\mu ,\stackrel{~}{𝒢}_S^{xx}=1`$ $`G_{xx}=1,G_{x\mu }=0,`$ where $`G_{\mu \nu }`$ is a inverse matrix of $`\stackrel{~}{𝒢}_S^{\mu \nu }`$ and plays the role of the natural open string metric . With using (2), the second expression in (16) gives the condition $$_xD=0_x𝒢_{\mu \nu }=_x_{\mu \nu }=0.$$ (22) When we return to (2) and use $$G_{\mu \nu }=𝒢_{\mu \nu }(2\pi \alpha ^{})^2_{\mu \kappa }𝒢^{\kappa \delta }_{\delta \nu },$$ (23) we get $$1=𝒢_{xx}(2\pi \alpha ^{})^2_{x\alpha }𝒢^{\alpha \beta }_{\beta x},$$ (24) where $`\alpha ,\beta =0,\mathrm{},p1,x=x^p`$. Since $`𝒢^{\alpha \beta }0`$, we obtain the natural solution of the previous equation in the form: $$𝒢_{xx}=1,𝒢_{x\alpha }=_{\alpha \beta }=0.$$ (25) Then we get $$det(𝒢_{\mu \nu }+(2\pi \alpha ^{})_{\mu \nu })=det\left(\begin{array}{cc}𝒢_{\alpha \beta }+(2\pi \alpha ^{})_{\alpha \beta }& 0\\ 0& 1\end{array}\right)=det(𝒢_{\alpha \beta }+(2\pi \alpha ^{})_{\alpha \beta }).$$ (26) When we combine $`\frac{dV}{dT}`$ with the second term in (19), we get the equation $$T^{\prime \prime }=\frac{1}{2\pi \alpha ^{}}\frac{dV}{dT},$$ (27) where $`T^{}=\frac{dT}{dx}`$. This equation has been solved in many textbooks (see, for example ) and we will follow their approach. The integration of the previous equation leads to $$\frac{dT}{\sqrt{V}}=\sqrt{\frac{2}{2\pi \alpha ^{}}}dx.$$ (28) This equation can be easily integrated for the potential given in (8) and we get $$T=T_0\mathrm{tanh}(\frac{m}{\sqrt{2}}x).$$ (29) With using (26), (28) the action (15) has a form $$S=C_pd^p\sigma 𝑑x\sqrt{det(𝒢_{\alpha \beta }+2\pi \alpha ^{}_{\alpha \beta })}2V(T(x)),$$ (30) where we have used $`V(T)+\frac{2\pi \alpha ^{}}{2}T^2=2V(T)`$. We can easily integrate over $`x`$ coordinate with using (22) and we get the final result $$S=T_{(p1)}d^p\sigma \sqrt{det(𝒢_{\alpha \beta }+2\pi \alpha ^{}_{\alpha \beta })},$$ (31) where the tension for D(p-1)-brane has a form $`T_{(p1)}=C_p{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x2V(T(x))=C_p2V_k{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x(1\mathrm{tanh}^2({\displaystyle \frac{m}{\sqrt{2}}}x))^2=`$ $`={\displaystyle \frac{2\pi }{g(4\pi ^2\alpha ^{})^{(p+1)/2}}}\left({\displaystyle \frac{8\sqrt{2}V_k}{3\pi }}\right)(4\pi ^2\alpha ^{})^{1/2},`$ where $`V_k=\frac{(2\pi \alpha ^{}m^2)^2}{16\lambda }`$. As was shown in , the vacuum value of the tachyon potential in the zeroth order approximation cancels about $`0.60`$ of the tension of the non-BPS D-brane, so we have the value of $`V_k`$ equal to $`V_k=0.60`$ and the previous equation gives the result $$T_{(p1)}=0.72\frac{2\pi }{g(4\pi ^2\alpha ^{})^{p/2}},$$ (33) which is in agreement with the result . We believe that the higher order correction to the potential as well as using the direct form of the action (without restriction to the linear approximation) (7) could give a correct value of the tension of a D(p-1)-brane. We must also stress that we do not obtain any constraints on the fermionic degrees of freedom. This follows from the fact that there are no interaction terms relating left-handed and right-handed spinors with the tachyon field, since these terms are not allowed through principles of T-duality covariance. However, this does not contradict the claim that tachyon condensation on the world-volume of a non-BPS Dp-brane leads to the action for BPS D(p-1)-brane, because we must also consider the tachyon condensation in the expression (10). It was shown in that the tachyon condensation in this term leads to the correct term for a BPS D(p-1)-brane. With using this result and (31) the whole action after tachyon condensation on the world-volume of a non-BPS Dp-brane has a form $$S=T_{p1}d^p\sigma \sqrt{det(𝒢_{\alpha \beta }+2\pi \alpha ^{}_{\alpha \beta })}+\mu _{p1}Ce^{2\pi \alpha ^{}}.$$ (34) When we assume that tachyon condensation leads to the correct values of D-brane tension $`T_{p1}`$ and D-brane charge $`\mu _{p1}`$, that (34) is supersymmetric action for D(p-1)-brane with $`\kappa `$-symmetry restored. ## 3 Other proposal for non-BPS D-brane action In the recent papers , the action for a non-BPS Dp-brane was proposed in the explicit form $$S=C_pd^{p+1}\sigma V(T)\sqrt{det(𝒢_{\mu \nu }+2\pi \alpha ^{}_{\mu \nu }+2\pi \alpha ^{}_\mu T_\nu T)}.$$ (35) This action is manifestly supersymmetric invariant and also T-duality covariant . This action obeys the property proposed in that for the tachyon equal to its vacuum value (this is the value of the tachyon that minimises the potential $`V(T)`$) $`T=T_0,V(T_0)=0`$ the action is equal to zero. This action incorporates in a very nice way the tachyon field and the interaction between the tachyon and the massless fields and suggests the deep symmetry between the tachyon and the other fields which was anticipated in . We would like to discuss the equation of motion for the tachyon obtained from the linearised form of the action (35). The linearised form of the action is $$S=C_pd^{p+1}\sigma \sqrt{det(𝒢_{\mu \nu }+2\pi \alpha ^{}_{\mu \nu })}\left(\frac{2\pi \alpha ^{}}{2}\stackrel{~}{𝒢}_S^{\mu \nu }_\mu T_\nu TV(T)+V(T)\right),$$ (36) from which we obtain the same equation of motion for tachyon as in (16) with function $`F`$ now given as $$F=\frac{2\pi \alpha ^{}}{2}V(T)\stackrel{~}{𝒢}_S^{\mu \nu }_\mu T_\nu T+V(T).$$ (37) The analysis of the this equation is the same as in the previous section and we obtain the same form of the constrains on the massless fields as before. Much interesting is the analysis the resulting equation for the tachyon which is $$\frac{dV}{dT}+\frac{2\pi \alpha ^{}}{2}\frac{dV}{dT}(T^{})^22\pi \alpha ^{}(VT^{})^{}=\frac{dV}{dT}+\frac{2\pi \alpha ^{}}{2}\frac{dV}{dT}(T^{})^22\pi \alpha ^{}(V^{}T^{}+VT^{\prime \prime })=0,$$ (38) where $`(\mathrm{})^{}=\frac{d}{dx}`$. We can immediately see that the solution $`T=T_0`$ is the solution of equation of motion. This result comes from the fact that $`V(T_0)=0,\frac{dV}{dT}|_{T=T_0}=0`$ and from the trivial fact that the derivative of a constant function is equal to zero. This confirms the results presented in . To obtain the other solution, we multiply the previous equation with $`T^{}`$ and we get $$V^{}+\frac{2\pi \alpha ^{}}{2}V^{}(T^{})^22\pi \alpha ^{}V^{}(T^{})^2\frac{2\pi \alpha ^{}}{2}V((T^{})^2)^{}=V^{}\frac{2\pi \alpha ^{}}{2}(V(T^{})^2)^{}=0,$$ (39) which can be easily integrated with the result $$V=\frac{2\pi \alpha ^{}}{2}VT^2+k,$$ (40) where $`k`$ is an integration constant. We determine this constant from the fact that in order to get the solution with the finite energy, the solution must approaches the vacuum value at spatial infinity, where we have $`V(T_0)=0,T^{}0`$. We immediately see that $`k=0`$. Than the next integration gives $$T=\sqrt{\frac{1}{\pi \alpha ^{}}}x,$$ (41) where this solution does not depend on the precise form of the tachyon potential. We can show that the solution of the equation of the motion is given in terms of the function $$T=\{\begin{array}{ccc}T_0,x<L& & \\ \sqrt{\frac{1}{\pi \alpha ^{}}}x,L<x<L& & \\ T_0,x>L& & \end{array},$$ (42) where the parameter $`L`$ is determined from the condition that for $`x=L`$ the tachyon field given (41) is equal to its vacuum value $$T_0=\sqrt{\frac{1}{\pi \alpha ^{}}}LL=T_0\sqrt{\pi \alpha ^{}}.$$ (43) We see that this solution in the zero slope limit $`\alpha ^{}0`$ reduces to the piecewise kink solution discussed in that depends only on the vacuum value of the tachyon field. The next calculation is the same as in the previous section. The tension of the resulting D-brane is given as $$T_{p1}=C_p_{\mathrm{}}^{\mathrm{}}𝑑xF(T)=2C_p_{\mathrm{}}^{\mathrm{}}𝑑xV(T),$$ (44) where we have used (40). Using (42) we obtain from the equation given above $$T_{p1}=2C_p_L^L𝑑xV(T)=2C_p\sqrt{\pi \alpha ^{}}_{T_0}^{T_0}𝑑TV(T),$$ (45) For the zeroth order approximation to the potential (8) we obtain the result $$T_{(p1)}=(\pi \alpha ^{})^{1/2}C_p\frac{32}{15}\lambda T_0^5=0.25\frac{(4\pi ^2\alpha ^{})^{1/2}}{\sqrt{2}}C_p=0.25T_{(p1)}^c.$$ (46) We must emphasise again that in the linear approximation this solution does not depend on the exact form of the tachyon potential, is depends only on the vacuum value of the tachyon field. However, we must stress that from these simple calculations we cannot determine the exact form of the action for a non-BPS D-brane. Perhaps more detailed calculations in the string theory could answer the question how a DBI action for a non-BPS D-brane looks like. ## 4 Conclusion In this paper we have proposed the form of the action for a non-BPS Dp-brane, which is manifestly supersymmetric invariant, T-duality covariant and in the linear approximation we have obtained through the tachyon condensation the supersymmetric action for a D(p-1)-brane with $`\kappa `$-symmetry restored. We have also discussed the tachyon kink solution obtained as a solution of the equation of motion which arises from the variation of the linearised action proposed in . We have seen the remarkable fact that we can get the solution which does not depend on the form of the tachyon potential explicitly, it is function of the tachyon vacuum value only. At present we cannot determine whether our proposal is the correct one only on the grounds of supersymmetry invariance and T-duality covariance. It seams to us that the more detailed calculations in string theory could determine the correct form of the DBI action for a non-BPS D-brane <sup>4</sup><sup>4</sup>4We thank A. A. Tseytlin for discussing this point.. It would be interesting to extend the action (7) to the non-Abelian case, following . This result could have the direct relation to the classification D-branes in K-theory . We have made some progress in this direction in , where we have tried to extend the action (5) to the non-Abelian case. It would be nice to see whether the action presented in could be modified in order to be related to the non-Abelian extension of the action (7). We hope to return to this question in the future. Acknowledgement: I would like to thank Zdeněk Kopecký and especially Rikard von Unge for conversations and critical comments.
warning/0004/nucl-th0004052.html
ar5iv
text
# NEUTRINO OSCILLATIONS AND THE SOLAR NEUTRINO PROBLEM ## 1 Introduction Part of the interest in neutrino astrophysics has to do with the fascinating interplay between nuclear and particle physics issues — e.g., whether neutrinos are massive and undergo flavor oscillations, whether they have detectable electromagnetic moments, etc. — and astrophysical phenomena, such as the clustering of matter on large scales, the processes responsible for the synthesis of nuclei, the mechanism for core-collapse supernovae, and the evolution of our sun. This summary addresses one of the oldest problems in neutrino astrophysics, the 30-year puzzle of the missing solar neutrinos. This puzzle grew out of attempts to test the standard theory of main sequence stellar evolution, but has now led to speculations about physics beyond the standard model of electroweak interactions. I will describe the work that defined the solar neutrino problem, the likelihood that its resolution is connected with massive neutrinos, and the hopes we have for future experiments. ## 2 Open Questions in Neutrino Physics The existence of the neutrino was first suggested by Wolfgang Pauli in a private letter dated December, 1930. The motivation was to solve an apparent problem with energy conservation in nuclear $`\beta `$ decay: the observable particles in the final state (the daughter nucleus and emitted electron) carried less energy than that released in the nuclear decay. Pauli suggested that an unobserved particle, the neutrino, accompanied the decay and accounted for the missing energy. A number of important developments followed Pauli’s suggestion. In 1934 Fermi suggested a theory of $`\beta `$ decay that was modeled after electromagnetism, except that there was no analog of the electromagnetic field: the interaction occurred at a point. (Apart from the missing aspect of parity violation, this was the correct reduction of today’s standard model to an effective theory.) In 1934 Bethe and Critchfield described the role of $`\beta `$ decay in thermonuclear reaction chains powering the stars $$(A,Z)(A,Z1)+e^++\nu _e$$ thus predicting that our sun produces an enormous neutrino flux. In 1956 Cowan and Reines succeeded in measuring neutrinos emitted by a reactor through the reaction $$\overline{\nu }_e+pn+e^+,$$ exploiting the positron and neutron coincidence. (The neutron was detected by $`(n,\gamma )`$ on a Cd neutron poison.) In 1957 the weak force mediating neutrino interactions was found to violate parity maximally. Later experiments found that the $`\nu _e`$ was replicated twice more in nature – the $`\nu _\mu `$ and $`\nu _\tau `$ – each accompanying a distinct charged lepton, $$\nu _ee^{}\nu _\mu \mu ^{}\nu _\tau \tau ^{}.$$ Finally all of this physics was embodied in the standard electroweak model, out of which came the prediction of a new neutral interaction mediating neutrino scattering. Despite all of this progress, a remarkable number of questions remain. We now believe neutrinos are massive, but still have no measurement of an absolute neutrino mass (only mass differences). Many models attribute the puzzle of neutrino mass — why these neutrinos are so much lighter than other standard model particles — to scales well beyond the standard model, but we lack independent experimental tools for probing these scales. We do not now the particle-antiparticle conjugation properties of neutrinos: because they carry no standard model charges, both the Dirac (distinct antiparticle) and Majorana (no distinction between particle and antiparticle) possibilities are open. An associated question is the existence of nonzero electromagnetic moments: magnetic, charge radius, anapole, and electric dipole. No nonzero moment has been measured. Finally, there are many questions about the role of neutrinos in astrophysics and cosmology. We suspect cosmic background neutrinos contribute to dark matter and may influence large-scale structure formation. However direct experimental attempts to measure background neutrinos have failed by many orders of magnitude to reach the expected density. Type II supernovae convert approximately 99% of the energy released in the infall into neutrinos of all flavors. Yet only $`\overline{\nu }_e`$s were detected from SN1987A. Supernova modelers predict that neutrinos play an essential role in the explosion mechanism and in the associated nucleosynthesis, yet there is disagreement about the success of neutrino-driven explosions. Finally, there is great interest in mounting searches for very high energy astrophysical neutrinos that might be associated with active galactic nuclei, gamma ray bursts, etc. Given all of these open questions, 70 years after Pauli’s original suggestion, it would be nice to have a few more answers. There is every indication that some answers will come with the resolution of the solar neutrino puzzle. ## 3 The Standard Solar Model Solar models trace the evolution of the sun over the past 4.7 billion years of main sequence burning, thereby predicting the present-day temperature and composition profiles of the solar core that govern neutrino production. Standard solar models (SSMs) share four basic assumptions: * The sun evolves in hydrostatic equilibrium, maintaining a local balance between the gravitational force and the pressure gradient. To describe this condition in detail, one must specify the equation of state as a function of temperature, density, and composition. * Energy is transported by radiation and convection. While the solar envelope is convective, radiative transport dominates in the core region where thermonuclear reactions take place. The opacity depends sensitively on the solar composition, particularly the abundances of heavier elements. * Thermonuclear reaction chains generate solar energy. The standard model predicts that over 98% of this energy is produced from the pp chain conversion of four protons into <sup>4</sup>He (see Fig. 1) $$4p^4\mathrm{He}+2e^++2\nu _e$$ (1) with proton burning through the CNO cycle contributing the remaining 2%. The sun is a large but slow reactor: the core temperature, $`T_c1.510^7`$ K, results in typical center-of-mass energies for reacting particles of $``$ 10 keV, much less than the Coulomb barriers inhibiting charged particle nuclear reactions. Thus reaction cross sections are small: in most cases laboratory measurements are only possible at higher energies, so that cross section data must be extrapolated to the solar energies of interest. * The model is constrained to produce today’s solar radius, mass, and luminosity. An important assumption of the standard model is that the sun was highly convective, and therefore uniform in composition, when it first entered the main sequence. It is furthermore assumed that the surface abundances of metals (nuclei with A $`>`$ 5) were undisturbed by the subsequent evolution, and thus provide a record of the initial solar metallicity. The remaining parameter is the initial <sup>4</sup>He/H ratio, which is adjusted until the model reproduces the present solar luminosity in today’s sun. The resulting <sup>4</sup>He/H mass fraction ratio is typically 0.27 $`\pm `$ 0.01, which can be compared to the big-bang value of 0.23 $`\pm `$ 0.01. Note that the sun was formed from previously processed material. The model that emerges is an evolving sun. As the core’s chemical composition changes, the opacity and core temperature rise, producing a 44% luminosity increase since the onset of the main sequence. The temperature rise governs the competition between the three cycles of the pp chain: the ppI cycle dominates below about 1.6 $`10^7`$ K; the ppII cycle between (1.7-2.3) $`10^7`$K; and the ppIII above 2.4 $`10^7`$K. The central core temperature of today’s SSM is about 1.55 $`10^7`$K. The competition between the cycles determines the pattern of neutrino fluxes. Thus one consequence of the thermal evolution of our sun is that the <sup>8</sup>B neutrino flux, the most temperature-dependent component, proves to be of relatively recent origin: the predicted flux increases exponentially with a doubling period of about 0.9 billion years. A final aspect of SSM evolution is the formation of composition gradients on nuclear burning timescales. Clearly there is a gradual enrichment of the solar core in <sup>4</sup>He, the ashes of the pp chain. Another element, <sup>3</sup>He, can be considered a catalyst for the pp chain, being produced and then consumed, and thus eventually reaching some equilibrium abundance. The timescale for equilibrium to be established as well as the final equilibrium abundance are both sharply decreasing functions of temperature, and therefore increasing functions of the distance from the center of the core. Thus a steep <sup>3</sup>He density gradient is established over time. The SSM has had some notable successes. From helioseismology the sound speed profile $`c(r)`$ has been very accurately determined for the outer 90% of the sun, and is in excellent agreement with the SSM. Such studies verify important predictions of the SSM, such as the depth of the convective zone. However the SSM is not a complete model in that it does not explain all features of solar structure, such as the depletion of surface Li by two orders of magnitude. This is usually attributed to convective processes that operated at some epoch in our sun’s history, dredging Li to a depth where burning takes place. The principal neutrino-producing reactions of the pp chain and CNO cycle are summarized in Table 1. The first six reactions produce $`\beta `$ decay neutrino spectra having allowed shapes with endpoints given by E$`{}_{}{}^{\mathrm{max}}{}_{\nu }{}^{}`$. Deviations from an allowed spectrum occur for <sup>8</sup>B neutrinos because the <sup>8</sup>Be final state is a broad resonance. The last two reactions produce line sources of electron capture neutrinos, with widths $``$ 2 keV characteristic of the temperature of the solar core. Measurements of the pp, <sup>7</sup>Be, and <sup>8</sup>B neutrino fluxes will determine the relative contributions of the ppI, ppII, and ppIII cycles to solar energy generation. As discussed above, and as later illustrations will show more clearly, this competition is governed in large classes of solar models by a single parameter, the central temperature $`T_c`$. The flux predictions of the 1998 calculations of Bahcall, Basu, and Pinsonneault (BP98) and of Brun, Turck-Chieze and Morel are included in Table 1. ## 4 Solar Neutrino Experiments and their Implications The first solar neutrino results were announced by Ray Davis Jr. and his Brookhaven collaborators in 1968, more than 30 years ago . Located deep within the Homestake Gold Mine in Lead, South Dakota, the detector consists of a 100,000 gallon tank of C<sub>2</sub>Cl<sub>4</sub>. Solar neutrinos are captured by $${}_{}{}^{37}\mathrm{Cl}(\nu ,e^{})^{37}\mathrm{Ar}.$$ As the threshold for this reaction is 0.814 MeV, the important neutrino sources are the <sup>7</sup>Be and <sup>8</sup>B reactions. The <sup>7</sup>Be neutrinos excite just the Gamow-Teller (GT) transition to the ground state, the strength of which is known from the electron capture lifetime of <sup>37</sup>Ar. The <sup>8</sup>B neutrinos can excite all bound states in <sup>37</sup>Ar, including the dominant transition to the isobaric analog state residing at an excitation energy of 4.99 MeV. The strength of excite-state GT transitions can be determined from the $`\beta `$ decay <sup>37</sup>Ca$`(\beta ^+)^{37}`$K, which is the isospin mirror reaction to <sup>37</sup>Cl$`(\nu ,e^{})^{37}`$Ar. The net result is that, for SSM fluxes, 78% of the capture rate should be due to <sup>8</sup>B neutrinos, and 15% to <sup>7</sup>Be neutrinos. The measured capture rate 2.56 $`\pm 0.16\pm 0.16`$ SNU (1 SNU = 10<sup>-36</sup> capture/atom/sec) is about one-third the SSM value. Similar radiochemical experiments were begun in January, 1990, and May, 1991, respectively, by the SAGE and GALLEX collaborations using a different target, <sup>71</sup>Ga. The special properties of this target include its low threshold and an unusually strong transition to the ground state of <sup>71</sup>Ge, leading to a large pp neutrino cross section (see Fig. 2). The experimental capture rates are $`66.6{}_{7.1}{}^{+6.8}_{4.0}^{+3.8}`$ (SAGE) and $`77.5\pm 6.2_{4.7}^{+4.3}`$ SNU (GALLEX) . The SSM prediction is about 130 SNU . Most important, since the pp flux is directly constrained by the solar luminosity in all steady-state models, there is a minimum theoretical value for the capture rate of 79 SNU, given standard model weak interaction physics. Note there are substantial uncertainties in the <sup>71</sup>Ga cross section due to <sup>7</sup>Be neutrino capture to two excited states of unknown strength. These uncertainties were greatly reduced by direct calibrations of both detectors using <sup>51</sup>Cr neutrino sources. Experiments of a different kind, Kamiokande II/III and SuperKamiokande, exploit water Cerenkov detectors to view solar neutrinos event-by-event. Solar neutrinos scatter off electrons, with the recoiling electrons producing the Cerenkov radiation that is then recorded in surrounding phototubes. Thresholds are determined by background rates; SuperKamiokande is currently operating with a trigger at approximately six MeV. The initial experiment, Kamiokande II/III, found a flux of <sup>8</sup>B neutrinos of (2.80 $`\pm 0.19\pm 0.33)10^6`$/cm<sup>2</sup>s after about a decade of measurement . Its much larger successor SuperKamiokande, with a 22.5 kiloton fiducial volume, yielded the result $`(2.42\pm 0.04\pm 0.06)10^6`$/cm<sup>2</sup>s after the first 825 days of measurements . This is about 48% of the SSM flux. This result continues to improve in accuracy. These results can be combined to limit the principal solar neutrino fluxes, under the assumption that no new particle physics distorts the spectral shape of the pp and <sup>8</sup>B neutrinos. One finds $`\varphi (pp)`$ $``$ $`0.9\varphi ^{\mathrm{SSM}}(pp)`$ $`\varphi (^7\mathrm{Be})`$ $``$ $`0`$ $`\varphi (^8\mathrm{B})`$ $``$ $`0.47\varphi ^{\mathrm{SSM}}(^8\mathrm{B}).`$ (2) A reduced <sup>8</sup>B neutrino flux can be produced by lowering the central temperature of the sun somewhat, as $`\varphi (^8`$B)$`T_c^{18}`$. However, such an adjustment, either by varying the parameters of the SSM or by adopting some nonstandard physics, tends to push the $`\varphi (^7`$Be)/$`\varphi (^8`$B) ratio to higher values rather than the low one of eq. (12), $$\frac{\varphi (^7\mathrm{Be})}{\varphi (^8\mathrm{B})}T_c^{10}.$$ (3) Thus the observations seem difficult to reconcile with plausible solar model variations: one observable ($`\varphi (^8`$B)) requires a cooler core while a second, the ratio $`\varphi (^7`$Be)/$`\varphi (^8`$B), requires a hotter one. This physics was nicely illustrated by Castellani et al. . These authors generated a series of nonstandard models by changing the S-factor for the p+p reaction, modifying the core metalicity, introducing weakly interacting massive particles as a new mechanism for energy transport, etc. The resulting core temperature $`T_c`$ and neutrino fluxes were then determined, and the latter were plotted as a function of the former. The pattern that emerges is striking (see Fig. 3): parameter variations producing the same value of $`T_c`$ produce remarkably similar fluxes. Thus $`T_c`$ provides an excellent one-parameter description of standard model perturbations. Figure 3 also illustrates the difficulty of producing a low ratio of $`\varphi (^7`$Be)/$`\varphi (^8`$B) when $`T_c`$ is reduced. This result is consistent with our earlier argument and shows that even extreme changes in quantities like the metallicity, opacities, or solar age, cannot produce the pattern of fluxes deduced from experiment (eq. (2)). Is it possible to change the solar model in a way that reduces the <sup>7</sup>Be/<sup>8</sup>B neutrino flux ratio? It is appears the answer is no in models where the nuclear reactions burn in equilibrium with plausible cross sections. However Cumming and Haxton pointed out that a possible exception was a nonequilibrium model in which the solar core is mixed on the timescale of <sup>3</sup>He evolution, about $`10^7`$ years. Thus the pp chain is prevented from reaching equilibrium. This suggestion has some physical plausibility because it allows the sun to burn more efficiently, with a cooler core and enhanced ppI terminations. The SSM <sup>3</sup>He profile is known to be overstable, as was first discussed by Dilke and Gough . Also the possibility of a persistent convective core powered by the <sup>3</sup>He gradient has been discussed in the literature. (The SSM core is convective for about $`10^8`$ years because of out-of-equilibrium burning of the CNO cycle.) A strong argument against a mixed core was offered by Bahcall et al. , who showed that homogenizing the core of the SSM led to very large changes in the helioseismology. While this is a sobering result, this test was not done in a self-consistent model. However, there is work in progress to test the helioseismology of a more realistic mixed core model – one where the <sup>4</sup>He content of the core, the temperature gradient, the nuclear reaction rates, and the luminosity are handled consistently . If the helioseismology remains unacceptable, this will rule out the only solar model conjecture for producing a reduced <sup>7</sup>Be/<sup>8</sup>B flux ratio, which is necessary to produce fluxes closer to those observed. However, there is a popular argument showing that no SSM change can completely remove the discrepancy with experiment: if one assumes undistorted neutrino spectra, no combination of pp, <sup>7</sup>Be, and <sup>8</sup>B neutrino fluxes fits the experimental results well . In fact, in an unconstrained fit, the required <sup>7</sup>Be flux is unphysical, negative by about 2.5$`\sigma `$. This is clearly a strong hint that one should look elsewhere for a solution! The remaining possibility is new neutrino physics. Suggested particle physics solutions of the solar neutrino problem include neutrino oscillations, neutrino decay, neutrino magnetic moments, and weakly interacting massive particles. Among these, the Mikheyev-Smirnov-Wolfenstein effect — neutrino oscillations enhanced by matter interactions — is widely regarded as perhaps the most plausible. ## 5 Neutrino Oscillations One odd feature of particle physics is that neutrinos, which are not required by any symmetry to be massless, nevertheless must be much lighter than any of the other known fermions. For instance, the current limit on the $`\overline{\nu }_e`$ mass is $`\text{ }<`$ 5 eV. The standard model requires neutrinos to be massless, but the reasons are not fundamental. Dirac mass terms $`m_D`$, analogous to the mass terms for other fermions, cannot be constructed because the model contains no right-handed neutrino fields. Neutrinos can also have Majorana mass terms $$\overline{\nu _L^c}m_L\nu _L\mathrm{and}\overline{\nu _R^c}m_R\nu _R$$ (4) where the subscripts $`L`$ and $`R`$ denote left- and right-handed projections of the neutrino field $`\nu `$, and the superscript $`c`$ denotes charge conjugation. The first term above is constructed from left-handed fields, but can only arise as a nonrenormalizable effective interaction when one is constrained to generate $`m_L`$ with the doublet scalar field of the standard model. The second term is absent from the standard model because there are no right-handed neutrino fields. None of these standard model arguments carries over to the more general, unified theories that theorists believe will supplant the standard model. In the enlarged multiplets of extended models it is natural to characterize the fermions of a single family, e.g., $`\nu _e`$, e, u, d, by the same mass scale $`m_D`$. Small neutrino masses are then frequently explained as a result of the Majorana neutrino masses. In the seesaw mechanism, $$M_\nu \left(\begin{array}{cc}0& m_D\\ m_D^T& m_R\end{array}\right).$$ (5) Diagonalization of this matrix produces one light neutrino, $`m_{\mathrm{light}}m_D(\frac{m_D}{m_R})`$, and one unobservably heavy, $`m_{\mathrm{heavy}}m_R`$. The factor ($`m_D`$/$`m_R`$) is the needed small parameter that accounts for the distinct scale of neutrino masses. The masses for the $`\nu _e,\nu _\mu `$, and $`\nu _\tau `$ are then related to the squares of the corresponding quark masses $`m_u`$, $`m_c`$, and $`m_t`$. Taking $`m_R10^{16}`$ GeV, a typical grand unification scale for models built on groups like SO(10), the seesaw mechanism gives the crude relation $$m_{\nu _e}:m_{\nu _\mu }:m_{\nu _\tau }210^{12}:210^7:310^3\mathrm{eV}.$$ (6) The fact that solar neutrino experiments can probe small neutrino masses, and thus provide insight into possible new mass scales $`m_R`$ that are far beyond the reach of direct accelerator measurements, has been an important theme of the field. Consider for simplicity just two neutrino flavors. The states of definite mass are the states that diagonalize the free Hamiltonian. Similarly the weak interaction eigenstates are the states of definite flavor, that is, the $`\nu _e`$ accompanies the positron in $`\beta `$ decay, and the $`\nu _\mu `$ accompanies the muon. There is every reason to assume that these two bases are not coincident, but instead are related by a nontrivial rotation, $`|\nu _e`$ $`=`$ $`\mathrm{cos}\theta _v|\nu _1+\mathrm{sin}\theta _v|\nu _2`$ $`|\nu _\mu `$ $`=`$ $`\mathrm{sin}\theta _v|\nu _1+\mathrm{cos}\theta _v|\nu _2`$ (7) where $`\theta _v`$ is the (vacuum) mixing angle. Consider a $`\nu _e`$ produced at time $`t`$=0 as a momentum eigenstate $$|\nu (t=0)=|\nu _e=\mathrm{cos}\theta _v|\nu _1+\mathrm{sin}\theta _v|\nu _2.$$ (8) The resulting probability for measuring a $`\nu _e`$ downstream then depends on $`\delta m^2=m_2^2m_1^2`$, $`P_{\nu _e}(t)`$ $`=`$ $`|\nu _e|\nu (t)|^2`$ (9) $`=`$ $`1\mathrm{sin}^22\theta _v\mathrm{sin}^2\left({\displaystyle \frac{\delta m^2t}{4k}}\right)1{\displaystyle \frac{1}{2}}\mathrm{sin}^22\theta _v`$ where the limit on the right is appropriate for large $`t`$. (When one properly describes the neutrino state as a wave packet, the large-distance behavior follows from the eventual separation of the mass eigenstates.) If the the oscillation length $$L_o=\frac{4\pi \mathrm{}cE}{\delta m^2c^4}$$ (10) is comparable to or shorter than one astronomical unit, a reduction in the solar $`\nu _e`$ flux would be expected in terrestrial neutrino oscillations. The suggestion that the solar neutrino problem could be explained by neutrino oscillations was first made by Pontecorvo in 1958, who pointed out the analogy with $`K_0\overline{K}_0`$ oscillations. From the point of view of particle physics, the sun is a marvelous neutrino source. The neutrinos travel a long distance and have low energies ($``$ 1 MeV), implying a sensitivity to $$\delta m^2\text{ }>10^{12}eV^2.$$ (11) In the seesaw mechanism, $`\delta m^2m_2^2`$, so neutrino masses as low as $`m_210^6`$ eV could be probed. In contrast, terrestrial oscillation experiments with accelerator or reactor neutrinos are typically limited to $`\delta m^2\text{ }>0.1`$ eV<sup>2</sup>. (Planned long-baseline experiments, though, will soon push below 0.01 eV<sup>2</sup>.) From the expressions above one expects vacuum oscillations to affect all neutrino species equally, if the oscillation length is small compared to an astronomical unit. This is somewhat in conflict with the data, as we have argued that the <sup>7</sup>Be neutrino flux is quite suppressed. Furthermore, there is a weak theoretical prejudice that $`\theta _v`$ should be small, like the Cabibbo angle. The first objection, however, can be circumvented in the case of “just so” oscillations where the oscillation length is comparable to one astronomical unit. In this case the oscillation probability becomes sharply energy dependent, and one can choose $`\delta m^2`$ to preferentially suppress one component (e.g., the monochromatic <sup>7</sup>Be neutrinos). This scenario has been explored by several groups and remains an interesting possibility. However, the requirement of large mixing angles remains. ## 6 The Mikheyev-Smirnov-Wolfenstein Mechanism In order to include matter effects, we first consider vacuum oscillations for the more general case $$|\nu (t=0)=a_e(t=0)|\nu _e+a_\mu (t=0)|\nu _\mu .$$ (12) from which one easily calculates $$i\frac{d}{dx}\left(\begin{array}{c}a_e\\ a_\mu \end{array}\right)=\frac{1}{4E}\left(\begin{array}{c}\delta m^2\mathrm{cos}2\theta _v\delta m^2\mathrm{sin}2\theta _v\\ \delta m^2\mathrm{sin}2\theta _v\delta m^2\mathrm{cos}2\theta _v\end{array}\right)\left(\begin{array}{c}a_e\\ a_\mu \end{array}\right).$$ (13) We have equated $`x=t,`$ that is, set $`c`$ = 1. Mikheyev and Smirnov showed in 1985 that the density dependence of the neutrino effective mass, a phenomenon first discussed by Wolfenstein in 1978, could greatly enhance oscillation probabilities: a $`\nu _e`$ is adiabatically transformed into a $`\nu _\mu `$ as it traverses a critical density within the sun. It became clear that the sun was not only an excellent neutrino source, but also a natural regenerator for cleverly enhancing the effects of flavor mixing. While the original work of Mikheyev and Smirnov was numerical, their phenomenon was soon understood analytically as a level-crossing problem. The vacuum oscillation evolution equation changes in the presence of matter to $$i\frac{d}{dx}\left(\begin{array}{c}a_e\\ a_\mu \end{array}\right)=\frac{1}{4E}\left(\begin{array}{c}2E\sqrt{2}G_F\rho (x)\delta m^2\mathrm{cos}2\theta _v\delta m^2\mathrm{sin}2\theta _v\\ \delta m^2\mathrm{sin}2\theta _v2E\sqrt{2}G_F\rho (x)+\delta m^2\mathrm{cos}2\theta _v\end{array}\right)\left(\begin{array}{c}a_e\\ a_\mu \end{array}\right)$$ (14) where G<sub>F</sub> is the weak coupling constant and $`\rho (x)`$ the solar electron density. The new contribution to the diagonal elements, $`2E\sqrt{2}G_F\rho (x)`$, represents the effective contribution to $`M_\nu ^2`$ that arises from neutrino-electron scattering. The indices of refraction of electron and muon neutrinos differ because the former scatter by charged and neutral currents, while the latter have only neutral current interactions. The difference in the forward scattering amplitudes determines the density-dependent splitting of the diagonal elements of the new matter equation. It is helpful to rewrite this equation in a basis consisting of the light and heavy local mass eigenstates (i.e., the states that diagonalize the right-hand side of the equation), $`|\nu _L(x)`$ $`=`$ $`\mathrm{cos}\theta (x)|\nu _e\mathrm{sin}\theta (x)|\nu _\mu `$ $`|\nu _H(x)`$ $`=`$ $`\mathrm{sin}\theta (x)|\nu _e+\mathrm{cos}\theta (x)|\nu _\mu .`$ (15) The local mixing angle is defined by $`\mathrm{sin}2\theta (x)`$ $`=`$ $`{\displaystyle \frac{\mathrm{sin}2\theta _v}{\sqrt{X^2(x)+\mathrm{sin}^22\theta _v}}}`$ $`\mathrm{cos}2\theta (x)`$ $`=`$ $`{\displaystyle \frac{X(x)}{\sqrt{X^2(x)+\mathrm{sin}^22\theta _v}}}`$ (16) where $`X(x)=2\sqrt{2}G_F\rho (x)E/\delta m^2\mathrm{cos}2\theta _v`$. Thus $`\theta (x)`$ ranges from $`\theta _v`$ to $`\pi /2`$ as the density $`\rho (x)`$ goes from 0 to $`\mathrm{}`$. If we define $$|\nu (x)=a_H(x)|\nu _H(x)+a_L(x)|\nu _L(x),$$ (17) the neutrino propagation can be rewritten in terms of the local mass eigenstates $$i\frac{d}{dx}\left(\begin{array}{c}a_H\\ a_L\end{array}\right)=\left(\begin{array}{cc}\lambda (x)& i\alpha (x)\\ i\alpha (x)& \lambda (x)\end{array}\right)\left(\begin{array}{c}a_H\\ a_L\end{array}\right)$$ (18) with the splitting of the local mass eigenstates determined by $$2\lambda (x)=\frac{\delta m^2}{2E}\sqrt{X^2(x)+\mathrm{sin}^22\theta _v}$$ (19) and with mixing of these eigenstates governed by the density gradient $$\alpha (x)=\left(\frac{E}{\delta m^2}\right)\frac{\sqrt{2}G_F\frac{d}{dx}\rho (x)\mathrm{sin}2\theta _v}{X^2(x)+\mathrm{sin}^22\theta _v}.$$ (20) The results above are quite interesting: the local mass eigenstates diagonalize the matrix if the density is constant. In such a limit, the problem is no more complicated than our original vacuum oscillation case, although our mixing angle is changed because of the matter effects. But if the density is not constant, the mass eigenstates in fact evolve as the density changes. This is the crux of the MSW effect. Note that the splitting achieves its minimum value, $`\frac{\delta m^2}{2E}\mathrm{sin}2\theta _v`$, at a critical density $`\rho _c=\rho (x_c)`$ $$2\sqrt{2}EG_F\rho _c=\delta m^2\mathrm{cos}2\theta _v$$ (21) that defines the point where the diagonal elements of the original flavor matrix cross. Our local-mass-eigenstate form of the propagation equation can be trivially integrated if the splitting of the diagonal elements is large compared to the off-diagonal elements, $$\gamma (x)=\left|\frac{\lambda (x)}{\alpha (x)}\right|=\frac{\mathrm{sin}^22\theta _v}{\mathrm{cos}2\theta _v}\frac{\delta m^2}{2E}\frac{1}{|\frac{1}{\rho _c}\frac{d\rho (x)}{dx}|}\frac{[X(x)^2+\mathrm{sin}^22\theta _v]^{3/2}}{\mathrm{sin}^32\theta _v}1,$$ (22) a condition that becomes particularly stringent near the crossing point, $$\gamma _c=\gamma (x_c)=\frac{\mathrm{sin}^22\theta _v}{\mathrm{cos}2\theta _v}\frac{\delta m^2}{2E}\frac{1}{\left|\frac{1}{\rho _c}\frac{d\rho (x)}{dx}|_{x=x_c}\right|}1.$$ (23) The resulting adiabatic electron neutrino survival probability , valid when $`\gamma _c1`$, is $$P_{\nu _e}^{\mathrm{adiab}}=\frac{1}{2}+\frac{1}{2}\mathrm{cos}2\theta _v\mathrm{cos}2\theta _i$$ (24) where $`\theta _i=\theta (x_i)`$ is the local mixing angle at the density where the neutrino was produced. The physical picture behind this derivation is illustrated in Fig. 4. One makes the usual assumption that, in vacuum, the $`\nu _e`$ is almost identical to the light mass eigenstate, $`\nu _L(0)`$, i.e., $`m_1<m_2`$ and $`\mathrm{cos}\theta _v`$ 1. But as the density increases, the matter effects make the $`\nu _e`$ heavier than the $`\nu _\mu `$, with $`\nu _e\nu _H(x)`$ as $`\rho (x)`$ becomes large. That is, the mixing angle at high density rotates to $`\pi /2`$. The special property of the sun is that it produces $`\nu _e`$s at high density that then propagate to the vacuum where they are measured. The adiabatic approximation tells us that if initially $`\nu _e\nu _H(x)`$, the neutrino will remain on the heavy mass trajectory provided the density changes slowly. That is, if the solar density gradient is sufficiently gentle, the neutrino will emerge from the sun as the heavy vacuum eigenstate, $`\nu _\mu `$. This guarantees nearly complete conversion of $`\nu _e`$s into $`\nu _\mu `$s, producing a flux that cannot be detected by the Homestake or SAGE/GALLEX detectors. But this does not explain the curious pattern of partial flux suppressions coming from the various solar neutrino experiments. The key to this is the behavior when $`\gamma _c\text{ }<`$ 1. Our expression for $`\gamma (x)`$ shows that the critical region for nonadiabatic behavior occurs in a narrow region (for small $`\theta _v`$) surrounding the crossing point, and that this behavior is controlled by the derivative of the density. This suggests an analytic strategy for handling nonadiabatic crossings: one can replace the true solar density by a simpler (integrable!) two-parameter form that is constrained to reproduce the true density and its derivative at the crossing point $`x_c`$. Two convenient choices are the linear $`(\rho (x)=a+bx)`$ and exponential $`(\rho (x)=ae^{bx})`$ profiles. As the density derivative at $`x_c`$ governs the nonadiabatic behavior, this procedure should provide an accurate description of the hopping probability between the local mass eigenstates when the neutrino traverses the crossing point. The initial and ending points $`x_i`$ and $`x_f`$ for the artificial profile are then chosen so that $`\rho (x_i)`$ is the density where the neutrino was produced in the solar core and $`\rho (x_f)=0`$ (the solar surface), as illustrated in in Fig. 5. Since the adiabatic result ($`P_{\nu _e}^{\mathrm{adiab}}`$) depends only on the local mixing angles at these points, this choice builds in that limit. But our original flavor-basis equation can then be integrated exactly for linear and exponential profiles, with the results given in terms of parabolic cylinder and Whittaker functions, respectively. That result can be simplified further by observing that the nonadiabatic region is generally confined to a narrow region around $`x_c`$, away from the endpoints $`x_i`$ and $`x_f`$. We can then extend the artificial profile to $`x=\pm \mathrm{}`$, as illustrated by the dashed lines in Fig. 5. As the neutrino propagates adiabatically in the unphysical region $`x<x_i`$, the exact soluation in the physical region can be recovered by choosing the initial boundary conditions $`a_L(\mathrm{})`$ $`=`$ $`a_\mu (\mathrm{})=\mathrm{cos}\theta _ie^{i_{\mathrm{}}^{x_i}\lambda (x)𝑑x}`$ $`a_H(\mathrm{})`$ $`=`$ $`a_e(\mathrm{})=\mathrm{sin}\theta _ie^{i_{\mathrm{}}^{x_i}\lambda (x)𝑑x}.`$ (25) That is, $`|\nu (\mathrm{})`$ will then adiabatically evolve to $`|\nu (x_i)=|\nu _e`$ as $`x`$ goes from $`\mathrm{}`$ to $`x_i`$. The unphysical region $`x>x_f`$ can be handled similarly. With some algebra a simple generalization of the adiabatic result emerges that is valid for all $`\delta m^2/E`$ and $`\theta _v`$ $$P_{\nu _e}=\frac{1}{2}+\frac{1}{2}\mathrm{cos}2\theta _v\mathrm{cos}2\theta _i(12P_{\mathrm{hop}})$$ (26) where P<sub>hop</sub> is the Landau-Zener probability of hopping from the heavy mass trajectory to the light trajectory on traversing the crossing point. For the linear approximation to the density , $$P_{\mathrm{hop}}^{\mathrm{lin}}=e^{\pi \gamma _c/2}.$$ (27) As it must by our construction, $`P_{\nu _e}`$ reduces to P$`{}_{\nu _e}{}^{}{}_{}{}^{\mathrm{adiab}}`$ for $`\gamma _c`$ 1. When the crossing becomes nonadiabatic (e.g., $`\gamma _c1`$ ), the hopping probability goes to 1, allowing the neutrino to exit the sun on the light mass trajectory as a $`\nu _e`$, i.e., no conversion occurs. Thus there are two conditions for strong conversion of solar neutrinos: there must be a level crossing (that is, the solar core density must be sufficient to render $`\nu _e\nu _H(x_i)`$ when it is first produced) and the crossing must be adiabatic. The first condition requires that $`\delta m^2/E`$ not be too large, and the second $`\gamma _c\text{ }>`$ 1. The combination of these two constraints, illustrated in Fig. 6, defines a triangle of interesting parameters in the $`\frac{\delta m^2}{E}\mathrm{sin}^22\theta _v`$ plane, as Mikheyev and Smirnov first found. A remarkable feature of this triangle is that strong $`\nu _e\nu _\mu `$ conversion can occur for very small mixing angles $`(\mathrm{sin}^22\theta 10^3`$), unlike the vacuum case. One can envision superimposing on Fig. 6 the spectrum of solar neutrinos, plotted as a function of $`\frac{\delta m^2}{E}`$ for some choice of $`\delta m^2`$. Since Davis sees some solar neutrinos, the solutions must correspond to the boundaries of the triangle in Fig. 6. The horizontal boundary indicates the maximum $`\frac{\delta m^2}{E}`$ for which the sun’s central density is sufficient to cause a level crossing. If a spectrum properly straddles this boundary, we obtain a result consistent with the Homestake experiment in which low energy neutrinos (large 1/E) lie above the level-crossing boundary (and thus remain $`\nu _e`$’s), but the high-energy neutrinos (small 1/E) fall within the unshaded region where strong conversion takes place. Thus such a solution would mimic nonstandard solar models in that only the <sup>8</sup>B neutrino flux would be strongly suppressed. The diagonal boundary separates the adiabatic and nonadiabatic regions. If the spectrum straddles this boundary, we obtain a second solution in which low energy neutrinos lie within the conversion region, but the high-energy neutrinos (small 1/E) lie below the conversion region and are characterized by $`\gamma 1`$ at the crossing density. (Of course, the boundary is not a sharp one, but is characterized by the Landau-Zener exponential). Such a nonadiabatic solution is quite distinctive since the flux of pp neutrinos, which is strongly constrained in the standard solar model and in any steady-state nonstandard model by the solar luminosity, would now be sharply reduced. Finally, one can imagine “hybrid” solutions where the spectrum straddles both the level-crossing (horizontal) boundary and the adiabaticity (diagonal) boundary for small $`\theta `$, thereby reducing the <sup>7</sup>Be neutrino flux more than either the pp or <sup>8</sup>B fluxes. What are the results of a careful search for MSW solutions satisfying the Homestake, Kamiokande/SuperKamiokande, and SAGE/GALLEX constraints? One solution, corresponding to a region surrounding $`\delta m^2610^6`$eV<sup>2</sup> and $`\mathrm{sin}^22\theta _v610^3`$, is the hybrid case described above. It is commonly called the small-angle solution. A second, large-angle solution exists, corresponding to $`\delta m^210^5`$eV<sup>2</sup> and $`\mathrm{sin}^22\theta _v`$ 0.6. (Variations on these solutions include oscillations to sterile neutrinos, oscillations modified by regeneration as neutrinos pass through the earth (day/night effects), and of course vacuum oscillations.) These solutions can be distinguished by their characteristic distortions of the solar neutrino spectrum. The survival probabilities $`P_{\nu _e}^{\mathrm{MSW}}`$(E) for the small- and large-angle parameters given above are shown as a function of E in Fig. 7. The MSW mechanism provides a natural explanation for the pattern of observed solar neutrino fluxes. While it requires profound new physics, both massive neutrinos and neutrino mixing are expected in extended models. The small-angle solution corresponds to $`\delta m^210^5`$ eV<sup>2</sup>, and thus is consistent with $`m_2`$ few $`10^3`$ eV. This is a typical $`\nu _\tau `$ mass in models where $`m_Rm_{\mathrm{GUT}}`$. This mass is also reasonably close to atmospheric neutrino values. On the other hand, if it is the $`\nu _\mu `$ participating in the oscillation, this gives $`m_R10^{12}`$ GeV and predicts a heavy $`\nu _\tau `$ 10 eV. Such a mass is of great interest cosmologically as it would have consequences for supernova physics, the dark matter problem, and the formation of large-scale structure. There are many interesting elaborations of the MSW effect not discussed here, but treated in many papers: spin-flavor oscillations induced by the solar magnetic field (the mass difference between $`\nu _e^L`$ and a sterile $`\nu _\mu ^R`$ is compensated by the matter effects); oscillations induced by density fluctuations ; “stochastic depolarization” effects in large random magnetic fields ; etc. There are also interesting effects associated with three neutrinos: if the solar neutrino problem is due to MSW $`\nu _e\nu _\mu `$ oscillations, one might expect a $`\nu _e\nu _\tau `$ crossing at still higher densities. This has led to many interesting speculations about the role of the MSW mechanism in supernova explosions and in supernova nucleosynthesis, and about the possibility that $`\nu _e\nu _\tau `$ oscillations governed by small mixing angles might be best probed using the supernova neutrino flux. ## 7 Other Neutrino Mass Evidence and Implications The solar neutrino problem is not the most compelling evidence for neutrino mass. SuperKamiokande’s analysis of the ratio of muon-like to electron-like atmospheric neutrino events confirmed that a dramatic anomaly exists. The quantity studied is $$R=\frac{(N_\mu /N_e)_{DATA}}{(N_\mu /N_e)_{MC}},$$ (28) the measured ratio of muon-like to electron-like neutrino events normalized to the expected ratio, based on Monte Carlo calculations of the production and interaction of atmospheric neutrinos. The SuperKamiokande results are $$R=0.63\pm 0.03(\mathrm{stat})\pm 0.05(\mathrm{syst})$$ (29) for sub-GeV events which were fully contained in the detector and $$R=0.65\pm 0.05(\mathrm{stat})\pm 0.08(\mathrm{syst})$$ (30) for fully- and partially-contained multi-GeV events. The results for $`R`$ are consistent among the four largest detectors used for atmospheric neutrinos (SuperK, Soudan II, IMB, Kamiokande). While this suggests neutrino oscillations, even stronger evidence for new physics comes from measurements of $`R`$ as a function of the zenith angle, $`\mathrm{\Theta }`$, between the vertical and neutrino direction. A down-going neutrino ($`\mathrm{\Theta }0^o`$) travels through the atmosphere above the detector (a distance of about 20 km), whereas an up-going neutrino ($`\mathrm{\Theta }180^o`$) has traveled through the entire Earth (a distance of about 13000 km). Hence a measurement of number of neutrinos as a function of the zenith angle yields information about their numbers as a function of the distance traveled. The zenith angle dependence of the electron and muon fluxes is shown in Fig. 8, plotted as a function of the reconstructed $`L/E_\nu `$. The muon neutrino flux drops with increasing distance, while the electron neutrino flux is approximately constant. This behavior is consistent with $`\nu _\mu \nu _\tau `$ oscillations. Measurements of up-going muons at Kamiokande and MACRO , produced by high energy muon neutrino interactions in the rock below the detectors, show a similar deficit. It is often pointed out that such results provide convincing evidence because the up/down difference in R is essentially self-normalizing: only weak geomagnetic effects are expected to break the isotropy of cosmic ray interactions in the atmosphere. Explanations of the SuperK anomaly in terms of $`\nu _\mu \nu _e`$ oscillations conflict with reactor oscillation limits, while the explanation $`\nu _\mu \nu _{sterile}`$ is disfavored in fits. The preferred solution is $`\nu _\mu \nu _\tau `$ characterized by $$\delta m_{23}^2(4\pm 2)10^3eV^2$$ $$0.8\text{ }<\mathrm{sin}^22\theta \text{ }<1.0.$$ The best fit mixing angle is approximately maximal, an intriguing and surprising result. One can take the the square root of the atmospheric $`\delta m^2`$ to obtain $`0.06`$ eV, a minimum value for a neutrino mass. This immediately establishes a lower bound on the neutrino contribution to dark matter of about 0.3% the closure density, a value not too different from the mass evident in visible stars, a remarkable result . There is one other indication of neutrino mass, the positive signal for $`\overline{\nu }_\mu \overline{\nu }_e`$ oscillations seen by the LSND group in a beam-stop experiment at Los Alamos . The experiment uses a 52,000 gallon tank of mineral oil and liquid scintillator, instrumented with 1220 phototubes. A neutrino event $`\overline{\nu }_e+pn+e^+`$ is detected by the coincidence of the positron and subsequent 2.2 MeV $`\gamma `$ ray from neutron capture, $`n+pd+\gamma `$. The signal is consistent with $`\overline{\nu }_\mu \overline{\nu }_e`$ oscillations in a narrow band that includes the ranges $`\delta m^2`$ 0.2 - 2.0 eV<sup>2</sup> and $`\mathrm{sin}^22\theta 0.030.003`$. A similar experiment at the Rutherford Laboratory, KARMEN, sees no oscillations, but has lower sensitivity . A recent combined analysis of the two experiments lowers the confidence level of the oscillation claim, but finds a parameter region consistent with both experiments. An improved experiment at Fermilab has been approved and should yield results in 2002. These results have some interesting implications. For example, we discussed the quadratic seesaw relation earlier, $`m_{\mathrm{light}}\frac{m_D^2}{m_R}`$. If we use the atmospheric $`\delta m^2`$ as a rough guide to the $`\nu _\tau `$ mass (or more correctly the $`\nu _3`$ mass, given the large atmospheric neutrino mixing angle), $`\nu _30.1`$ eV, and adopt for $`m_D`$ the corresponding third-generation quark mass, $`m_Dm_{top}`$ 180 GeV, one obtains $`m_R0.310^{15}`$ GeV. This is a value reasonable close to the supersymmetric grand unified scale of $`10^{16}`$ GeV, a startling result. It has inspired some to hope that current neutrino results are giving us our first glimpse of physics at the GUT scale. One puzzling aspect of atmospheric, solar, and LSND neutrino results is that they require three independent $`\delta m^2`$s. That is, they do not respect the relation $$\delta m_{21}^2+\delta m_{13}^2+\delta m_{32}^2=0.$$ (31) Thus either one of more of the experiments must be attributed to some phenomenon other than neutrino oscillations, or a fourth neutrino is required. That neutrino must be sterile to avoid constraints imposed by the known width of the $`Z_0`$. ## 8 Outlook The argument that the solar neutrino problem is due to neutrino oscillations is, in a sense, circumstantial: this conclusion is derived from combining several experiments, no one of which requires new particle physics. There is no direct observation of new physics analogous to the zenith angle dependence of the SuperKamiokande atmospheric results. For this reason there is great interest in a new experiment now taking data in the Creighton nickel mine in Sudbury, Ontario, 6800 feet below the surface. The Sudbury Neutrino Observatory (SNO) has a central acrylic vessel filled with one kiloton of very pure (99.92%) heavy water, surrounded by a shield of 7.5 kilotons of ordinary water. SNO can detect electron neutrinos through the charged current reaction $$\nu _e+dp+p+e^{}$$ (32) The Cerenkov light from the outgoing electron is then recorded in the array of 9800 phototubes that surround SNO’s central vessel. The spectrum of produced electrons is quite hard, making reconstruction of the energy of the $`\nu _e`$ easier than in the case of neutrino-electron elastic scattering. Thus the experimenters may be able to detect distortions of the neutrino spectrum resulting from the MSW effect. SNO will also study the neutral current reaction $$\nu _x(\overline{\nu }_x)+d\nu _x(\overline{\nu }_x)+p+n$$ (33) by detecting the produced neutron either through $`(n,\gamma )`$ reactions on salt dissolved in the heavy water or in <sup>3</sup>He proportional counters. In this way the experimenters will obtain an integral measurement of the flux of active neutrinos, independent of flavor. Thus a neutral current signal clearly larger than the corresponding $`\nu _e`$ signal would show that heavy-flavor neutrinos comprise a portion of the solar neutrino flux, providing definite proof of new physics. The SNO collaboration is expected to make its first announcement of results for the charged current reaction as early as summer, 2000. In future years, as the SNO neutral and charged current results become precise and as SuperKamiokande continues to amass data, the nature of the solar neutrino problem should become much clearer. The hope is that these results, in combination with other solar neutrino results (Borexino, GNO, iodine), with new atmospheric and (possibly) supernova neutrino measurements, and with precision tests of oscillations at accelerators and reactors, will allow us to completely characterize the neutrino mass matrix, providing a window on physics well beyond the standard model. This work was supported in part by the US Department of Energy. ## References
warning/0004/hep-th0004174.html
ar5iv
text
# 𝑆⁢𝑈⁢(𝑁) Skyrmions from Instantons11footnote 1To appear in Nonlinearity ## 1 Introduction In this paper we construct a class of cylindrically symmetric $`SU(N)`$ instantons and calculate some of their properties, like the topological charges. This construction involves harmonic maps of the plane into C CIP<sup>N-1</sup>. The symmetry of the solutions determines the dependence of the fields on the three dimensional polar angles and leaves unknown only the dependence on the three dimensional radius ($`r`$) and the Euclidean time $`(\tau )`$. These solutions describe instantons with the same spatial location but centered at different times and with different scales. Atiyah and Manton have observed that computing the holonomy of $`SU(2)`$ instantons in $`\text{I}\text{R}^4`$ generates configurations in $`\text{I}\text{R}^3`$ which are good approximations to solutions of the Skyrme model. Here, we extend their construction to $`SU(N)`$ and derive explicitly some $`SU(3)`$ spherically symmetric skyrmions from the $`SU(3)`$ cylindrically symmetric instantons. In addition, we connect these skyrmion solutions with the ones obtained using the harmonic maps of $`S^2`$ to C CIP<sup>N-1</sup> . Perhaps we should make it clear that there are several approaches to studying cylindrically symmetric instantons and spherically symmetric skyrmions. The main aim of this paper is not the construction of new instanton or skyrmion solutions, but rather to gain a better understanding of the correspondence between them and harmonic maps. In particular, using the Atiyah-Manton procedure , we obtain explicit closed form approximations for the profile functions of the $`SU(N)`$ Skyrme fields, which until now could only be determined numerically. ## 2 $`SU(N)`$ Instantons Instantons are solutions of the $`SU(N)`$ Yang-Mills equations in $`\text{I}\text{R}^4`$ which are derived from the action functional $$S=\frac{1}{16\pi ^2}\text{tr}(F_{ij}^2)d^4x$$ (2.1) which is expressed in terms of topological charge units, ie $`S|k|`$. Here $`k`$ is the topological charge (or Pontryagin index) which counts the number of instantons of the configuration. \[Since we are in the self-dual sector, ie $`F=F`$, the topological charge $`k`$ is given by (2.1)\]. $`A_i`$, for $`i=0,1,2,3`$, is the $`su(N)`$-valued gauge potential, with field strength $`F_{ij}=_iA_j_jA_i+[A_i,A_j]`$ and covariant derivative $`D_i=_i+[A_i,]`$. The finiteness of the action implies that the field strength must go to zero at spatial infinity, which means that the gauge field $`A_i`$ must be a pure gauge at spatial infinity. Variation of the action (2.1) gives the second order Yang-Mills equations $$D_iF_{ij}=0.$$ (2.2) The starting point for our investigation is the introduction of the coordinates $`u,\overline{u},z,\overline{z}`$ on $`\text{I}\text{R}^4.`$ In terms of the usual spherical coordinates $`r,\theta ,\phi `$ the Riemann sphere variable is $`z=e^{i\phi }\mathrm{tan}(\theta /2)`$, while $`u=r+i\tau `$. Using these coordinates the Yang-Mills equations (2.2) take the form $`D_u((u+\overline{u})^2F_{u\overline{u}})+(1+|z|^2)^2(D_zF_{u\overline{z}}+D_{\overline{z}}F_{uz})`$ $`=`$ $`0`$ (2.3) $`D_{\overline{u}}((u+\overline{u})^2F_{u\overline{u}})(1+|z|^2)^2(D_zF_{\overline{u}\overline{z}}+D_{\overline{z}}F_{\overline{u}z})`$ $`=`$ $`0`$ (2.4) $`D_uF_{\overline{u}z}+D_{\overline{u}}F_{uz}{\displaystyle \frac{1}{(u+\overline{u})^2}}D_z((1+|z|^2)^2F_{z\overline{z}})`$ $`=`$ $`0`$ (2.5) $`D_uF_{\overline{u}\overline{z}}+D_{\overline{u}}F_{u\overline{z}}+{\displaystyle \frac{1}{(u+\overline{u})^2}}D_{\overline{z}}((1+|z|^2)^2F_{z\overline{z}})`$ $`=`$ $`0`$ (2.6) while the corresponding action (2.1) becomes $$S=\frac{1}{8\pi ^2}\text{tr}\left(4F_{u\overline{u}}^2+\frac{8(1+|z|^2)^2}{(u+\overline{u})^2}(|F_{uz}|^2+|F_{u\overline{z}}|^2)\frac{4(1+|z|^2)^4}{(u+\overline{u})^4}F_{z\overline{z}}^2\right)r^2𝑑r𝑑\tau \frac{2idzd\overline{z}}{(1+|z|^2)^2}.$$ (2.7) A complex gauge can, at least for the self-dual case, be chosen so that $$A_{\overline{u}}=0,A_{\overline{z}}=0.$$ (2.8) The full equations (2.4) and (2.6) become then the total $`_{\overline{u}}`$ and $`_{\overline{z}}`$ derivative, respectively, of the self-dual Yang-Mills equation $$_{\overline{u}}A_u+\frac{(1+|z|^2)^2}{(u+\overline{u})^2}_{\overline{z}}A_z=0$$ (2.9) while the other two become simply $`F_{uz}=0`$, ie they define the gauge $$A_z=H^1H_z,A_u=H^1H_u$$ (2.10) where $`HSL(N,\text{C}\text{ })`$ is a Hermitian matrix and subscripts denote partial differentiation. To proceed further we need to briefly recall some results about harmonic maps of the two-dimensional C CIP<sup>N-1</sup> sigma model. See Zakrzewski for a more detailed account of two-dimensional sigma models and their solutions. ### 2.1 Harmonic Maps The harmonic map (or sigma model) equations for the C CIP<sup>N-1</sup> model are given by $$[P_{z\overline{z}},P]=0$$ (2.1) where $`P`$ is an $`N\times N`$ Hermitian projector. One set of solutions to these equations are the instantons given by $$P(f)=\frac{ff^{}}{|f|^2}$$ (2.2) where $`f(z)`$ is an $`N`$-component column vector which is a holomorphic function of $`z`$ and whose degree is equal to the topological charge of the sigma model. Another set of solutions are the anti-instantons, which have the same form but this time $`f`$ is an anti-holomorphic function, and then the sigma model topological charge is minus the degree of $`f.`$ For $`N=2`$ these are all the finite action solutions, but for $`N>2`$ there are other non-instanton solutions. These can be described by introducing the operator $`\mathrm{\Delta }`$ defined by its action on any vector $`f\text{C}\text{ }^N`$ as $$\mathrm{\Delta }f=_zf\frac{f(f^{}_zf)}{|f|^2}$$ (2.3) and then define further vectors $`\mathrm{\Delta }^kf`$ by induction: $`\mathrm{\Delta }^kf=\mathrm{\Delta }(\mathrm{\Delta }^{k1}f)`$. To proceed further we note the following useful properties of $`\mathrm{\Delta }^kf`$ when $`f`$ is holomorphic: $`(\mathrm{\Delta }^kf)^{}\mathrm{\Delta }^lf=0,kl`$ (2.4) $`_{\overline{z}}\left(\mathrm{\Delta }^kf\right)=\mathrm{\Delta }^{k1}f{\displaystyle \frac{|\mathrm{\Delta }^kf|^2}{|\mathrm{\Delta }^{k1}f|^2}},_z\left({\displaystyle \frac{\mathrm{\Delta }^{k1}f}{|\mathrm{\Delta }^{k1}f|^2}}\right)={\displaystyle \frac{\mathrm{\Delta }^kf}{|\mathrm{\Delta }^{k1}f|^2}}.`$ (2.5) These properties either follow directly from the definition of $`\mathrm{\Delta }`$ or are easy to prove . It is also convenient to define projectors $`P_k`$ corresponding to the family of vectors $`\mathrm{\Delta }^kf`$ as $$P_k=P(\mathrm{\Delta }^kf),k=0,\mathrm{},N1.$$ (2.6) Applying $`\mathrm{\Delta }`$ a total of $`N1`$ times to a holomorphic vector gives an anti-holomorphic vector, so that a further application of $`\mathrm{\Delta }`$ gives the zero vector and hence no corresponding projector. The projectors $`P_k`$ are solutions of the harmonic map equations (2.1) and all solutions can be found in this way by starting with an appropriate holomorphic vector $`f`$. In the C CIP<sup>1</sup> case the operator $`\mathrm{\Delta }`$ converts a holomorphic vector to an anti-holomorphic vector, that is, instantons to anti-instantons and these are all the solutions in this case. Note that the projectors obtained from this sequence always satisfy the relation $`_{k=0}^{N1}P_k=1`$. ### 2.2 Constructing the Instantons The self-dual Yang-Mills equation (2.9) after the gauge choice (2.10) is equivalent to the single equation for $`H`$ $$_{\overline{u}}(H^1H_u)+\frac{(1+|z|^2)^2}{(u+\overline{u})^2}_{\overline{z}}(H^1H_z)=0.$$ (2.1) This is similar to an equation introduced by Jarvis for studying monopoles, under a dimensional reduction of the time $`\tau `$. As we are going to illustrate, exact $`SU(N)`$ instanton solutions of (2.1) can be obtained using harmonic maps, ie by assuming that the field $`H`$ is of the form $$H=\mathrm{exp}\left\{g_0\left(P_0\frac{1}{N}\right)+g_1\left(P_1\frac{1}{N}\right)+\mathrm{}+g_{N2}\left(P_{N2}\frac{1}{N}\right)\right\}$$ (2.2) where $`g_i=g_i(u,\overline{u})`$ for $`i=0,\mathrm{},N2,`$ are arbitrary functions of $`u`$ and $`\overline{u}`$. Recall that the projector $`P_{N1}`$ is a linear combination of the other projectors plus the identity matrix, which is why it is not included in the above formula. The above ansatz is motivated by our recent study of Bogomolny monopoles and their construction in terms of harmonic maps. In order for our ansatz (2.2) to give solutions to (2.1), the harmonic maps used must have spherical symmetry — essentially the factors of $`(1+|z|^2)^2`$ which appear in (2.1) must be cancelled. The required harmonic maps are obtained by applying the above procedure to the initial holomorphic vector $$f=(f_{N1},\mathrm{},f_j,\mathrm{},f_0)^t,\text{where}f_j=z^j\sqrt{\left(\genfrac{}{}{0pt}{}{N1}{j}\right)}$$ (2.3) and $`\left(\genfrac{}{}{0pt}{}{N1}{j}\right)`$ denote the binomial coefficients. For a discussion of the spherical symmetry of these maps see Ref. . Here we merely point out that it is at least plausible that the required factors do indeed cancel since $`|f|^2=(1+|z|^2)^{N1}.`$ We shall illustrate this explicitly in the following with some examples. $`SU(2)`$ Case There are simplifying special cases for which we are able to perform the construction explicitly, the easiest example being the rational map $`f=(z,1)^t`$. For $`N=2`$, there is only one profile function $`g_0`$ and our ansatz (2.2) reduces the self-dual equation (2.1) to the following differential equation $$(u+\overline{u})^2g_{0u\overline{u}}+2(e^{g_0}1)=0.$$ (2.4) Moreover, the topological charge $`k`$ is given by $$k=\frac{1}{4\pi }\left(4g_{0u\overline{u}}^2+\frac{4}{r^2}e^{g_0}|g_{0u}|^2+\frac{1}{r^4}(e^{g_0}1)^2\right)r^2𝑑r𝑑\tau $$ (2.5) where we have used the fact that $`i𝑑z𝑑\overline{z}(1+|z|^2)^2=2\pi `$. To actually solve (2.4), we first let $`g_0=2\mathrm{ln}(\frac{u+\overline{u}}{2})+2\rho _0`$ (where $`\rho _0`$ is a new unknown function). Then equation (2.4) becomes $$4\rho _{0u\overline{u}}=e^{2\rho _0}$$ (2.6) which is the so-called Liouville equation and can be solved explicitly using conformal invariance. In this case, the function $`g_0`$ is $$g_0=2\mathrm{ln}\left(\frac{(u+\overline{u})|dh/du|}{1|h|^2}\right).$$ (2.7) Here $`h`$ is, an analytic function of $`u`$, of the form (see Ref. ) $$h=\underset{i=1}{\overset{k+1}{}}\frac{a_iu}{\overline{a}_i+u}$$ (2.8) and the $`a_i`$ are an arbitrary set of complex numbers (some of them perhaps equal) constrained to have Re $`a_i>0`$. Then (2.7) provide the most general solution of (2.1) with cylindrical symmetry and finite action. For general $`k+1`$, the total multiplicity of the zeros of $`h`$ in the right half plane is always $`k`$; therefore this solution describes $`k`$ instantons. The imaginary part of the zeroes of $`h`$ determines the location of the instantons along the time axis, while the real part determines the instanton scales . For the special case where $`h=(a_1u)^2/(\overline{a}_1+u)^2`$ (for $`a_1=\lambda +i\lambda `$) the topological charge $`k`$ of our solution (2.7) is equal to one, ie the configuration consists of one instanton solution located at the origin with scale $`\lambda `$. $`SU(3)`$ Case For $`N=3`$ there are two profile functions $`g_0`$ and $`g_1`$ and equation (2.1) reduces to $`(u+\overline{u})^2g_{0u\overline{u}}+2(e^{g_1}1)+2(e^{g_0g_1}1)`$ $`=`$ $`0`$ $`(u+\overline{u})^2g_{1u\overline{u}}+4(e^{g_1}1)2(e^{g_0g_1}1)`$ $`=`$ $`0.`$ (2.9) It is immediately clear that there is a symmetry under the interchange of indices $`01`$, when applied simultaneously to $`g_i`$ for $`i=0,1`$. As we will show later, this symmetry can be used to derive special instantons which involve a smaller number of profile functions and projectors. In addition, the topological charge becomes $`k`$ $`=`$ $`{\displaystyle \frac{1}{3\pi }}{\displaystyle }\{4(g_{0u\overline{u}}^2+g_{1u\overline{u}}^2g_{0u\overline{u}}g_{1u\overline{u}})+{\displaystyle \frac{6}{r^2}}e^{g_0g_1}(|g_{0u}|^2+|g_{1u}|^2g_{0u}g_{1\overline{u}}g_{0\overline{u}}g_{1u})+`$ $`{\displaystyle \frac{6}{r^2}}e^{g_1}|g_{1u}|^2+{\displaystyle \frac{1}{r^4}}(3e^{2(g_0g_1)}3e^{(g_0g_1)}+3+3e^{2g_1}+3e^{g_1}3e^{g_0})\}r^2drd\tau `$ The profile functions equations that we obtain, ie (2.9), are related to those derived from the ansatz based approach of Bais et al and the methods employed there can be adapted to solve for the functions explicitly. For the case in which the functions $`g_k`$ are time independent, we have shown that , the equations (2.11) are precisely the equations for spherically symmetric monopoles. To solve (2.9), we first let $`g_0=2\mathrm{ln}\left({\displaystyle \frac{(u+\overline{u})^2}{4}}\right)+2\rho _0+2\rho _1`$ $`g_1=2\mathrm{ln}\left({\displaystyle \frac{(u+\overline{u})}{2}}\right)+4\rho _02\rho _1`$ (2.11) where $`\rho _0`$ and $`\rho _1`$ are arbitrary functions of $`u`$ and $`\overline{u}`$. Then (2.9) simplifies to $$4\rho _{0u\overline{u}}=e^{4\rho _02\rho _1},4\rho _{1u\overline{u}}=e^{4\rho _12\rho _0}$$ (2.12) which are a set of coupled Liouville equations. We have two sets of solutions to (2.12). The first one is just the maximal embedding of $`SU(2)`$ in $`SU(3)`$ and occurs when $`\rho _0=\rho _1`$, that is for $`g_0=2g_1`$ (ie, using the symmetry). Then the system (2.12) simplifies to the Liouville equation and the solution is just Witten’s solution embedded in $`SU(3)`$, ie $`g_1`$ is given by (2.7). In this case, the topological charge (LABEL:duoins) is exactly four times the $`SU(2)`$ one (given by (2.5)). Therefore, for the special case $`h=(a_1u)^2/(\overline{a}_1+u)^2`$, the configuration consists of four instantons — placed on top of each other in space. The second set of solutions to (2.11) describes an irreducible $`SU(3)`$ instanton for which $`\rho _0={\displaystyle \frac{1}{2}}\mathrm{ln}\left[{\displaystyle \frac{3|dh/du|^2}{|h|^{1/3}(1|h|)^2(1+2|h|)}}\right]`$ $`\rho _1={\displaystyle \frac{1}{2}}\mathrm{ln}\left[{\displaystyle \frac{3|dh/du|^2}{|h|^{2/3}(1|h|)^2(2+|h|)}}\right]`$ (2.13) where $`h`$ is given by (2.8). For the special case where $`h=(a_1u)^2/(\overline{a}_1+u)^2`$ (for $`a_1=\lambda +i\lambda )`$ the topological charge (LABEL:duoins) of our solution (2.11) is equal to two. ## 3 $`SU(N)`$ Skyrmions The Skyrme model is a nonlinear field theory which provides a good description of low energy hardon physics. To have finite-energy configurations, one must require that the field $`U(\stackrel{}{x},t)SU(N)`$ goes to a constant matrix (say 1) at spatial infinity: $`U1`$ as $`|\stackrel{}{x}|\mathrm{}`$. This effectively compactifies the three-dimensional Euclidean space onto $`S^3`$ and hence implies that the Skyrme field can be considered as a map from $`S^3`$ into $`SU(N)`$; and therefore it can be classified by the third homotopy group $`\pi _3(SU(N))=Z`$ or, equivalently, by the integer valued winding number $$B=\frac{1}{24\pi ^2}_{R^3}\epsilon _{ijk}\text{tr}\left(_iUU^1_jUU^1_kUU^1\right)d^3\stackrel{}{x}$$ (3.1) which is topological invariant. This winding number classifies the solitonic sectors in the model, and as Skyrme has argued , it may be identified with the baryon number $`B`$ of the field configuration. In the static limit, the energy of the Skyrme model is $$E=\frac{1}{12\pi ^2}_{R^3}\left\{\frac{1}{2}\text{tr}\left(_iUU^1\right)^2\frac{1}{16}\text{tr}[_iUU^1,_jUU^1]^2\right\}d^3\stackrel{}{x}$$ (3.2) which is expressed in the same units as the baryon number. There is a lower bound on the energy of a given configuration in terms of the baryon number, ie $`E|B|`$. Since the model is not integrable explicit skyrmion solutions are not known and therefore, must be obtained by solving the equations numerically. In what follows using the Atiyah-Manton approach and harmonic maps we construct explicitly approximations to the $`SU(N)`$ skyrmions. ### 3.1 Harmonic Maps Recently in , $`SU(N)`$ spherically symmetric skyrmion fields have been constructed from the harmonic maps of $`S^2`$ to C CIP<sup>N-1</sup> (which are not embeddings of the $`SU(2)`$ fields). In fact, the Skyrme field involves the introduction of $`N1`$ projectors, ie $$U=\mathrm{exp}\left\{ig_{0_{Sk}}\left(P_0\frac{1}{N}\right)+ig_{1_{Sk}}\left(P_1\frac{1}{N}\right)+\mathrm{}+ig_{(N2)_{Sk}}\left(P_{N2}\frac{1}{N}\right)\right\}$$ (3.3) where $`g_{i_{Sk}}=g_{i_{Sk}}(r)`$ for $`i=0,\mathrm{},N2`$ are the profile functions. Moreover, the energy (3.2) becomes $`E`$ $`=`$ $`{\displaystyle \frac{1}{6\pi }}{\displaystyle }r^2dr\{{\displaystyle \frac{1}{N}}\left({\displaystyle \underset{i=0}{\overset{N2}{}}}\dot{g}_{i_{Sk}}\right)^2+{\displaystyle \underset{i=0}{\overset{N2}{}}}\dot{g}_{i_{Sk}}^2+{\displaystyle \frac{1}{2r^2}}{\displaystyle \underset{k=1}{\overset{N1}{}}}(\dot{g}_{k_{Sk}}\dot{g}_{(k1)_{Sk}})^2D_k+{\displaystyle \frac{2}{r^2}}{\displaystyle \underset{k=1}{\overset{N1}{}}}D_k`$ (3.4) $`+{\displaystyle \frac{1}{4r^4}}(D_1^2+{\displaystyle \underset{k=1}{\overset{N2}{}}}(D_kD_{k+1})^2+D_{N1}^2)\}`$ where $`D_k=k(Nk)(1\mathrm{cos}(g_{k_{Sk}}g_{(k1)_{Sk}}))`$. In addition, the topological charge (3.1) takes the form $$B=\frac{1}{2\pi }\underset{i=0}{\overset{N2}{}}(i+1)(Ni1)\left(g_{i_{Sk}}g_{(i+1)_{Sk}}\mathrm{sin}(g_{i_{Sk}}g_{(i+1)_{Sk}})\right)_{r=0}^{r=\mathrm{}}.$$ (3.5) As $`g_{i_{Sk}}(\mathrm{})=0`$ (required for the finiteness of the energy) the only contributions to the topological charge comes from $`g_i(0)`$. This way a family of exact spherically symmetric solutions of the $`SU(N)`$ Skyrme model has been obtained. In fact for each $`SU(N)`$ model the Skyrme field involving $`N1`$ projectors leads to an exact solutions involving $`N1`$ profile functions $`g_{i_{Sk}}`$. These profile functions $`g_{i_{Sk}}`$, which satisfy $`N1`$ coupled nonlinear ordinary differential equations and can be solved numerically, are exhibited in Ref. . Next we will derive explicitly analytic forms for $`g_{i_{Sk}}`$ in (3.3) which are good approximations to the ones obtained numerically in . ### 3.2 Derivation of Skyrmions It has been proposed by Atiyah and Manton that a finite-dimensional manifold of Skyrme fields can be generated from $`SU(2)`$ self-dual Yang-Mills fields. The main idea of the Atiyah-Manton scheme is to construct the holonomy $$U(\stackrel{}{x})=𝒯\text{exp}\left(_{\mathrm{}}^{\mathrm{}}A_\tau (\stackrel{}{x},\tau )𝑑\tau \right)$$ (3.1) where $`𝒯`$ denotes time-ordering and $`x=(\stackrel{}{x},\tau )`$ denotes the time line through $`\stackrel{}{x}`$. The fundamental topological result is that if we consider this holonomy as a Skyrme field then it has baryon number $`B=k`$, where $`k`$ is the topological charge of the gauge potential $`A_i`$ (ie, instanton number). The fields can be explicitly computed in some special cases, by integrating simple expressions. The gauge field $`A_\tau `$, using the harmonic map ansatz (2.2) and due to the gauge choice (2.8), takes the form $$A_\tau iA_u=ig_{0u}(\frac{1}{N}P_0)ig_{1u}(\frac{1}{N}P_1)\mathrm{}ig_{(N2)u}(\frac{1}{N}P_{N2}).$$ (3.2) By substituting (3.2) into the Atiyah-Manton formula (3.1) and comparing with (3.3) we see that the profile functions of the instanton and Skyrme fields are related. In fact, the profile functions of the $`SU(N)`$ Skyrme fields $`g_{i_{Sk}}`$ can be determined analytically through the relation $$g_{i_{Sk}}=_{\mathrm{}}^{\mathrm{}}g_{iu}(r,\tau )𝑑\tau $$ (3.3) for $`i=0,\mathrm{},N2`$. \[Recall that, these profile functions (3.3) are approximations to the ones obtained numerically in \]. Next, we will determine analytically the $`g_{i_{Sk}}`$ profile functions of the Skyrme field, for the simplest cases of $`SU(2)`$ and $`SU(3)`$. This way, we construct a class of spherically symmetric skyrmions and calculate some of their properties, such as their energies and baryon numbers. Recall that, their baryon number is equal to the topological charge of the corresponding instanton configuration they are derived from. $`SU(2)`$ Case There is only one profile function in this case, $`g_{0_{Sk}}`$, which can be obtained from (3.3) for $`g_0`$ given by (2.7). This gives the well-known one spherically symmetric skyrmion solution, with $$g_{0_{Sk}}=2\pi \left[1(1+\lambda ^2/r^2)^{1/2}\right].$$ (3.4) This Skyrme field has first been obtained in . It can be shown that the configuration consists of one skyrmion (due to (3.5)) and the energy obtained from (3.4) has minimum at $`\lambda ^2=2.11`$ and is equal to $`E=1.2432`$, ie within 1% of the numerically determined value 1.232. $`SU(3)`$ Case For $`N=3`$ there are two profile functions $`g_{0_{Sk}}`$ and $`g_{1_{Sk}}`$ which again can be evaluated explicitly using (3.3). \[Recall that we have two set of solutions for the instanton functions\]. When $`g_0=2g_1`$ for $`g_1`$ given by (2.7), we end up having one profile function for the Skyrme field which coincides with the profile function of a single $`SU(2)`$ skyrmion, ie $$g_{1_{Sk}}=2\pi \left[1(1+\lambda ^2/r^2)^{1/2}\right].$$ (3.5) Here we note that as $`g_{0_{Sk}}(0)=4\pi `$ and $`g_{1_{Sk}}(0)=2\pi `$ the topological charge (3.5) of our solution is four (so is the instanton number). The energy obtained from (3.4) of this configuration is exactly four times the energy of the one $`SU(2)`$ skyrmion, ie $`E=4\times 1.2432`$. This compares with the energy of the numerically solution $`4\times 1.232`$ determined in . \[Again the minimum occurs at $`\lambda ^2=2.11`$\]. In the case where the two instanton profile functions are given by equations (2.11) and (2.13), after performing the integration (3.3) we get $`g_{0_{Sk}}`$ $`=`$ $`2\pi \left(1{\displaystyle \frac{\lambda +3r}{2\sqrt{9\lambda ^2+6\lambda r+9r^2}}}+{\displaystyle \frac{\lambda 3r}{2\sqrt{9\lambda ^26\lambda r+9r^2}}}\right)`$ $`g_{1_{Sk}}`$ $`=`$ $`\pi \left(1+{\displaystyle \frac{\lambda +3r}{\sqrt{9\lambda ^2+6\lambda r+9r^2}}}+{\displaystyle \frac{2(\lambda 3r)}{\sqrt{9\lambda ^26\lambda r+9r^2}}}\right).`$ (3.6) Since $`g_{0_{Sk}}(0)=g_{1_{Sk}}(0)=2\pi `$ the baryon number (3.1) is two (recall, $`k=2`$); the interpretation of this solution is therefore that it contains two skyrmions. The approximate energy obtained from (3.4) is then $`E=2.39815`$ which is in good agreement with the true value $`2.3764`$ determined in . This solutions has, first, been obtained by Balachandran et al . ## 4 Conclusion We have studied in some detail the construction of $`SU(N)`$ instanton from harmonic maps. Explicit solutions have been obtained in the case of cylindrical symmetry and we have shown how these solutions involve harmonic maps of the plane into C CIP<sup>N-1</sup>. In addition, we have generate approximate solutions of the $`SU(N)`$ Skyrme model using the Atiyah-Manton procedure, ie by computing instanton holonomies. The corresponding skyrmions are spherically symmetric, their energies are in good agreement with their exact value and their baryon number is equal to the instanton number. The same approach can be applied to generate other soliton approximations in lower dimensions (see for example, Ref. ), leading to the fact that they might be many connections between solitons and instantons in varying spacetime dimensions. ## Acknowledgements Many thanks to Paul Sutcliffe for useful discussions and to the Nuffield Foundation for a newly appointed lecturer award.
warning/0004/cond-mat0004422.html
ar5iv
text
# Reversible Boolean Networks I: Distribution of Cycle Lengths ## 1 Introduction ### 1.1 Review of dynamical boolean networks In recent years considerable effort has been devoted to the study of the development of complexity in dynamical systems. Complexity is observed in such different examples as ecosystems (see ), spin glasses (see ), and quite broadly through the biological sciences. One thread of activity involves the study of the behavior of dynamical systems consisting of $`N`$ variables $`\sigma _t^j`$ (the site label $`j=1,\mathrm{},N`$), where each $`\sigma _t^j`$ is Boolean, so that it takes on one of two values that we choose to be $`\pm 1`$. The configuration of the system at time $`t`$ is characterized by a ‘state’ $`\mathrm{\Sigma }_t`$: $`\mathrm{\Sigma }_t=(\sigma _t^1,\sigma _t^2,\mathrm{},\sigma _t^j,\mathrm{},\sigma _t^N).`$ (1) The time development is given by saying that the state at time $`t+1`$ is a prescribed function of the state at time $`t`$, i.e. $`\mathrm{\Sigma }_{t+1}=(\mathrm{\Sigma }_t).`$ (2) The mapping can also be written $`\sigma _{t+1}^j=F^j(\mathrm{\Sigma }_t)\text{ for }j=1,\mathrm{},N`$ (3) where the $`F^j`$ also take on the values $`\pm 1`$. The number of different states of the system is finite; it is $$\omega =2^N,$$ (4) and the mapping function $``$ is independent of time. Therefore, starting from any state, eventually the system falls into a cyclic behavior and follows that cycle forever. Typically, each of the functions $`F^j(\mathrm{\Sigma })`$ is picked so that it depends upon exactly $`K`$ distinct input spin variables in the vector $`\mathrm{\Sigma }`$. A random choice is made to determine which components, $`\sigma ^k`$ will appear in each $`F^j(\mathrm{\Sigma })`$; this fixes the ‘wiring’ of the realization. Once $`K`$ and $`N`$ and the assignment $`\sigma ^k`$’s are fixed, the each mapping function $`F^j(\mathrm{\Sigma })`$ is selected at random from the set of all Boolean functions of $`K`$ Boolean variables. The randomly chosen $`N`$ sets of $`K`$ input spin variables and the functions $`F^j`$ compose a realization. Given an initial configuration of the system, a realization will completely define the system’s behavior. The realization is picked at the beginning of the calculation for the system and remains independent of time. Since there are $`2^{2^K}`$ Boolean functions of $`K`$ Boolean variables and $`({}_{N}{}^{K})`$ different ways to assign $`K`$ distinct input spin variables, as $`K`$ and/or $`N`$ become large the number of different realizations is truly huge. This kind of model is often called a Kauffman net because Stuart Kauffman developed a program of study for generic maps of this type. Later, this program was extended by Derrida, Flyvberg, Parisi, and others. Quantities of interest that have been studied include the distribution of cycle-lengths and the number of starting points which will eventually lead to a given cycle. (These points form what is called the basin of attraction of the cycle.) One calculates the above quantities by enumeration or by Monte Carlo simulation for each realization and then averages over realizations with the same values of $`N`$ and $`K`$. The Kauffman net is said to be ‘dissipative’ since several different states may map into one. Thereby information is lost. The behavior of Kauffman nets are interesting and surprising. Large $`K`$-values produce a complex time-behavior which closely resembles Parisi’s theory of spin glasses. For large $`K`$, the cycle lengths grow exponentially with the number of spins $`N`$. Conversely, for $`K=0`$ or $`K=1`$, the cycles tend to be short. At the ‘critical’ value, $`K=2`$, typical cycle lengths grow as a power of $`N`$, and for large $`N`$ the probability of observing a cycle of length $`L`$ varies as a power of $`L`$. (For a study of critical properties see Refs and .) This three-phase structure is typical of phase transition problems in which an ordered and a disordered phase are separated by a critical phase line. ### 1.2 A time-reversible network Thus, a great deal is known about the behavior of Kauffman nets, which can be viewed as a class of generic dynamical mapping problems. But not all problems are generic. For example, many of the systems considered in Hamiltonian mechanics are reversible. Such systems have the property that some transformation of the coordinates (for example changing the sign of all velocities) makes the system retrace its previous path. Thus a forward motion and its inverse are equally possible. This paper is devoted to a study of the behavior of discrete reversible maps. In contrast to a dissipative system, in a finite and reversible dynamical system every possible state is in exactly one cycle. Because one and only one state at time $`t`$ maps into a predefined state at time $`t+1`$, any cycle can be traversed equally well forward or backward. There are no basins of attraction in this kind of system. The long-term properties are then described by giving the number of cycles of length $`l`$, $`N(l)`$. It turns out that for the smaller values of $`K`$, $`N(l)`$ is an oscillating function of $`l`$ in which the small prime divisors of $`l`$ play a major role. We shall study this effect in a companion paper. For now, we focus on the gross scaling properties of $`N(l)`$, by using a cumulative distribution $$S(l)=\frac{_{j=l+1}^{\mathrm{}}jN(j)}{_{k=1}^{\mathrm{}}kN(k)},$$ (5) which gives the probability of finding a cycle of length greater than $`l`$ by picking the realizations and the initial cycle element at random. We call $`S(l)`$ a “Survival probability”. #### 1.2.1 Definition of the model We construct our time-reversible maps using the method of two time-slices that was studied for discrete systems by Fredkin and collaborators.<sup>1</sup><sup>1</sup>1The construction of time-reversible models by using variables from two different times has a long history. Imagine doing a calculation in ordinary classical mechanics using small but discrete time-steps. The behavior depends upon the position and velocity of each particle, but one might wish to do the analysis in terms of positions alone. To do this, one works with a state defined by the positions at two closely neighboring times. The difference in position at the two times gives an estimate of the velocity vector. In this way one can construct a two slice model of particle behavior. As in the usual dissipative Kauffman model, our basic variable is a list, $`\mathrm{\Sigma }`$, of the $`N`$ ‘spin variables’, $`\sigma ^j`$: $`\mathrm{\Sigma }=(\sigma ^1,\sigma ^2,\mathrm{},\sigma ^j,\mathrm{},\sigma ^N).`$ (6) The value of $`\mathrm{\Sigma }`$ is given for each integer value of the time, and written as $`\mathrm{\Sigma }_t`$. In our reversible mappings the spin configuration at time $`t+1`$, $`\mathrm{\Sigma }_{t+1}`$, depends upon $`\mathrm{\Sigma }_t`$ and $`\mathrm{\Sigma }_{t1}`$ according to the rule $`\sigma _{t+1}^j=\sigma _{t1}^jF^j(\mathrm{\Sigma }_t),`$ (7) where the $`F`$’s are picked exactly as in the dissipative Kauffman net. Since the $`\sigma `$’s take on the values $`\pm 1`$, our model can be written in the equivalent form $`\sigma _{t+1}^j\sigma _{t1}^j=F^j(\mathrm{\Sigma }_t),`$ (8) which exhibits a quite manifest time reversal invariance. The models given by equation (7) are the subject of this paper. The information needed to predict future time steps is called the state of the system. In our case, the state at time $`t`$, $`𝒮_t`$, is given by two time slices or substates $`\mathrm{\Sigma }_t`$ and $`\mathrm{\Sigma }_{t1}`$ as $`𝒮_t=\left({\displaystyle \genfrac{}{}{0pt}{}{\mathrm{\Sigma }_{t1}}{\mathrm{\Sigma }_t}}\right),`$ (9) and the full mapping is of the form $$𝒮_{t+1}=(𝒮_t).$$ (10) The history of the system is given by listing the substates in order as $$\mathrm{\Sigma }_0,\mathrm{\Sigma }_1,\mathrm{\Sigma }_2,\mathrm{},\mathrm{\Sigma }_n$$ If there are $`N`$ spins, the volume of the state space is $`\mathrm{\Omega }=2^{2N}=\omega ^2.`$ (11) Since the dynamics are deterministic, $`\mathrm{\Omega }`$ gives the length of the longest possible cycle. However, as we shall see, that length is never attained. ### 1.3 Questions to be asked This paper concerns the distribution of cycle lengths in the time-reversible models. Figure 1 shows plots of probabilities of observing a cycle of length larger than $`l`$ for systems with $`N=10`$ and various $`K`$, averaged over realizations. For $`K=1`$ the cycles are very short; for $`K=N`$ they have a wide range of lengths, but the longest ones have length of order $`2^N`$, much less than the number of points in the state space, $`2^{2N}`$. For $`K=2`$, a wide range of cycle lengths is seen, including some lengths which considerably exceed the ones in the $`K=N`$ system. For smaller values of $`K`$, the number of cycles of length $`l`$ tends to be a strongly oscillatory function of $`l`$. We shall discuss this oscillation in a subsequent publication. To understand these results for the cycle lengths, we shall need to understand in some detail the interplay between the quenched randomness of the system and the time-reversal symmetry. We will find that the cycles can be divided into two classes; those that are symmetric under time-reversal and those that are not. There are profound differences between the behaviors of these two types of cycles. The companion paper will discuss the growth of the ‘Hamming distance’ between configurations, defined as the the number of Boolean variables which are unequal in the two configurations. We shall compare two configurations that initially differ by a single spin flip, and see how this distance depends upon the number of iterations. For small $`K`$, the growth in the distance may be very slow; for large $`K`$ we see much more rapid growth. Our simulational data of cycle length distributions suggest but do not prove the existence of phase transition. However, in the companion publication we shall see that the behavior of the Hamming distance can be used to demonstrate convincingly the presence of a type of percolation transition at a value of $`K`$ of about 1.6. ### 1.4 Outline of the rest of paper In the next section, we discuss some special features of time-reversible models and their implications for classifying different kinds of cycles. The section after that is devoted to the limiting cases, $`K=0`$ and $`N`$. Section four describes the structures seen at intermediate $`K`$. The appendices cover some more peripheral issues. ## 2 Time reversal invariance At first sight, it is not obvious that time-reversal invariance should have any important effect on the distribution of cycle lengths. However, G. Birkhoff and later John M. Greene and others showed one important mechanism by which the time-reversal process works to set the cycle length. ### 2.1 Symmetry points In our model there are special points in the phase space which we might call mirrors. A mirror produces a time-reflected motion in the sequence, for example $`\mathrm{},\mathrm{\Sigma }_3,\mathrm{\Sigma }_2,\mathrm{\Sigma }_1,\mathrm{\Sigma }_1,\mathrm{\Sigma }_2,\mathrm{\Sigma }_3,\mathrm{}`$ (12a) or, as another example $`\mathrm{},\mathrm{\Sigma }_3,\mathrm{\Sigma }_2,\mathrm{\Sigma }_1,\mathrm{\Sigma }_0,\mathrm{\Sigma }_1,\mathrm{\Sigma }_2,\mathrm{\Sigma }_3,\mathrm{}`$ (12b) We call the first type (equation (12a)) a twin configuration and the second type (equation (12b)) a sandwich configuration. A sketch of the mirrors is given in figure 2. The phase space contains many of these mirrors. A typical and important cyclic motion is for the sequence of substates to hit a mirror, be reflected, hit another mirror, be reflected once more, and thereby be forced into a cyclic behavior. More explicitly, if we record an orbit of the time-reversible model using a sequence of substates, for example $`\mathrm{},\mathrm{\Sigma }_{t1},\mathrm{\Sigma }_t,\mathrm{\Sigma }_{t+1},\mathrm{}`$, we may call the state $`𝒮`$-value defined as $`\left(\genfrac{}{}{0pt}{}{\mathrm{\Sigma }_t}{\mathrm{\Sigma }_{t+1}}\right)`$ at which $`\mathrm{\Sigma }_{t+1}=\mathrm{\Sigma }_t`$ a twin special point, and another $`𝒮`$-value a sandwich special point, if $`\mathrm{\Sigma }_{t+2}`$ in the sequence equals $`\mathrm{\Sigma }_t`$. Thus a twin point is any state of the form $$𝒮=\left(\genfrac{}{}{0pt}{}{\mathrm{\Sigma }}{\mathrm{\Sigma }}\right),$$ (13) so that the cycle will spread out in a palindromic fashion before and after this special point in the pattern of equation (12a). Since the entire volume of the state space is $`\mathrm{\Omega }=\omega ^2`$ and since there are $`\omega `$ of these invariant points, the chance that a randomly chosen point in the state space is a twin point is $`1/\omega `$. Correspondingly, a sandwich point appears when there is some value of $`\mathrm{\Sigma }^{(s)}`$ for which $`F^j(\mathrm{\Sigma }^{(s)})=1\text{ for all }j`$ (14) (see equation (7)). A sandwich point gives rise to a sequence of the form of equation (12b). The number of sandwich points depends on the realization. For example, if $`F^j1`$ for any $`j`$, then there are no sandwich points at all. For a given realization, we denote the number of substates that have the property of equation (14) by $`m`$, and we denote each such substate by the symbol $`\mathrm{\Sigma }_\alpha ^{(s)}`$; the $`\alpha `$ label differentiates between the $`m`$ different sandwich substates. Since for any substate $`\mathrm{\Gamma }`$ a state of the form $`𝒮=\left(\genfrac{}{}{0pt}{}{\mathrm{\Gamma }}{\mathrm{\Sigma }_\alpha ^{(s)}}\right)`$ is a sandwich state, there must be $`m\omega `$ of these sandwich states. Appendix A discusses the properties of the sandwich points. As shown there, the average over realizations $`m`$ is unity. When $`K=N`$ and $`N`$ is large, $`m`$ follows a Poisson distribution, $$P(m)=\frac{1}{m!e}\text{(}N\mathrm{}\text{)}.$$ (15) When $`K`$ is small, there are large realization-to-realization fluctuations in $`m`$, and the average over realizations of higher powers of $`m`$, for example $`m^2`$, can grow rapidly as a function of $`N`$. In fact, when $`N2^{(2^K)}`$, the vast majority of realizations have $`m=0`$. ### 2.2 Inversions and cycles Two states $`𝒮`$ and $`𝒮^{}`$ are time reversed images of one another if $`𝒮=\left(\genfrac{}{}{0pt}{}{\mathrm{\Sigma }_1}{\mathrm{\Sigma }_2}\right)`$ while $`𝒮^{}=\left(\genfrac{}{}{0pt}{}{\mathrm{\Sigma }_2}{\mathrm{\Sigma }_1}\right)`$. All cycles belong in one of two classes. The first class, which we call special cycles, each contains at least one pair of time-reversed images of one another. The other class, called regular cycles, contain no such pairs. In Appendix B we show that each special cycle of length greater than one contains exactly two distinct special points. Furthermore, if the points are both twins or both sandwiches, the cycle has even length; if they are of different types, the cycle length is odd. ### 2.3 Counting special points and special cycles The arguments in subsection 2.1 imply that for a given realization there are $`m\omega `$ sandwich points and $`\omega `$ twin points. Among them there are $`m`$ points that are both twin and sandwich points (consisting of a sandwich substate followed by the same substate), all of which produce $`m`$ cycles of length one. Thus, there are $`(m+1)\omega m`$ distinct special points in the state space. The total number of special cycles is the number derived from the points which are both twins and sandwiches, $`m`$, plus the number derived from all the other special points, $`[(m+1)\omega 2m]/2`$, yielding $`\text{number of special cycles}={\displaystyle \frac{(m+1)\omega }{2}}.`$ (16) The importance of these special points in determining cycle properties will become clear in the following sections. ## 3 Limiting cases This section discusses the behavior of reversible Boolean nets for the cases $`K=0`$ and $`K=N`$. ### 3.1 $`K=0`$ When $`K=0`$ the evolutions of the different $`\sigma ^j`$’s are uncorrelated. Each $`\sigma _t^j`$ repeats after either one, two, or four steps. (In contrast, in the Kauffman net after the first step each of the $`\sigma ^j`$’s remains constant.) For large $`N`$, each system is likely to contain at least one $`\sigma ^j`$ with period four. Hence cycles of period four will dominate. There will be roughly $`N(4)\omega ^2/4`$ such cycles. The uncorrelated cycles of the spins produce a hamming distance which can have value zero or one and has period one, two or four. ### 3.2 $`K=N`$ In this part we first review the results for the dissipative Kauffman net $$\sigma _{}^{j}{}_{t+1}{}^{}=F^j(\mathrm{\Sigma }_t)$$ and then proceed to discuss the time-reversible case $$\sigma _{}^{j}{}_{t+1}{}^{}=F^j(\mathrm{\Sigma }_t)\sigma _{}^{j}{}_{t1}{}^{}.$$ #### 3.2.1 Dissipative case The case in which $`K`$ has its maximum value, $`K`$=$`N`$, was first analyzed for the dissipative Kauffman model by Derrida. This case has the simplifying feature that a change of a single spin changes the input of every function $`F^j`$. Since the functions are chosen randomly, this means that every new input configuration $`\mathrm{\Sigma }_t`$ leads to an output configuration $`\mathrm{\Sigma }_{t+1}`$ that is picked at random from the whole phase space, with its volume $$\omega =2^N.$$ This process continues until the time $`T`$ at which $`\mathrm{\Sigma }_T=\mathrm{\Sigma }_\tau `$ for some $`\tau <T`$. After that, the system cycles repeatedly through the sequence $`\mathrm{\Sigma }_\tau ,\mathrm{},\mathrm{\Sigma }_{T1}`$. To find the distribution of orbit lengths, we first calculate $`p_n`$, the probability that starting from a randomly chosen initial state at time $`t=0`$, the orbit closes at time $`n`$. To do this, we define $`q_n`$ to be the probability that the cycle remains unclosed after $`n`$ steps. The probability that a closure event occurs at time zero is $`p_0=0`$, and thus $`q_0=1`$. At the time one, the system has a probability $`p_1=1/\omega `$ of falling into the initial value and a probability $`q_1=1p_1`$ of not doing so. At time two, there are two possible cycle-closures, since the new element can be the same as either the zeroth or the first element. Thus the conditional probability of a closure at time two, given that the closure event did not occur at any earlier time, is $`2/\omega `$. Similarly, the conditional probability of a closure event at time $`t=n`$, given that the system has not closed at time $`t=n1`$, is just $`n/\omega `$. Thus the likelihood of a cycle closure at step $`n`$ is $`p_n={\displaystyle \frac{n}{\omega }}q_{n1}`$ (17a) and correspondingly the $`q_n`$ satisfy $`q_n=\left(1{\displaystyle \frac{n}{\omega }}\right)q_{n1}.`$ (17b) The solution to equation (17b) with $`q_0=1`$ is $`q_n={\displaystyle \underset{j=1}{\overset{n}{}}}(1j/\omega ).`$ (18) We shall see that $`q_n1`$ unless $`n\omega `$. Therefore, we can write $`\mathrm{ln}q_n={\displaystyle \underset{j=1}{\overset{n}{}}}\mathrm{ln}(1j/\omega )`$ (19) and expand the right hand side for small $`j/\omega `$, yielding<sup>2</sup><sup>2</sup>2One can also write $`q_n=\omega ^n\omega !/(\omega n)!`$ and expand the factorials using Sterling’s approximation. $`q_n=\mathrm{exp}[n(n+1)/2\omega ].`$ (20) The probability $`p_n`$ of obtaining a cycle closure at time $`t=n`$ is then $`p_n={\displaystyle \frac{n}{\omega }}q_{n1}={\displaystyle \frac{n}{\omega }}e^{n(n+1)/2\omega }.`$ (21) To obtain $`𝒫(L)`$, the probability that a given starting point is in the basin of attraction of a cycle of length $`L`$, we note that a closure event at time $`t=n`$ yields with equal probability all cycle lengths up to $`n`$. Therefore, $`𝒫(L)={\displaystyle \underset{n=L}{\overset{\mathrm{}}{}}}{\displaystyle \frac{p_n}{n}},`$ (22) which is well-approximated by $`𝒫(L)`$ $``$ $`{\displaystyle _{x=L}^{\mathrm{}}}{\displaystyle \frac{1}{\omega }}e^{\frac{x(x+1)}{2\omega }}𝑑x`$ (23) $`=`$ $`e^{1/8\omega }{\displaystyle \frac{\pi }{2}}\left(\mathrm{Erf}\left({\displaystyle \frac{1+2L}{2\sqrt{2\omega }}}\right)\right).`$ (24) Note that the probability distribution of cycle lengths is asymptotically Gaussian (for large lengths), and that the length of a typical cycle is of order $`\sqrt{\omega }`$. Derrida has pointed out that the random hopping assumption (or annealed approximation) is exact for the dissipative Kauffman net with $`K=N`$. When $`K<N`$, the random hopping assumption is no longer exact. However, the $`K=N`$ results agree well with the simulational data for large but finite $`K`$. #### 3.2.2 $`K=N`$: Reversible case As we shall see, the behavior of the reversible model in certain parameter regimes is also well-described by a model in which the time-development can be considered to be random until a closure event occurs. However, the analysis of $`K=N`$ limit is more subtle for the reversible model than for the dissipative case. Wrong calculation: leave out special points. To illustrate some of the complications that arise when we consider the reversible model, we first present a naive (and wrong) adaptation to the reversible system of the argument in subsection 3.2.1. Note that in the reversible models, the state at time t, $`𝒮_t`$, depends on the spin configurations, or substates, at two times, $`\mathrm{\Sigma }_{t1}`$ and $`\mathrm{\Sigma }_t`$. We once again consider a sequence of states $`𝒮_0,𝒮_1,𝒮_2,\mathrm{}`$ and assume that the map induces “random hopping” through the state space (each $`𝒮_j`$ chosen with equal probability from all allowed possibilities). If the state $`𝒮_n`$ happens to be the same as $`𝒮_0`$, then the cycle closes, with the cycle-length being $`L=n`$. Note that at the $`n`$th step there is only one possible output that will give closure, $`𝒮_n=𝒮_0`$; the $`n1`$ other values of $`𝒮_j`$, $`1j<n`$, are impossible because each cycle must be traversable both forward and backward. Therefore, if the cycle has not closed in the first $`n1`$ steps, the total number of allowed possibilities for $`𝒮_n`$ is $`\mathrm{\Omega }n`$, of which only one will give closure. This argument yields an estimate for $`\rho _n`$, the probability of closure at the $`n`$th step, given that the orbit has not closed previously: $$\rho _n=\frac{1}{\mathrm{\Omega }n}.$$ (25) Therefore, this estimate implies that $`p_l`$, the probability that the orbit closes at the $`l`$th step, should be $`p_l`$ $`=`$ $`\rho _l{\displaystyle \underset{k=1}{\overset{l1}{}}}\left(1\rho _k\right)`$ (26) $``$ $`{\displaystyle \frac{\mathrm{exp}[l/\mathrm{\Omega }]}{\mathrm{\Omega }}}`$ (27) in the limit of large $`\mathrm{\Omega }`$. However, equation (27) is wrong. Looking back at figure 1, one sees that for $`K=N=10`$, the average cycle length is of order $`\omega =2^N10^3`$. However, equation (27) implies an average cycle length of order $`\mathrm{\Omega }=2^{2N}10^6`$. A more accurate calculation. The problem with equation (27) is that we have ignored the role of the special points. A sequence of substates of the form $`\mathrm{\Sigma }_{t_1},\mathrm{\Sigma }_{t_1+1},\mathrm{},\mathrm{\Sigma }_{t_1+j},\mathrm{\Sigma }^{},\mathrm{\Sigma }^{}`$ is reflected at the twin point and must continue $`\mathrm{\Sigma }_{t_1+j}`$,$`\mathrm{}`$,$`\mathrm{\Sigma }_{t_1+1},\mathrm{\Sigma }_{t_1}`$. Similarly, a sequence of substates of the form $`\mathrm{\Sigma }_{t_1},\mathrm{\Sigma }_{t_1+1},\mathrm{},\mathrm{\Sigma }_{t_1+j},\mathrm{\Sigma }^{},\mathrm{\Sigma }_{t_1+j+2},\mathrm{\Sigma }^{}`$ is reflected at the sandwich point. After the first special point has been hit, the orbit retraces and then continues until a second special point is reached. Once the second special point is reached, say at time $`t=t^{}`$, the orbit is reflected again, and closure in less than $`t^{}`$ additional steps is guaranteed. Since twin points are hit with probability $`1/\omega `$ and sandwiches with probability $`m/\omega `$ at each time step, this mechanism yields orbit lengths of order $`\omega `$ rather than the $`O(\mathrm{\Omega })`$ result of equation (27). In the dissipative case discussed above, when $`K=N`$, the calculation of cycle lengths is exact. We have not been able to do that well in the reversible case. However, we present here a simple approximation for the distribution of cycle lengths that is rather accurate when $`\omega `$ is large and $`l`$ is less than or of the same order as $`\omega `$. We begin by specifying a realization of the network. Given this realization, we consider a sequence of substates $`G_l=\mathrm{\Sigma }_0,\mathrm{\Sigma }_1,\mathrm{\Sigma }_2,\mathrm{},\mathrm{\Sigma }_l,\mathrm{\Sigma }_{l+1}`$ (28) produced by the map. We define a regular sequence to be one which has neither closed nor reached a special point. A regular sequence can be used to construct a part of either a regular or a special cycle. We define the sequence (28) to be a sequence of length $`l`$. Given a regular sequence $`G_l`$, we may produce a $`G_{l+1}`$ by evolving $`G_l`$ for another step. We define the probability $`q(l)`$ as the fraction of the realizations in which $`G_{l+1}`$ is also a regular sequence and the probability $`\rho (l)`$ as the fraction in which it is not. We take $`G_l`$ as a regular sequence and now imagine calculating the next substate $`\mathrm{\Sigma }_{l+2}`$. As an approximation, assume that all $`\mathrm{\Sigma }_{l+2}`$ appear with equal probability, $`\frac{1}{\omega }`$. The probability that $`\mathrm{\Sigma }_{l+2}=\mathrm{\Sigma }_{l+1}`$ (i.e., $`\mathrm{\Sigma }_{l+1}`$ is a twin point) is $`1/\omega `$, and the probability that $`\mathrm{\Sigma }_{l+2}=\mathrm{\Sigma }_l`$ (i.e., $`\mathrm{\Sigma }_{l+1}`$ is a sandwich point) is $`m/\omega `$. There is also a chance $`1/\omega `$ that $`\mathrm{\Sigma }_{l+2}=\mathrm{\Sigma }_0`$. In this last case, the orbit will close with no special points if, in addition, $`\mathrm{\Sigma }_{l+3}=\mathrm{\Sigma }_1`$. Thus, this estimate yields a probability of closure without special points that is of order $`1/\mathrm{\Omega }`$, as in our naive estimate above. Therefore, $`\rho (l)`$, the probability of a closure event at step $`l`$, given that the sequence has not closed previously, is the sum of two terms: the probability for closure to a regular sequence $`\rho ^R1/\mathrm{\Omega }`$, and the probability of getting a special point $`\rho ^S(m+1)/\omega `$, so that $`\rho (l)=\rho ^R(l)+\rho ^S(l)(m+1)/\omega .`$ (29) In addition to ignoring the possibility that $`\mathrm{\Sigma }_{l+2}`$ has already appeared in the sequence, we have ignored terms of relative order $`1/\omega `$. This is quite reasonable for large $`N`$. Now we can derive expressions for the number of cycles of different types. There are $`\mathrm{\Omega }`$ different starting configurations for sequences. We require that the first substate $`\mathrm{\Sigma }_0`$ not be a sandwich substate and that the second substate $`\mathrm{\Sigma }_1`$ not be equal to $`\mathrm{\Sigma }_0`$, Therefore, the fraction of sequences of the form $`\mathrm{\Sigma }_0,\mathrm{\Sigma }_1`$ (e.g., $`l=0`$) that are regular is $`(11/\omega )(1m/\omega )(1(m+1)/\omega )`$. Each iteration reduces the fraction of sequences that are regular by a factor of $`1\rho `$, until after $`l`$ steps we find that the number of regular sequences, $`N_{RS}(l)`$, is (where again we disregard terms of relative order $`l/\omega `$ and smaller) $`N_{RS}(l)=\mathrm{\Omega }(1(m+1)/\omega )(1\rho )^l\mathrm{\Omega }e^{(m+1)l/\omega }.`$ (30) Since the probability of closure to a regular cycle is $`1/\mathrm{\Omega }`$, the probability that a randomly chosen point in the phase space is part of a regular cycle that closes in $`l`$ steps, $`P_R(l)`$, is $`P_R(l)=N_{RS}(l)/\mathrm{\Omega }`$. Because each cycle of length $`l`$ is found by starting at any of $`l`$ points on the cycle, the average number of regular cycles which close after $`l`$ steps is $`N_R(l)={\displaystyle \frac{N_{RS}(l)}{\mathrm{\Omega }l}}l^1e^{(m+1)l/\omega }.`$ (31) A very small proportion of the points in the state space are, in fact, parts of regular cycles. The number which take part in regular cycles of all lengths, $`\text{M}_R`$, is $`\text{M}_R={\displaystyle \underset{j=0}{\overset{\mathrm{}}{}}}jN_R(j){\displaystyle \frac{\omega }{m+1}}.`$ (32) This number is indeed much smaller than the state space volume $`\mathrm{\Omega }`$. We now turn our attention to the special cycles, which dominate the state space in this high $`K`$ limit. All of them can be found by starting at a special point and iterating until a second special point is reached after $`l`$ steps. Then the orbit reverses itself and closes after an additional $`l`$ steps. There are three kinds of special cycles, twin-twin, sandwich-sandwich, and twin-sandwich. There are $`\omega m`$ different ways to choose the initial point if it is a sandwich point and $`\omega `$ ways to choose it if it is a twin point, so the number of twin-twin cycles and sandwich-sandwich cycles of length $`l`$ are: $`N_{tt}(l)={\displaystyle \frac{1}{2}}e^{(m+1)l/(2\omega )}`$ (33a) $`N_{ss}(l){\displaystyle \frac{m^2}{2}}e^{(m+1)l/(2\omega )}.`$ (33b) The factors of two arise in equation (33) because each cycle of these types is found twice by this method. Similarly, the number of odd (sandwich-twin) special cycles is $`N_{st}(l)me^{(m+1)l/(2\omega )}.`$ (33c) Note that this estimate for the distribution of special cycle lengths depends exponentially on $`l`$, in contrast to the Gaussian dependence for the dissipative model. To check these conclusions we plot in figure 3 $`N(l)`$, the number of cycles of length $`l`$ as a function of $`l`$ averaged over realizations. The realizations used all had $`m=0`$. The theoretical estimates (equations 31 and 33) agree very well with the numerical results. Average over $`𝐦`$. Our results of equations (31) and (33) depend upon the number of sandwich special points, $`m`$. In this section, we shall denote averages over $`m`$ by $`<>`$. In Appendix A we show that for $`K=N`$, the probability distribution for $`m`$ is a Poisson distribution $`\rho (m)={\displaystyle \frac{e^\lambda \lambda ^m}{m!}}\text{ with }\lambda =1.`$ (34) Averaging equation (31) using the weight defined in equation (34) gives that the realization average of the number of regular cycles of length $`l`$, $`<N_R(l)>`$, is $`<N_R(l)>=l^1\mathrm{exp}\left[(e^{l/\omega }1)l/\omega \right].`$ (35a) Similarly, for the various kinds of special cycles $`<N_{tt}(l)>`$ $`=`$ $`{\displaystyle \frac{1}{2}}\mathrm{exp}\left[(e^{l/(2\omega )}1)l/(2\omega )\right]`$ (35b) $`<N_{ts}(l)>`$ $`=`$ $`\mathrm{exp}\left[(e^{l/(2\omega )}1)l/\omega \right]`$ (35c) $`<N_{ss}(l)>`$ $`=`$ $`{\displaystyle \frac{e^{l/(2\omega )}+1}{2}}\mathrm{exp}\left[(e^{l/(2\omega )}1)l/\omega \right].`$ To test equation (35) against simulations, in figure 4 we plot $`N_S(l)`$, the number of special cycles averaged over realizations, against $`l`$. As discussed in subsection (2.2), special cycles formed by two twin points or two sandwich points have an even length, while the ones with one twin point and one sandwich point have an odd length. Therefore, $`N_S(l)=<N_{ss}(l)>+<N_{tt}(l)>`$ if $`l`$ is even and $`N_S(l)=<N_{ts}(l)>`$ if $`l`$ is odd. We find that $`N_S(l)`$ oscillates because of this difference between even cycle lengths and odd cycle lengths, leading to a two-branch structure in $`N_S(l)`$. This structure was indeed observed in our simulations. The random hopping analysis of the $`K=N`$ situation gives a Hamming-distance behavior which is exactly the same in the dissipative and reversible systems. In both cases, the systems start from a pair of configurations which differ in the value of one spin at time $`t`$. By the next time-step, the configuration will be random, so that half the spins will be “wrong”. Thus the Hamming distance immediately comes to the value $`N/2`$, and stays there. ## 4 Intermediate $`K`$ Section 3 outlines both qualitative and quantitive pictures of the limiting cases $`K=0`$ and $`N`$. In this section, we will discuss the system’s behavior for intermediate $`K`$ values. As $`K`$ is increased, there is an evolution from dynamically independent clumps of spins to a situation in which there is random hopping over the whole state-space. Figure 5 shows numerical results of $`S_S(l)`$ and $`S_R(l)`$, the survival functions for special cycles and regular cycles separately, for various intermediate $`K`$ values. Recall that a survival function of cycle length $`l`$ is defined to be the probability of observing cycles with length greater than $`l`$ (see equation (5)). Noninteger values of $`K`$ are produced by mixtures of spins with the neighboring integer $`K`$-values. The survival functions for special cycles, $`S_S(l)`$, as shown in part (A), shows no surprising structure. As $`K`$ decreases from 2, the observed cycle lengths get shorter and shorter. For $`K3`$, the survival probabilities approach in a uniform manner the result of equation (35). The regular cycles are different. For $`K>5`$ we see a uniform approach to the $`K=N`$ result. At $`K=1.6`$, there seems to be a power law behavior with $`S_R(l)l^1`$. Then for larger values of $`K`$, $`S_R(l)`$ seems to decrease faster than power law. Finally, $`K=3`$ and $`K=4`$ show a remarkable bump at large values of $`l/2^N`$, indicating that there is a new process going on at large $`l`$. The difference between part (A) and (B) implies that the relative importance of the two pieces of cycle-closing mechanism depends strongly on K. Moreover, the relative numbers of regular and special cycles are also strongly $`K`$-dependent. This point is illustrated in figure 6, which shows the ratio of the number of regular cycles to that of special cycles as a function of $`K`$. For small $`K`$, several subsets of the spins may become uncoupled, and most cycles are regular cycles. Conversely special cycles are more likely for large $`K`$. ### 4.1 Anomalously Long Regular Cycles for Intermediate $`K`$ Values Notice the bump in figure 5 part (B) for $`K=3`$ and $`K=4`$ in the region $`1\widehat{l}\omega `$. This bump arises from a group of regular cycles which are anomalously long. More careful study shows that in fact regular cycles split into two groups, which scale differently. To demonstrate this effect, we plot in figure 7 the survival functions $`S_R(l)`$ for $`K=3`$ and for a range of $`N`$ values against the cycle length $`l`$ normalized by $`2^N`$. Manifestly the distribution of short cycles scales with $`\omega =2^N`$. This scaling can be explained in the following way. There are $`(m+1)2^N`$ special points in the state-space of the system, as discussed in section 2.3. To obtain a regular cycle, the system must not hit any of the special points. The probability of hitting a special point $`1/2^N`$, so the lengths of the regular cycles scale as $`2^N`$. For a more explicit and quantitative reasoning, we may appeal to the regular cycle result based on the random hopping assumption described by equation (35), which provides a reasonably accurate approximation of the system when the system is sufficiently well-connected. Note that in equation (35), the regular cycle distribution clearly scales as $`2^N`$. While we have presented an intuitive explanation for the scaling of the relatively short regular cycles, one may still wonder why there is a population of anomalously long cycles and why it appears and disappears as $`K`$ increases. To attack these issues, a number of realizations which produce extra long cycles were studied and similarities among them were observed. In general, a realization and the initial configuration have to satisfy the following two conditions to be able to generate an anomalously long cycle: The realization should not have any sandwich points, thereby annulling the mechanism to close a cycle using them, and the functions assigned to the spins together with the choice of initial configurations should prevent a system from hitting a twin point. For small $`K`$ values, one can easily find realizations without any sandwich points. For instance, if any spin is assigned the function that is “-1” for all inputs, then the realization has no sandwich points. In fact, in Appendix A, we prove that almost all realizations have no sandwich points for finite $`K`$ and large enough $`N`$ ($`2^{(2^K)}`$). The twin points in the cycle could be avoided in many ways. For example, when a function assigned to a spin equals to the constant $`+1`$, if this spin starts from $`\sigma _0=+1`$ and $`\sigma _1=1`$, it will follow the progression of $`+1`$, $`1`$, $`+1`$, $`1`$, … , and the system never hits a twin point since $`\sigma _t\sigma _{t+1}`$ always holds. Since there is a probability $`1/2^{(2^K)}`$ that any given spin is assigned the function “$`+1`$”, and a probability $`1/2`$ that this spin starts with $`\sigma _0=\sigma _1`$, the probability $`p_2`$ that at least one spin is in the “$`++`$” progression is $`p_2`$ $`=`$ $`1(1({\displaystyle \frac{1}{2}})({\displaystyle \frac{1}{2^{(2^K)}}}))^N`$ (36) $``$ $`1\mathrm{exp}({\displaystyle \frac{N}{2^{(2^K1)}}})(2^{(2^K)}1).`$ Thus, when $`N2^{(2^K)}`$, almost all starting configurations and realizations will not hit a twin point. Another way to avoid twin points consists of two spins that are assigned identical inputs and functions. In this case, it is possible to choose an initial configuration for these two spins, which will stop the system from forming a twin point. For instance, a system in which spin 1 and spin 2 are assigned identical input spins and functions starting from $`\sigma _{t=0}^1=\sigma _{t=1}^1=1`$ and $`\sigma _{t=0}^2=1`$, $`\sigma _{t=1}^2=1`$ can never hit a twin point. It is also easy to prescribe certain simple progressions for a few spins and stop the system from hitting a twin point. For example, two spins both with period $`3`$ evolving following the pattern $`\sigma _0:++++++\mathrm{}`$ $`\sigma _1:++++++\mathrm{}`$ can prevent the system from forming a special cycle. The mechanisms preventing the system from hitting a twin point always appear to be related to some small piece of the system that evolves independently from other parts of the system. The couplings are such that this piece is not affected by anything outside itself (though in general it can and does affect the rest of the system). We call a piece like this a local structure. Local structures are to be discussed in detail in the next paper in this series. Note that local structures are very unlikely for sufficiently large $`K`$ values, where the spins of the system are quite correlated, unless $`N`$ is enormous. When $`K`$ is small, local structures occur much more frequently. When many a local structures are present, almost all the realizations and initial configurations have no special points. Thus the presence of local structures leads to very long regular cycles. Since for a given $`N`$ the local structures become less probable as $`K`$ increases, it is natural to find the number of regular cycles decreases as $`K`$. Section 3.2.2 presents a naive theory of hopping in state space that ignores the role of the special points and predicts that the distribution of orbit lengths in a system of $`N`$ spins should scale as $`2^{2N}`$. This naive theory fails qualitatively when $`N=K`$ because the special points induce orbit closure in order $`2^N`$ steps. However, if local structures prevent the system from ever reaching a special point, then the mechanism for closing the orbits in order $`2^N`$ steps does not operate and it is plausible that the typical orbit lengths will be much longer than $`2^N`$. We test this idea by calculating orbit lengths in systems with one spin assigned the function $`1`$ (so it follows the sequence $`(+1,+1,1,1,+1,+1,\mathrm{})`$) and a second individual spin assigned the function $`+1`$ and the initial condition $`(+1,1)`$ (so its evolution is $`(+1,1,+1,1,\mathrm{})`$). Such $`(+,+,,),(+,,+,)`$ systems can never reach a special point. As figure 8 demonstrates, in these systems orbit lengths grow with $`N`$ much faster than $`2^N`$; the numerical results are consistent with $`2^{2N}`$ scaling. We believe that the orbit lengths in systems with $`N2^{2^K}`$ and randomly chosen functions cannot grow faster than $`2^{2N(1ϵ(K))}`$, where $`ϵ(K)`$ is of order $`2^{2^K}`$. This is because of order $`N2^{2^K}`$ spins have input functions $`1`$ or $`1`$ and hence cycle with periods $`1,`$ $`2`$, and $`4`$. More generally, we expect interesting crossover phenomena to occur when $`K`$ is both large and of the order of $`log_2(log_2N)`$. Even for $`K=3`$, simulating systems with large enough $`N`$ to permit numerical exploration of these effects is computationally prohibitive. ### 4.2 Average cycle length versus $`N`$ for different values of $`K`$ When $`K`$ is large, the random hopping approximation works well and the cycle length distribution scales as $`\omega =2^N`$. Figure 5 demonstrates that the distribution of cycle lengths does not change dramatically while K decreases until $`K`$ is quite small. Therefore, it is reasonable to expect the average cycle length to increase exponentially with $`N`$ when $`K`$ is large. On the other hand, when $`K=0`$, the average cycle length is bounded above by 4. For $`K=1`$, our simulations and analytic arguments indicate that it scales as $`\mathrm{log}N`$; the results will be presented in the companion publication. For values of $`K`$ in the range \[1.4, 1.7\], the survival functions $`S(l)`$ decay roughly as a power law over three decades in cycle length $`l`$, as shown in figure 9. This result is consistent with the presence of a phase transition. The companion paper on the behavior of the Hamming distance presents more compelling evidence for a phase transition at $`K1.6`$. ## 5 Summary This paper addresses the dynamics of a Boolean network model of $`N`$ elements with $`K`$ inputs with time-reversible dynamics. We present the general setup of the model and introduce the concept of special points and the distinction between special cycles and regular cycles. The relation between special points and the properties of the cycles is demonstrated. We show that the numbers of special points and special cycles for each realization are proportional to $`\omega 2^N`$, where $`N`$ is the number of variables in the system. We determine the probability distribution of cycle lengths as well as the survival functions. In limiting case $`K=0`$, the cycle length is bounded above by 4 and the probability that a cycle length is 4 approaches 1 as $`N`$ increases. For $`K=N`$, within a random hopping approximation we calculate the survival functions for regular cycles and for special cycles, which agree with simulational data extremely well. Finally, we present the simulational results for survival functions for intermediate $`K`$ values. A population of anomalously long regular cycles scaling as $`2^{2N}`$ is found for small $`K`$ values and explained based on the notions of special points and local structures. The correlation between typical cycle length and the K values of the networks is studied, and we find that the typical cycle length increases logarithmically with $`N`$ when $`K<1.4`$, exponentially when $`K>1.7`$, and following a power law when $`K`$ falls in between; these results are compatible with the presence of a phase transition for $`K`$ somewhere in the range of $`[1.4,1.7]`$. ## Appendix A: Some statistical properties of sandwich points The text discusses twin and sandwich symmetry points which induce closure of orbits in the reversible Kauffman model. Each type arises when the substate $`\mathrm{\Sigma }_t`$ is such that each of the functions of these inputs takes on a target value. For a twin point, $`\sigma _{t+1}^j=\sigma _t^j`$, so the target function is different at every time step. For a sandwich point, for all $`t`$ the target function is $`F^i=1`$ for every $`i`$. To calculate orbit lengths, we need to compute the probability that a symmetry point of either type occurs at each time $`t`$. For a given realization of couplings, the number of substates for which the functions take on a particular value can vary. Because the target value for the twin point is different for different substates, whereas the target values for the sandwich points are the same at all times, the statistics of the two types of symmetry points are different. The process we consider is one in which couplings and an initial condition are chosen, and then the system evolves in time. We assume that this time evolution yields a random sampling of all possible spin configurations or substates. At each time $`t`$ we examine whether a symmetry point of either type has been reached. Let $`m`$ be the number of substates for which $`F^j=1`$ for all $`j`$ (the criterion for a sandwich point), and $`m_t`$ be the number of spin configurations for which each function takes on the target value for a twin point at time $`t`$. At a time $`t`$, the probability of being at a twin point is $`m_t/\omega `$, whereas the probability of being at a sandwich point is $`m/\omega `$. On average it takes many trials before a special point is reached; the probability of having observed a sandwich point after $`T`$ trials $`_{t=1}^Tm_t/\omega mT/\omega `$ = $`T/\omega `$, where $``$ is the average over realizations, whereas the probability of having observed a sandwich symmetry point is $`mT/\omega `$. We wish to calculate the probability that a randomly chosen realization has a given value of $`m`$. Now each output takes on a given value with probability $`1/2`$, so on average the probability that $`N`$ outputs all have given target values is $`(1/2)^N`$, implying that the realization average $`m=2^N`$. However, if one of the functions happens to be $`F^i=1`$, then clearly there are no sandwich points. We wish to find $`P_K(m)`$, the fraction of all possible realizations of the couplings for a given $`K`$ that yield each value of $`m`$. First we consider $`K=N`$. This case is particularly relevant because when $`K`$ is large, essentially all the orbits close because of the symmetry points. Here, the functions can be viewed as mapping a given input substate into a randomly chosen output substate. Since there are $`\omega =2^N`$ possible output substates, of which one is the target, each input configuration has a probability $`1/\omega `$ of having its output be the target. $`P_{K=N}(0)`$, the probability that no input configurations has as its output the target configuration, is $`(12^N)^{2^N}`$, which, as $`N\mathrm{}`$, approaches $`1/e`$. The probability that exactly one of the $`2^N`$ different inputs yields the target output is $`(2^N)(2^N)(12^N)^{2^N1}1/e`$. Similarly, $`P_{K=N}(m)`$, the probability that exactly $`m`$ of the $`2^N`$ different inputs yields the target output is $`P_{K=N}(m)`$ $`=`$ $`{\displaystyle \frac{\left(2^N\right)!}{m!(2^Nm)!}}\left(2^N\right)^m\left(12^N\right)^{2^Nm}`$ (37) $``$ $`{\displaystyle \frac{1}{m!e}}\text{(}N\mathrm{}\text{}m2^N\text{)}.`$ Thus, $`P_{K=N}(m)`$ is a Poisson distribution. Since $`P_{K=N}(m)`$ falls off quickly as $`m`$ gets large, clearly it is consistent to assume that $`m2^N`$. However, when $`K`$ is finite and $`N`$ is large enough, we expect the behavior of $`P_K(m)`$ to differ qualitatively from the $`K=N`$ result. We expect that almost all configurations will have $`m=0`$ for any finite $`K`$ as $`N\mathrm{}`$. We have argued before that if in a realization a spin is assigned a function that equals to constant -1, the realization has no sandwich point. Also, if two spins are assigned functions $`F^1`$ and $`F^2`$ such that $$F^1+F^2=0$$ for all inputs, there can be no sandwich point for the realization. There are many other possible mechanisms that lead to $`m=0`$. Clearly, the probability that a realization has at least one spin function of -1 bounds below the probability that it has no sandwich point. Among $`2^{(2^K)}`$ possible functions that can be assigned to one spin, one is $`1`$ for all inputs. Assuming that all functions are equally likely to be picked and that the function choices for different spins are independent, the probability that no spin is assigned the constant function -1 is $$(1\frac{1}{2^{(2^K)}})^N\mathrm{exp}(\frac{N}{2^{(2^K)}})(N2^{(2^K)}).$$ Thus, $`P_K(m=0)`$ is bounded by the probability that the realization has at least one function that is $`1`$; or $`P_K(m=0)(1\mathrm{exp}(\frac{N}{2^{(2^K)}}))`$. This result implies that whenever $`K`$ is finite, in a large enough system sandwich points cause orbit closure only in a vanishingly small fraction of realizations. However, when $`K`$ is not small, realizations with sandwich points are rare only when $`N`$ is enormous (when $`N2^{2^K}`$). ## Appendix B: Relation between special points and cycles Here we prove the results used in section 2 that 1) each special cycle contains exactly two special points, 2) that cycles with two special points of the same kind have even cycle lengths, and 3) cycles with different kinds of special points have odd lengths. To prove that each special cycle contains two and only two special points, we first consider a cycle of even length $`2n`$, $$\mathrm{\Sigma }_0,\mathrm{\Sigma }_1,\mathrm{},\mathrm{\Sigma }_{2n1}.$$ Suppose there is a twin point in the cycle. By relabeling the cycle, we can get $$\mathrm{\Sigma }_{n1}=\mathrm{\Sigma }_n$$ by definition. Now $`\mathrm{\Sigma }_{n1t}=\mathrm{\Sigma }_{n+t}`$ since the cycle is time reversible, so that $`\mathrm{\Sigma }_0=\mathrm{\Sigma }_{2n1}`$ when we take $`t=n1`$. Thus we find another twin point in this even-length cycle. If there is a sandwich point at $`n`$, then $$\mathrm{\Sigma }_{n1}=\mathrm{\Sigma }_{n+1}$$ and $`\mathrm{\Sigma }_{n1t}=\mathrm{\Sigma }_{n+1+t}`$, thus $$\mathrm{\Sigma }_1=\mathrm{\Sigma }_{2n1}.$$ Here we find another sandwich point at 0. Similarly, when the cycle is odd, we will find a twin point in the presence of a sandwich point, and vice versa. By now we have proven that if there is a special point in the cycle, then there has to be another. The statement that the cycle length being odd or even depends on whether the special points are of the same kind, is also clear from the above argument. To finish the proof, we need to demonstrate that no orbit can contain more than two special points. Assume that an orbit of length $`L`$ with more than two special points exists. Choose the origin of time so that $`\mathrm{\Sigma }_{Lj}=\mathrm{\Sigma }_j`$ (one does this by placing a sandwich substate at $`t=0`$ or placing twin substates at $`t=1`$ and $`t=L`$), and let $`P`$ be the smallest value for which $`\mathrm{\Sigma }_{Pn}=\mathrm{\Sigma }_{P+n}`$ for all $`n`$<sup>3</sup><sup>3</sup>3If the special point at $`P`$ is a twin point, then $`2P`$ is odd, otherwise it is even.; by assumption, $`P<L/2`$. Then we must have simultaneously $`\mathrm{\Sigma }_{Lj}=\mathrm{\Sigma }_j`$ and $`\mathrm{\Sigma }_{Pn}=\mathrm{\Sigma }_{P+n}`$. Letting $`j=Pn`$ yields $`\mathrm{\Sigma }_{LP+n}=\mathrm{\Sigma }_{Pn}=\mathrm{\Sigma }_{P+n}`$. Letting $`q=P+n`$, we obtain $`\mathrm{\Sigma }_{L2P+q}=\mathrm{\Sigma }_q`$. Thus the orbit period is $`L2P`$, which is strictly less than $`L`$, so we have a contradiction. We can also show that these two special points must be different from one another. Suppose not. Consider an orbit of length $`L`$, and choose the origin of time so that $`𝒮_0`$ and $`𝒮_P`$ are the same special point, with, by assumption, $`P<L1`$. Applying the map yields $`𝒮_j=𝒮_{(P+j)}`$ for any $`j`$, so the orbit repeats after $`P`$ steps. But this contradicts the assumption that $`L`$ is strictly greater than $`P`$.
warning/0004/physics0004005.html
ar5iv
text
# Lorentz-Covariant Hamiltonian Formalism. ## I Introduction A remarkable formulation of classical dynamics is provided by Hamiltonian mechanics. This is an old subject. However, new discoveries are still been made; we quote two examples among several: the Arnold duality transformations, which generalize the canonical transformations , and the extensions of the Poisson brackets to differential forms and multi-vector fields by A.Cabras and M.Vinogradov . In this context the transition from classical to relativistic mechanics raises the question of Hamiltonian covariance, the physical significance of which is discussed by Goldstein . In the first part of this paper we briefly recall the Poisson brackets approach and the covariant Hamiltonian formalism. Then we introduce new brackets to study the dynamics associated to this covariant Hamiltonian, which define an algebraic structure between position and velocity, and does not have an explicit formulation. We examine the close link between these brackets and those used by Feynman for his derivation of the Maxwell equations . A very interesting way to arrive at the same sort of result was found by Souriau in the frame of his symplectic classical mechanics . In the final part of this work we consider the dynamics in curved space, using Christoffel symbols, covariant derivatives, and curvature tensors expressed in terms of these brackets. ## II Brief review of Analytic Mechanics ### A Poisson brackets The dynamics of a classical particle in a 3-dimensional flat space with vector position $`q^i`$ and vector momentum $`p_i`$ ($`i=1,2,3)`$ is defined by the Hamilton equations: $$\{\begin{array}{c}\stackrel{i}{\stackrel{.}{q}}=\frac{dq^i}{dt}=\frac{H}{p_i}\\ \\ \underset{i}{\overset{.}{p}}=\frac{dp_i}{dt}=\frac{H}{q^i}\end{array}$$ (1) where the Hamiltonian $`H(q^i,p_i)`$ is a form on the phase space ( the cotangent fiber space). They can be also expressed in a symmetric manner by means of Poisson brackets: $$\{\begin{array}{c}\stackrel{i}{\stackrel{.}{q}}=\{q^i,H\}\\ \\ \underset{i}{\overset{.}{p}}=\{p_i,H\}\end{array}$$ (2) These brackets are naturally defined as skew symmetric bilinear maps on the space of functions on the phase space in the following form: $$\{f,g\}=\frac{f}{q^i}\frac{g}{p_i}\frac{g}{q^i}\frac{f}{p_i}$$ (3) ### B Covariant Hamiltonian Except in the electromagnetic situation, the Hamiltonian is not the total energy when it is time-dependent, and its generalization to relativistic problems with the $`M_4`$ Minkowski space is not trivial because it is not Lorentz covariant. In the electromagnetic case the answer to this question is given by the introduction of the following covariant expression : $$H=u^\mu p_\mu L=u^\mu (mu_\mu +\frac{q}{c}A_\mu )\text{ }$$ (4) where $`L`$ is the usual invariant electromagnetic Lagrangian : $$L=\frac{1}{2}m\text{ }u^\mu u_\mu +\frac{q}{c}\text{ }u^\mu A_\mu $$ (5) and $`u^\mu `$ the quadri-velocity defined by means of the proper time $`t_p`$, here used as an invariant parameter: $$u^\mu =\frac{dx^\mu }{dt_p}$$ (6) Finally we have the covariant Hamiltonian $$H=\frac{1}{2}m\text{ }u^\mu u_\mu $$ (7) with the corresponding eight Hamilton equations: $$\{\begin{array}{c}\frac{H}{p_\mu }=\frac{dx^\mu }{dt_p}=u^\mu \\ \\ \frac{H}{x^\mu }=\frac{dp_\mu }{dt_p}\end{array}$$ (8) It is interesting to recall that this structure is only possible in the situation where the potential can be put in a covariant manner as in the electromagnetism theory. ## III Lorentz covariant Hamiltonian and brackets formalism Now we want to generalize the relation between the usual non covariant relativistic Hamiltonian and the Poisson brackets to a covariant Hamiltonian $`H`$ and new formal brackets introduced in the frame of the Minkowski space. It is important to remark that, in a different manner, P.Bracken also studied the relation between this Feynman problem and the Poisson brackets. In this context a ”dynamic evolution law” is given by means of a one real parameter group of diffeomorphic transformations : $`g\text{ }(\text{ }IR\text{ }\times \text{ }M_4\text{ })M_4:\text{ }g(\tau ,x)=g^\tau x=x(\tau )`$ The ”velocity vector” associated to the particle is naturally introduced by: $$\stackrel{\mu }{\stackrel{.}{x}}=\frac{d}{d\tau }g^\tau x^\mu $$ (9) where the ”time” $`\tau `$ is not identified with the proper time as we shall see later. The derivative with respect to this parameter of an arbitrary function defined on the tangent bundle space can be written, by means of the covariant Hamiltonian, as: $$\frac{df(x,\stackrel{.}{x},\tau )}{d\tau }=[H,\text{ }f(x,\stackrel{.}{x},\tau )]+\frac{f(x,\stackrel{.}{x,\tau })}{\tau }$$ (10) where for $`H`$ we take the following definition: $$H=\frac{1}{2}\text{ }m\text{ }\frac{dx^\mu }{d\tau }\text{ }\frac{dx_\mu }{d\tau }=\frac{1}{2}\text{ }m\stackrel{\mu }{\stackrel{.}{x}}\underset{\mu }{\overset{.}{x}}$$ (11) Equation (10) giving the dynamic of the system, is the definition of our new brackets structure, and is the fundamental equation of the paper. We require for these new brackets the usual first Leibnitz law: $$[A,BC]=[A,B]C+[A,C]B$$ (12) and the skew symmetry: $$[A,B]=[B,A]$$ (13) where the quantities $`A,`$ $`B`$ and $`C`$ depend of $`x^\mu `$ and $`\stackrel{\mu }{\stackrel{.}{x}}`$. In the case of the vector position $`x^\mu (\tau )`$ we have from (10): $$\stackrel{\mu }{\stackrel{.}{x}}=[H,x^\mu ]=m\text{ }[\stackrel{\nu }{\stackrel{.}{x}},x^\mu ]\underset{\nu }{\overset{.}{x}}$$ (14) and we easily deduce that: $$m\text{ }[\stackrel{\nu }{\stackrel{.}{x}},x^\mu ]=g^{\mu \nu }$$ (15) where $`g^{\mu \nu }`$ is the metric tensor of the Minkowski space. As in the Feynman approach the time parameter is not the proper time. To see this we borrow Tanimura’s argument . Consider the relation $$g^{\mu \nu }\frac{dx^\mu }{dt_p}\frac{dx^\nu }{dt_p}=1$$ (16) which implies $$[\stackrel{\lambda }{\stackrel{.}{x}},g^{\mu \nu }\frac{dx^\mu }{dt_p}\frac{dx^\nu }{dt_p}]=0.$$ (17) and is in contradiction with: $$[\stackrel{\lambda }{\stackrel{.}{x}},g^{\mu \nu }\frac{dx^\mu }{d\tau }\frac{dx^\nu }{d\tau }]=\frac{2}{m}\stackrel{\lambda }{\stackrel{.}{x}}.$$ (18) But differently from Feynman, the fact that $`g^{\mu \nu }`$ is the metric is a consequence of the formalism and is not imposed by hand. In addition, contrary to Feynman, we do not need to impose the Leibnitz condition: $$\frac{d}{d\tau }[A,B]=[\frac{dA}{d\tau },B]+[A,\frac{dB}{d\tau }]$$ (19) ($`A`$ and $`B`$ being position- and velocity-dependent functions) because the time derivative is given by the fundamental equation (10). We impose the usual locality property: $$[x^\mu ,x^\nu ]=0$$ (20) which directly gives for an expandable function of the position or the velocity the following useful relations: $$\{\begin{array}{c}[x^\mu ,f(\stackrel{.}{x})]=\frac{1}{m}\frac{f(\stackrel{.}{x})}{\underset{\mu }{\overset{.}{x}}}\\ \\ [\stackrel{\mu }{\stackrel{.}{x}},f(x)]=\frac{1}{m}\frac{f(x)}{x_\mu }\end{array}$$ (21) which reduce in the particular cases of the position and velocity to: $$\{\begin{array}{c}[x^\mu ,\stackrel{\nu }{\stackrel{.}{x}}]=\frac{1}{m}g^{\mu \rho }\frac{\stackrel{\nu }{\stackrel{.}{x}}}{\stackrel{\rho }{\stackrel{.}{x}}}=\frac{1}{m}\frac{\stackrel{\nu }{\stackrel{.}{x}}}{\underset{\mu }{\overset{.}{x}}}=\frac{g^{\mu \nu }}{m}\\ \\ [\stackrel{\mu }{\stackrel{.}{x}},x^\nu ]=\frac{1}{m}g^{\mu \rho }\frac{x^\nu }{x^\rho }=\frac{1}{m}\frac{x^\nu }{x_\mu }=\frac{g^{\mu \nu }}{m}\end{array}$$ (22) To compute the bracket between two components of the velocity we require in addition the Jacobi identity: $$[[\stackrel{\mu }{\stackrel{.}{x}},\stackrel{\nu }{\stackrel{.}{x}}],x^\rho ]+[[x^\rho ,\stackrel{\mu }{\stackrel{.}{x}}],\stackrel{\nu }{\stackrel{.}{x}}]+\left[[\stackrel{\nu }{\stackrel{.}{x}},x^\rho ]\stackrel{\mu }{\stackrel{.}{x}}\right]=0$$ (23) which by using (15) gives: $$[\stackrel{\mu }{\stackrel{.}{x}},\stackrel{\nu }{\stackrel{.}{x}}]=\frac{N^{\mu \nu }(x)}{m}$$ (24) where $`N^{\mu \nu }(x)`$ is a skew symmetric tensor. The second derivative of the position vector is: $$\stackrel{\mu }{\stackrel{..}{x}}=\frac{d\stackrel{\mu }{\stackrel{.}{x}}}{d\tau }=[H,\stackrel{\mu }{\stackrel{.}{x}}]=N^{\mu \nu }\underset{\nu }{\overset{.}{x}}$$ (25) and we write: $$F^{\mu \nu }=\frac{m}{q}N^{\mu \nu }$$ (26) in order to recover the Lorentz equation of motion. remark 1. We can easily calculate: $$[H,H]=\frac{1}{4}m^2[\underset{\mu }{\overset{.}{x}}\stackrel{\mu }{\stackrel{.}{x}},\underset{\nu }{\overset{.}{x}}\stackrel{\nu }{\stackrel{.}{x}}]=\text{ }q\underset{\mu }{\overset{.}{x}}\underset{\nu }{\overset{.}{x}}F^{\mu \nu }=0$$ (27) and then deduced: $$\frac{dH}{d\tau }=\frac{H}{\tau }$$ (28) which is the expected result. In the same manner, we get for the $`4`$-orbital momentum: $`{\displaystyle \frac{dL^{\mu \nu }}{d\tau }}`$ $`=`$ $`m{\displaystyle \frac{d}{d\tau }}(x^\mu \stackrel{\nu }{\stackrel{.}{x}}\stackrel{\mu }{\stackrel{.}{x}}x^\nu )=m(x^\mu \stackrel{\nu }{\stackrel{..}{x}}\stackrel{\mu }{\stackrel{..}{x}}x^\nu )`$ (29) $`=`$ $`q(x^\mu F^{\nu \rho }\underset{\rho }{\overset{.}{x}}x^\nu F^{\mu \rho }\underset{\rho }{\overset{.}{x}})=[H,L^{\mu \nu }]`$ (30) as expected. ## IV Maxwell equations Our formal construction will give the Maxwell equations because it leads to the fundamental result (15) which is the starting point of Feynman’s proof of the first group of Maxwell equations. The difference is that our main property is equation (10) and not the Leibnitz rule (19). So our derivation will be obtained differently and will give in addition the two groups of Maxwell equations. $``$ To be general, we choose like in , the following definition for the gauge curvature: $$[\stackrel{.}{x^\mu },\stackrel{.}{x^\nu }]=\text{ }\frac{1}{m^2}(qF^{\mu \nu }+g^{}F^{\mu \nu })$$ (31) where g will be interpreted as the magnetic charge of the Dirac monopole, the \*-operation being the Hodge duality. $``$ A simple derivative gives: $$\frac{d(qF^{\mu \nu }(x)+g^{}F^{\mu \nu }(x))}{d\tau }=q^\rho F^{\mu \nu }(x)\underset{\rho }{\overset{.}{x}}+g^{}_{}{}^{\rho }F^{\mu \nu }(x)\underset{\rho }{\overset{.}{x}}$$ (32) and by means of the fundamental relation (10) we obtain: $`{\displaystyle \frac{d(qF^{\mu \nu }(x)+g^{}F^{\mu \nu }(x))}{d\tau }}`$ $`=`$ $`[H,qF^{\mu \nu }(x)+g^{}F^{\mu \nu }(x)]`$ (33) $`=`$ $`{\displaystyle \frac{m^3}{q}}\text{ }[\stackrel{\rho }{\stackrel{.}{x}},[\stackrel{\mu }{\stackrel{.}{x}},\stackrel{\nu }{\stackrel{.}{x}}]]\underset{\rho }{\overset{.}{x}}`$ (34) Now using the Jacobi identity we rewrite this expression as: $`{\displaystyle \frac{d(qF^{\mu \nu }(x)+g^{}F^{\mu \nu }(x))}{d\tau }}`$ $`=`$ $`{\displaystyle \frac{m^3}{q}}([\stackrel{\mu }{\stackrel{.}{x}},[\stackrel{\nu }{\stackrel{.}{x}},\stackrel{\rho }{\stackrel{.}{x}}]]\underset{\rho }{\overset{.}{x}}+\text{ }[\stackrel{\nu }{\stackrel{.}{x}},[\stackrel{\rho }{\stackrel{.}{x}},\stackrel{\mu }{\stackrel{.}{x}}]]\underset{\rho }{\overset{.}{x}})\underset{\rho }{\overset{.}{x}}`$ (35) $`=`$ $`q(^\mu F^{\nu \rho }+\text{ }^\nu F^{\rho \mu }x)\underset{\rho }{\overset{.}{x}}g(^\mu {}_{}{}^{}F_{}^{\nu \rho }+\text{ }^\nu {}_{}{}^{}F_{}^{\rho \mu }x)\underset{\rho }{\overset{.}{x}}`$ (36) By comparing equations (32) and (LABEL:dfdt3) we deduce the following field equation: $$q(^\mu F^{\nu \rho }+^\nu F^{\rho \mu }+^\rho F^{\mu \nu })+g(^\mu {}_{}{}^{}F_{}^{\nu \rho }+^\nu {}_{}{}^{}F_{}^{\rho \mu }+^\rho {}_{}{}^{}F_{}^{\mu \nu })=0$$ (38) that is: $$\{\begin{array}{c}^\mu F^{\nu \rho }+^\nu F^{\rho \mu }+^\rho F^{\mu \nu }=gN^{\mu \nu \rho }\\ \\ ^\mu {}_{}{}^{}F_{}^{\nu \rho }+^\nu {}_{}{}^{}F_{}^{\rho \mu }+^\rho {}_{}{}^{}F_{}^{\mu \nu }=qN^{\mu \nu \rho }\end{array}$$ (39) where $`N^{\mu \nu \rho }`$ is a tensor to be interpreted. Using the differential forms language defined on the Minkowski space $`(M_4)`$ we write the preceding equations in a compact form: $$\{\begin{array}{c}dF=gN\\ \\ d^{}F=qN\end{array}$$ (40) where $`F`$ and $`{}_{}{}^{}F`$ $`^2(M_4)`$ and $`N`$ $`^3(M_4)`$ . If we put: $$\{\begin{array}{c}gN=^{}k\\ \\ qN=^{}j\end{array}$$ (41) where $`j`$ and $`k`$ $`^1(M_4)`$, we deduce: $$\{\begin{array}{c}\delta F=j\\ \\ dF=^{}k\end{array}$$ (42) $`\delta `$ is the usual codifferential $`\delta :^k(M_4)^{k1}(M_4)`$ defined here as: $`\delta =()^{k(4k+1)+1}(^{}d^{})`$ Interpreting the $`1`$-forms $`j`$ and $`k`$ as the electric and magnetic four dimensional current densities, we obtained the two groups of Maxwell equations in the presence of a magnetic monopole. The situation without monopole is evidently obtained by putting the $`1`$-form $`k`$ equal to zero. We easily see by means of the Poincaré theorem that: $$\delta ^2F=\delta j=0$$ (43) which is nothing else that the current density continuity equation: $$_\mu \text{ }j^\mu =m[\underset{\mu }{\overset{.}{x}},\text{ }j^\mu ]=0,$$ (44) From the skew property of the brackets, we can choose: $$j^\mu =\rho \stackrel{\mu }{\stackrel{.}{x}},$$ (45) $`\rho `$ is the charge density whose dynamic evolution is given by: $$\frac{d\rho }{d\tau }=[H,\rho ]=m[\stackrel{\mu }{\stackrel{.}{x}},\rho ]\underset{\mu }{\overset{.}{x}}=\left(^\mu \rho \right)\underset{\mu }{\overset{.}{x}}=^\mu j_\mu =0$$ (46) We see that $`H`$ automatically takes into account the gauge curvature. It plays the role of a Hamiltonian not with the usual Poisson brackets, but with new four-dimensional brackets which can be related to for example, those used by Feynman in his derivation of Maxwell equations as published by Dyson . ## V Application to a curved space In this section we extend the previous analysis to the case of a general space time metric $`g_{\mu \nu }(x)`$. In this case we define the covariant Hamiltonian from the usual fundamental quadratic form $`ds^2`$ in the following manner: $`H`$ $`=`$ $`{\displaystyle \frac{1}{2}}m\left({\displaystyle \frac{ds}{d\tau }}\right)^2={\displaystyle \frac{1}{2}}mg_{\mu \nu }(x)\stackrel{\mu }{\stackrel{.}{x}}\stackrel{\nu }{\stackrel{.}{x}}`$ In the same manner asin section $`3`$, we can prove the relation between the metric tensor and the brackets structure: $`m\text{ }[\stackrel{\nu }{\stackrel{.}{x}},x^\mu ]=g^{\mu \nu }(x)`$ The law of motion is: $`\stackrel{\mu }{\stackrel{..}{x}}`$ $`=`$ $`[H,\stackrel{\mu }{\stackrel{.}{x}}]={\displaystyle \frac{1}{2}}m[g_{\nu \rho },\stackrel{\mu }{\stackrel{.}{x}}]\stackrel{\nu }{\stackrel{.}{x}}\stackrel{\rho }{\stackrel{.}{x}}+m[\stackrel{\nu }{\stackrel{.}{x}},\stackrel{\mu }{\stackrel{.}{x}}]\underset{\nu }{\overset{.}{x}}`$ (47) $`=`$ $`{\displaystyle \frac{1}{2}}\text{ }^\mu g_{\nu \rho }\stackrel{\nu }{\stackrel{.}{x}}\stackrel{\rho }{\stackrel{.}{x}}N^{\nu \mu }\underset{\nu }{\overset{.}{x}}`$ (48) where we define $`N^{\mu \nu }(x,\stackrel{.}{x})`$ as: $$[\stackrel{\mu }{\stackrel{.}{x}},\stackrel{\nu }{\stackrel{.}{x}}]=\frac{N^{\mu \nu }(x,\stackrel{.}{x})}{m}$$ (49) Note that this tensor is now velocity-dependent, in contrast to the Minkowski case. By means of equation (23) and (48), we deduce the relation: $$\frac{N^{\mu \nu }}{\underset{\rho }{\overset{.}{x}}}=^\nu g^{\rho \mu }^\mu g^{\rho \nu }$$ (50) then: $$N^{\mu \nu }(x,\stackrel{.}{x})=\text{ }(^\mu g^{\rho \nu }^\nu g^{\rho \mu })\underset{\rho }{\overset{.}{x}}+n^{\mu \nu }(x)$$ (51) where the tensor $`n^{\mu \nu }(x)`$ is only position dependent. If we introduce this equation in (48), we find: $`\stackrel{\mu }{\stackrel{..}{x}}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\text{ }^\mu g_{\nu \rho }\stackrel{\nu }{\stackrel{.}{x}}\stackrel{\rho }{\stackrel{.}{x}}(^\mu g^{\rho \nu }^\nu g^{\rho \mu })\underset{\nu }{\overset{.}{x}}\underset{\rho }{\overset{.}{x}}+n^{\mu \nu }(x)x_\nu `$ (52) $`=`$ $`{\displaystyle \frac{1}{2}}\text{ }^\mu g^{\nu \rho }\underset{\nu }{\overset{.}{x}}\underset{\rho }{\overset{.}{x}}(^\mu g^{\rho \nu }{\displaystyle \frac{1}{2}}^\nu g^{\rho \mu }{\displaystyle \frac{1}{2}}^\rho g^{\nu \mu })\underset{\nu }{\overset{.}{x}}\underset{\rho }{\overset{.}{x}}+n^{\mu \nu }(x)\underset{\nu }{\overset{.}{x}}`$ (53) $`=`$ $`\mathrm{\Gamma }^{\nu \rho \mu }\underset{\nu }{\overset{.}{x}}\underset{\rho }{\overset{.}{x}}+n^{\mu \nu }(x)\underset{\nu }{\overset{.}{x}}`$ (54) where we have defined the Christoffel symbols by: $`\mathrm{\Gamma }^{\nu \rho \mu }`$ $`=`$ $`{\displaystyle \frac{1}{2}}\text{ }\left([\stackrel{\rho }{\stackrel{.}{x}},[\stackrel{\nu }{\stackrel{.}{x}},x^\mu ]][\stackrel{\nu }{\stackrel{.}{x}},[\stackrel{\rho }{\stackrel{.}{x}},x^\mu ]][\stackrel{\mu }{\stackrel{.}{x}},[\stackrel{\rho }{\stackrel{.}{x}},x^\nu ]]\right)`$ (55) $`=`$ $`{\displaystyle \frac{1}{2}}\text{ }(^\rho g^{\nu \mu }^\nu g^{\rho \mu }^\mu g^{\rho \nu })`$ (57) Comparing with the usual law of motion of a particle in an electromagnetic field, as in the situation of a flat space, we can put: $$F^{\mu \nu }(x)=\frac{m}{q}\text{ }n^{\mu \nu }(x)$$ (58) and get the equation of motion of a particle in a curved space: $$m\text{ }\frac{d\stackrel{\mu }{\stackrel{.}{x}}}{d\tau }=m\text{ }\mathrm{\Gamma }_{\nu \rho }^\mu \stackrel{\nu }{\stackrel{.}{x}}\stackrel{\rho }{\stackrel{.}{x}}q\text{ }F^{\nu \mu }\underset{\upsilon }{\overset{.}{x}}$$ (59) so that: $$[H,\stackrel{\mu }{\stackrel{.}{x}}]=\text{ }\mathrm{\Gamma }_{\nu \rho }^\mu \stackrel{\nu }{\stackrel{.}{x}}\stackrel{\rho }{\stackrel{.}{x}}\frac{q}{m}\text{ }F^{\nu \mu }\underset{\upsilon }{\overset{.}{x}}$$ (60) Note the difference between the two tensor $`N^{\mu \nu }`$ and $`F_{\mu \nu }`$ whose definitions are: $$\{\begin{array}{c}[\stackrel{\mu }{\stackrel{.}{x}},\stackrel{\nu }{\stackrel{.}{x}}]=\frac{N^{\mu \nu }}{m}=g^{\mu \rho }g^{\nu \sigma }\frac{N_{\rho \sigma }}{m}\\ [\underset{\mu }{\overset{.}{x}},\underset{\nu }{\overset{.}{x}}]=\frac{F_{\mu \nu }}{m}=g_{\mu \rho }g_{\nu \sigma }\frac{F^{\rho \sigma }}{m}\end{array}$$ (61) and more generally: $$\{\begin{array}{c}[\stackrel{\mu }{\stackrel{.}{x}},f(\stackrel{.}{x},\tau )]=\frac{N^{\mu \nu }}{m}\frac{f(\stackrel{.}{x},\tau )}{\stackrel{\nu }{\stackrel{.}{x}}}\\ [\underset{\mu }{\overset{.}{x}},f(\stackrel{.}{x},\tau )]=\frac{F_{\mu \nu }}{m}\frac{f(\stackrel{.}{x},\tau )}{\underset{\nu }{\overset{.}{x}}}\end{array}$$ (62) As in the case of flat Minkowski space, it is not difficult to recover the two groups of Maxwell equations with or without monopoles. In this last case we must take the following definition for the dual field: $${}_{}{}^{}F_{}^{\mu \nu }=\frac{1}{2\sqrt{g}}\epsilon ^{\mu \nu \rho \sigma }F_{\rho \sigma }.$$ (63) Now we will show that the covariant derivative and the curvature tensor can be naturally introduced with our formalism. ### A Covariant derivative As in the flat-space case, the equation of motion can be rewritten in the two following manners: $$m\text{ }\frac{d\stackrel{\mu }{\stackrel{.}{x}}}{d\tau }=m\text{ }\mathrm{\Gamma }_{\nu \rho }^\mu \stackrel{\nu }{\stackrel{.}{x}}\stackrel{\rho }{\stackrel{.}{x}}q\text{ }F^{\nu \mu }\underset{\upsilon }{\overset{.}{x}}$$ (64) and: $$m\text{ }\frac{d\stackrel{\mu }{\stackrel{.}{x}}}{d\tau }=m\text{ }\frac{\stackrel{\mu }{\stackrel{.}{x}}}{x^\nu }\stackrel{\nu }{\stackrel{.}{x}}$$ (65) we then put: $$\frac{\stackrel{\mu }{\stackrel{.}{x}}}{x^\nu }=\mathrm{\Gamma }_{\nu \rho }^\mu \stackrel{\rho }{\stackrel{.}{x}}+\frac{q}{m}F^\mu {}_{\nu }{}^{}=[H^{},\stackrel{\mu }{\stackrel{.}{x}}]$$ (66) From equation (66), a covariant derivative can be defined by means of the brackets. For an arbitrary vector we put: $$m\text{ }[\underset{\nu }{\overset{.}{x}},V^\mu (x)]=\frac{V^\mu (x)}{x^\nu }$$ (67) We then define as the usual covariant derivative: $$[D_\nu ,V^\mu ]=\frac{V^\mu }{x^\nu }+\mathrm{\Gamma }_{\nu \rho }^\mu V^\rho $$ (68) and for an arbitrary mixed tensor: $$[D_\nu ,T^\mu {}_{\sigma }{}^{}]=\frac{T^\mu _\sigma }{x^\nu }+\mathrm{\Gamma }_{\nu \rho }^\mu T^\rho {}_{\sigma }{}^{}\mathrm{\Gamma }_{\nu \sigma }^\rho T^\mu _\rho $$ (69) For the particular case of the velocity we get: $$[D_\nu ,\stackrel{\mu }{\stackrel{.}{x}}]=\frac{\stackrel{\mu }{\stackrel{.}{x}}}{x^\nu }+\mathrm{\Gamma }_{\nu \rho }^\mu \stackrel{\rho }{\stackrel{.}{x}}=\frac{q}{m}F^\mu _\nu $$ (70) and in addition we recover the standard result: $`[D_\nu ,g^{\mu \nu }]=0`$ ### B Curvature tensor From this definition of the covariant derivative we can naturally express a curvature tensor by means of the brackets. Let’s compute the following expressions: $`[D^\mu ,[D^\nu ,V^\rho ]]`$ $`=`$ $`[\stackrel{\mu }{\stackrel{.}{x}},^\nu V^\rho +\mathrm{\Gamma }_\sigma ^{\nu \rho }V^\sigma ]+\mathrm{\Gamma }_\alpha ^{\mu \nu }(^\alpha V^\rho +\mathrm{\Gamma }_\sigma ^{\alpha \rho }V^\sigma )+\mathrm{\Gamma }_\alpha ^{\mu \rho }(^\nu V^\alpha +\mathrm{\Gamma }_\sigma ^{\alpha \nu }V^\sigma )`$ (71) $`=`$ $`^\mu ^\nu V^\rho +^\mu (\mathrm{\Gamma }_\sigma ^{\nu \rho })V^\sigma +\mathrm{\Gamma }_\sigma ^{\nu \rho }(^\mu V^\sigma )+\mathrm{\Gamma }_\alpha ^{\mu \nu }(^\alpha V^\rho +\mathrm{\Gamma }_\sigma ^{\alpha \rho }V^\sigma )`$ (73) $`+\mathrm{\Gamma }_\alpha ^{\mu \rho }(^\nu V^\alpha +\mathrm{\Gamma }_\sigma ^{\alpha \nu }V^\sigma )`$ and therefore: $`[D^\mu ,[D^\nu ,V^\rho ]][D^\nu ,[D^\mu ,V^\rho ]]`$ $`=`$ $`^\mu (\mathrm{\Gamma }_\sigma ^{\nu \rho })V^\sigma ^\nu (\mathrm{\Gamma }_\sigma ^{\mu \rho })V^\sigma +\mathrm{\Gamma }_\alpha ^{\mu \rho }\mathrm{\Gamma }_\sigma ^{\alpha \nu }V^\sigma \mathrm{\Gamma }_\alpha ^{\nu \rho }\mathrm{\Gamma }_\sigma ^{\alpha \mu }V^\sigma `$ (75) $`+\mathrm{\Gamma }_\alpha ^{\nu \mu }(^\alpha V^\rho +\mathrm{\Gamma }_\sigma ^{\alpha \rho }V^\sigma )\mathrm{\Gamma }_\alpha ^{\mu \nu }(^\alpha V^\rho +\mathrm{\Gamma }_\sigma ^{\alpha \rho }V^\sigma )`$ $`=`$ $`R^{\mu \nu \rho }{}_{\sigma }{}^{}V_{}^{\sigma }+\mathrm{\Omega }^{\mu \nu }{}_{\alpha }{}^{}D_{}^{\alpha }V^\rho `$ (76) where we have introduced the torsion tensor $`\mathrm{\Omega }^{\mu \nu }_\alpha `$ $`=\mathrm{\Gamma }_\alpha ^{\nu \mu }\mathrm{\Gamma }_\alpha ^{\mu \nu }=0,`$ and the curvature tensor $`R^{\mu \nu \rho }_\sigma `$ . Due to symmetric property of the Christoffel symbols, the curvature tensor is reduced to: $$R^{\mu \nu \rho }{}_{\sigma }{}^{}V_{}^{\sigma }=^\mu (\mathrm{\Gamma }_\sigma ^{\nu \rho })V^\sigma ^\nu (\mathrm{\Gamma }_\sigma ^{\mu \rho })V^\sigma +\mathrm{\Gamma }_\alpha ^{\mu \rho }\mathrm{\Gamma }_\sigma ^{\alpha \nu }V^\sigma \mathrm{\Gamma }_\alpha ^{\nu \rho }\mathrm{\Gamma }_\sigma ^{\alpha \mu }V^\sigma $$ (77) The Jacobi identity gives: $$[D^\mu ,[D^\nu ,V^\rho ]]+[D^\nu ,[V^\rho ,D^\mu ]]+[V^\rho ,[D^\mu ,D^\nu ]]=0$$ (78) that is: $$[D^\mu ,[D^\nu ,V^\rho ]][D^\nu ,[D^\mu ,V^\rho ]]=[[D^\mu ,D^\nu ],V^\rho ]=0$$ (79) and finally: $$[[D^\mu ,D^\nu ],V^\rho ]=R^{\mu \nu \rho }{}_{\sigma }{}^{}V_{}^{\sigma }$$ (80) remark 2. We can also define the Ricci and the electromagnetic energy-impulsion tensors, but we were unable to deduce the Einstein equation from this formalism. Naturally, we can write this equation with our brackets as a constraint equation. remark 3. We can generalize the covariant derivative in including the skew-symmetric tensor $`F^\mu _\nu `$ in the definition. For this we take into account the gauge curvature for the determination of the new covariant derivative. For a vectorial function of the velocity we write: $$[\mathrm{\Delta }_\nu ,f^\mu (\stackrel{.}{x})]=\frac{f^\mu (\stackrel{.}{x})}{x^\nu }+\mathrm{\Gamma }_{\nu \rho }^\mu f^\rho (\stackrel{.}{x})\frac{q}{m}F{}_{\rho \nu }{}^{}\frac{f^\mu (\stackrel{.}{x})}{\underset{\rho }{\overset{.}{x}}}$$ (81) and then for the velocity: $$[\mathrm{\Delta }_\nu ,\stackrel{\mu }{\stackrel{.}{x}}]=\frac{\stackrel{\mu }{\stackrel{.}{x}}}{x^\nu }+\mathrm{\Gamma }_{\nu \rho }^\mu \stackrel{\rho }{\stackrel{.}{x}}\frac{q}{m}F{}_{}{}^{\mu }{}_{\nu }{}^{}=0$$ (82) The covariant derivatives, are then simultaneously covariant under both local internal and external gauges. If we want to keep a synthetic form for the formulas using the curvature and torsion tensors, we must suppose for an arbitrary vector the relation: $$[\mathrm{\Delta }_\nu ,V^\mu ]=\frac{V^\mu }{x^\nu }+\mathrm{\Gamma }_{\nu \rho }^\mu V^\rho A_\nu V^\mu $$ (83) where the vector $`A_\nu `$ is defined by the following equation: $$F^{\mu \nu }=m\left([\stackrel{\mu }{\stackrel{.}{x}},A^\nu ][\stackrel{\nu }{\stackrel{.}{x}},A]^\mu \right)$$ (84) therefore we have: $$[\mathrm{\Delta }^\mu ,[\mathrm{\Delta }^\nu ,V^\rho ]][\mathrm{\Delta }^\nu ,[\mathrm{\Delta }^\mu ,V^\rho ]]=[[\mathrm{\Delta }^\mu ,\mathrm{\Delta }^\nu ],V^\rho ]=R^{\mu \nu \rho }{}_{\sigma }{}^{}V_{}^{\sigma }+\mathrm{\Omega }^{\mu \nu }{}_{\alpha }{}^{}\mathrm{\Delta }_{}^{\alpha }V^\rho +F^{\mu \nu }V^\rho $$ (85) We define a new ”generalized” curvature tensor which matches the local electromagnetism internal symmetry with the local external symmetry: $$\stackrel{\mu \nu \rho }{\stackrel{\mathrm{\_}}{R}}{}_{\sigma }{}^{}V_{}^{\sigma }=R^{\mu \nu \rho }{}_{\sigma }{}^{}V_{}^{\sigma }+F^{\mu \nu }V^\rho $$ (86) then: $$[[\mathrm{\Delta }^\mu ,\mathrm{\Delta }^\nu ],V^\rho ]=\stackrel{\mu \nu \rho }{\stackrel{\mathrm{\_}}{R}}{}_{\sigma }{}^{}V_{}^{\sigma }$$ (87) ## VI Conclusion The goal of this work was to study the dynamic associated with the Lorentz-covariant Hamiltonian well known in analytic mechanic. For this, we introduced a four dimensional bracket structure which gives an algebraic structure between the position and velocity and generalizes the Poisson brackets. This leads us to introduce a new time parameter which is not the proper time, but is the conjugate coordinate of this covariant Hamiltonian. This formal construction allows to recover the two groups of Maxwell equations in flat space. This approach is close to the one used by Feynman in his own derivation of the first group of Maxwell equations. The principal interest of this method, besides the phase space formalism, is in the study of theories with gauges symmetries because it avoids the introduction of the non-gauge invariant momentum. Our formalism can be directly extrapolated to the curved space, where the principal notions are introduced in a natural manner. A five-dimensional structure can also be studied by considering the $`\tau `$ parameter as a fifth coordinate. In such a case equations take a simpler form, particularly the group of Maxwell equations, but the meaning of this new coordinate is still difficult to interpret, and could be perhaps understood in the context of Kaluza-Klein compactification. Just after finishing this work we received a paper referring to the covariant Hamiltonian in the context of Feynman’s proof of the Maxwell equations . Acknowledgment: We would like to thank Y.Grandati for helpful discussions.
warning/0004/cond-mat0004415.html
ar5iv
text
# Mobile Bipolarons in the Adiabatic Holstein-Hubbard Model in 1 and 2 dimensions. ## 1 Introduction It is well-known for many decades since that a single electron submitted to an electron phonon coupling may generate a polaron that is a quasi-particle consisting into an electron localized in a self consistent lattice potential. For a local electron phonon interaction, the formation of a polaron always occurs in 1D. At small coupling, this polaron becomes large and highly mobile. On the opposite, in 2D (and more dimensions), the formation of a polaron requires a large enough coupling . Then, the polarons are always small that is mostly localized on a single site and not mobile. If there is no electron repulsion, two polarons form a bipolaron with two electrons in a spin singlet state localized in the same potential well. When there is a strong enough electronic repulsion, the bipolaron may be broken into two unbound polarons. However, in the intermediate regime, we are going to show that there are new bounded bipolaronic states with interesting properties. There have been many tentative theories about bipolaronic superconductivity for many years (see for a review). In the well-known BCS theory valid at weak electron phonon coupling, the electron pairs (Cooper pairs) form self consistently in the superconducting phase only and are spatially very extended. For the bipolaronic superconductivity, it is speculated that when there are pre-existent bounded pairs of electrons (small bipolaron), that they may behave as a quantum boson liquid and condense into a superfluid state. However, the most serious critiscismes to these theories is that when the bipolarons exist in 2 and 3 D models (with realistic physical parameters), their effective mass is so huge that their quantum character completely disappear. They should condense into spatially ordered or disordered structures which could be just considered as a chemical bond ordering. Expansions at large electron phonon coupling , modelize the bipolaronic system as an array of coupled quantum spins $`1/2`$ with two types of coupling. The spin is $`+1/2`$ at a site occupied by a bipolaron and $`1/2`$ in the opposite case. There is an $`xy`$ coupling between neighboring spins describing the bipolaron tunnelling and which yields the bipolaron band width and a $`zz`$ coupling representing the potential interaction between the bipolarons. If the $`xy`$ term dominates, the spin ordering occurs in the plane $`xy`$ and the structure is superconducting. If the $`z`$ term dominates, one get a spatial ordering of the pseudospin in the $`z`$ direction that is a spatial ordering of the bipolarons (or chemical bonds) in the real space. The real systems where the effective mass of the electrons are usually much smaller than the atomic masses are generally close to the adiabatic limit. In that case the $`xy`$ couplings turns out to be negligible compared to the $`z`$ interaction which eliminates any possibility of superconductivity but instead of, a spatial bipolaron ordering. It is clear that this regime is quite well-described within a purely adiabatic approximation (the chemist approximation). However, it might exist special situations where this $`xy`$ coupling could be enhanced to relative large values and dominates the $`z`$ coupling thus favoring a bipolaronic superconductivity at unusually high temperatures. The purpose of this paper is to show that in a restricted region of the parameter space only, the competition between the electron phonon coupling and the electron-electron repulsion could produce a sharp increase in the mobility of bipolarons while simultaneously the binding energy of the bipolaron remains significantly large thus favoring superconductivity as conjectured in . We test these ideas on the simplest model which is the Holstein Hubbard model, where both electron-phonon interactions and electron-electron interaction are involved. In a first step, we consider the adiabatic model where the quantum fluctuations of the lattice are neglected and shall look at the bipolaron classical mobility only. The next step will be done in a forthcoming work, where it will be shown that when the bipolarons are very mobile, the adiabatic approximation is not valid because they are very sensitive to the quantum lattice fluctuation by producing a sharp increase of the bipolaron bandwidth. ## 2 The Model: definitions In order to make the physical parameters explicit, we write the Holstein-Hubbard Hamiltonian with its full set of parameters: $$=T\underset{<i,j>,\sigma }{}C_{i,\sigma }^+C_{j,\sigma }+\underset{i}{}\mathrm{}\omega _0(a_i^+a_i)+gn_i(a_i^++a_i)+\upsilon n_{i,}n_{i,}$$ (1) where $`T`$ is the transfer integral between nearest neighbor sites $`<i,j>`$ of the lattice, of the electrons represented by the standard fermion operators $`C_{i,\sigma }^+`$ and $`C_{j,\sigma }`$ at site $`i`$ with spin $`\sigma =`$ or $``$. $`a_i^+`$ and $`a_i`$ are standard creation and annihilation boson operators of phonons. $`\mathrm{}\omega _0`$ is the phonon energy of a dispersionless optical phonon branch. $`g`$ is the constant of the onsite electron phonon coupling. The onsite electron-electron interaction is represented by a Hubbard term with positive coupling $`\upsilon `$. Choosing the energy $`E_0=8g^2/\mathrm{}\omega _0`$ as energy unit and introducing the position and momentum operators: $`u_i`$ $`={\displaystyle \frac{\mathrm{}\omega _0}{4g}}(a_i^++a_i)`$ (2) $`p_i`$ $`=i{\displaystyle \frac{2g}{\mathrm{}\omega _0}}(a_i^+a_i)`$ (3) we obtain the dimensionless hamiltonian: $$H=\underset{i}{}\left(\frac{1}{2}u_i^2+\frac{1}{2}u_in_i+Un_in_i\right)\frac{t}{2}\underset{<i,j>,\sigma }{}C_{i,\sigma }^+C_{j,\sigma }+\frac{\gamma }{2}\underset{i}{}p_i^2$$ (4) The parameters of the system are now: $$E_0=8g^2/\mathrm{}\omega _0U=\frac{\upsilon }{E_0}t=\frac{T}{E_0}\gamma =\frac{1}{4}(\frac{\mathrm{}\omega _0}{2g})^4$$ (5) As soon as the electron phonon coupling $`g`$ becomes reasonably large that is larger than the phonon energy $`\mathrm{}\omega _0`$, $`\gamma `$ becomes very small. The adiabatic approximation is obtained by taking $`\gamma =0`$. We shall assume this condition from nowon. Then $`\{u_i\}`$ commutes with the Hamiltonian and can be taken as a scalar variable. For a given set of $`\{u_i\}`$, the 2-electron ground-state of the Hamiltonian $$H_{ad}=\underset{i}{}\left(\frac{1}{2}u_i^2+\frac{1}{2}u_in_i+Un_in_i\right)\frac{t}{2}\underset{<i,j>,\sigma }{}C_{i,\sigma }^+C_{j,\sigma }$$ (6) has to be searched among singlet states with the form $$|\mathrm{\Psi }>=\underset{i,j}{}\psi _{i,j}C_{i,}^+C_{j,}^+|\mathrm{}>$$ (7) where $`\psi _{i,j}=\psi _{j,i}`$ is normalized $`_{i,j}|\psi _{i,j}|^2=1`$. Then the energy of the system depends on $`|\psi >=\{\psi _{i,j}\}`$ and $`\{u_i\}`$ as $$F(\{\psi _{i,j}\},\{u_i\})=\underset{i}{}\left(\frac{1}{2}u_i^2+u_i\rho _i+U|\psi _{i,i}|^2\right)\frac{t}{2}<\psi |\mathrm{\Delta }|\psi >$$ (8) where $$\rho _i=\frac{1}{2}\underset{j}{}(|\psi _{i,j}|^2+|\psi _{j,i}|^2)$$ (9) is the half electronic density at site $`i`$ and $`\mathrm{\Delta }`$ is a discrete Laplacian operator in $`2d`$ dimensions ($`d`$ being the initial lattice dimension). ## 3 The anticontinuous limit The minimization of $`F(\{\psi _{i,j}\},\{u_i\})`$ with respect to the normalized electronic state $`\{\psi _i^{}\}`$ yields the electronic eigenequation $$\frac{t}{2}\mathrm{\Delta }\psi _{i,j}+\left(\frac{1}{2}(u_i+u_j)+U\delta _{i,j}\right)\psi _{i,j}=F_{el}(\{u_i\})\psi _{i,j}$$ (10) where $`F_{el}`$ is the electronic ground-state energy in the potential generated by $`\{u_i\}`$. Then for $`t>0`$, the wave function $`\psi _{i,j}`$ has to be real, positive and symmetric. The adiabatic potential for the atoms is then $$F_{ad}(\{u_i\})=\underset{i}{}\frac{1}{2}u_i^2+F_{el}(\{u_i\})$$ (11) It has infinitely many minima close to the anticontinuous limit $`t=0`$. An adiabatic configuration $`\{u_i\}`$ is metastable when it is a local minima of 11 . The minimization of $`F(\{\psi _{i,j}\},\{u_i\})`$ can also be done first with respect to $`\{u_i\}`$ which yields $`u_i+\rho _i=0`$. Substitution in 8, yields a different variational form on $`\{\psi _{i,j}\}`$ non equivalent to 11 $$F_\psi (\{\psi _{i,j}\},\{u_i\})=\underset{i}{}\left(\frac{1}{2}\rho _i^2+U\underset{i}{}|\psi _{i,i}|^2\right)\frac{t}{2}<\psi |\mathrm{\Delta }|\psi >$$ (12) The extremalization of 12 with respect to $`\psi _{i,j}`$ with the condition of normalization, yields a generalized discrete nonlinear Schroedinger equation on a $`2d`$-dimensional lattice $$\frac{t}{2}\mathrm{\Delta }\psi _{i,j}+\left(\frac{1}{2}(\rho _i+\rho _j)+U\delta _{i,j}\right)\psi _{i,j}=F_{el}\psi _{i,j}$$ (13) corresponding to 10. This equation 13 has also an anticontinuous limit at $`t=0`$ which is different of those of model 11. It yields more states at this limit because it is not required that the electronic eigen states be in its ground-state with respect to the lattice potential. This allows one to involve electronic excitations but it has of course the same ground-states. The classification of these states will not be discussed here. However for eq.13 at $`t=0`$, the phases of each complex number $`\psi _{i,j}`$ is arbitrary. This situation is analogous to the anticontinuous limit of the breather problem where the phase of the uncoupled oscillators is degenerate (see ref. for details). The consequence is that the implicit function theorem cannot be applied directly for proving the possible continuation of the solutions at $`t=0`$. For the breather problem, a trick has been to consider first only time reversible solutions thus removing this phase degeneracy but this was not absolutely necessary ). In our model, it is also convenient to consider in a first step only real electronic wave functions. Then, these real normalized solutions at $`t=0`$ can be continued for $`t0`$ not too large. This is not a restriction for finding the electronic ground-states because we know that the corresponding electronic state is real, positive and symmetric. This continuation can be done by numerical methods. Most of the real solutions of 13 are unstable for $`t`$ small but some of them can recover stability at larger $`t`$ and even become the ground-state. We calculated numerically with a high accuracy a few number of these solutions rather well localized on a small number of sites (also taking advantage of its spatial symmetries if any) which we believed to be possible candidate for being a ground-state. By comparing their energies one with each others, the bipolarons state are found which are presumably the exact ground-state. In the domain of parameter we explored, we found mostly three kinds of bipolaronic states with a significantly large binding energy (see fig.1) as ground-states in some domain of $`U`$ and $`t`$. <sup>1</sup><sup>1</sup>1However, there are unexplored domains for large $`U`$ where the ground-states will be different but their very weak binding energy makes that they are less interesting at the present stage of our approach. They are obtained by continuation from the following solutions of eq. 13 at $`t=0`$: 1-the standard onsite bipolaron denoted (S0) at given site $`i`$ where $$\psi _{i,i}=1\text{and}\psi _{m,n}=0\text{else for}(m,n)(i,i)$$ (14) 2- the two-site singlet bipolaron , localized on two nearest neighbor sites $`i`$ and $`j`$ denoted (S1) which was named ”Spin Resonant” bipolaron in ref.. We also observed ground-states which are singlet states where the polarons are at distance $`2`$ (denoted (S2) but with a weaker binding energy. $`\psi _{i,j}`$ $`=`$ $`\psi _{j,i}={\displaystyle \frac{1}{\sqrt{2}}}\text{and}`$ $`\psi _{m,n}`$ $`=`$ $`0\text{else for}(m,n)(i,j)\text{and}(m,n)(j,i)`$ (15) 3-and in two dimensions, a new unexpected solution which is the quadrisinglet localized bipolaron denoted (QS) which is localized at the given site $`i`$ and its four nearest neighbors $`j_\nu `$ ($`\nu =1,..,4`$). It is the linear combination of four singlets located on the four nearest neighbor bonds $`<i,j_\nu >`$ $`\psi _{i,j_\nu }`$ $`=`$ $`\psi _{j_\nu ,i}={\displaystyle \frac{1}{\sqrt{8}}}\text{and}`$ $`\psi _{m,n}`$ $`=`$ $`0\text{else for}(m,n)(i,j_\nu )\text{and}(m,n)(j_\nu ,i)`$ (16) This solution can be viewed as a localized RVB state similar to those proposed by Anderson some years ago in the pure Hubbard model in 2D as a theory for superconductivity in cuprates. The physical origin of the binding of the bipolaron in states (S1), (S2).. and (QS) can be interpreted of magnetic origin. When the Hubbard term increases too much, bipolaron (S0) could break into two polarons far apart but then both of them are magnetic with a spin $`1/2`$. They should interact by an antiferromagnetic exchange coupling, as predicted by standard perturbation theories. Then, for moderately large Hubbard terms, it remains more favorable to reduce the distance between the two polarons for gaining a magnetic energy by forming a singlet state (S1), (S2)…. We have no simple interpretation to explain why bipolaron (QS) which has a more complex structure and could become the most stable in 2D models but is not in 1D (as far in our domains of investigation). ## 4 Bipolaronic Ground-states in One dimension The size of the system was chosen large enough compared to the bipolaron size (in our calculations up to the size 41 for spatially symmetric states). As expected, we found that for any value of $`t`$ (tested up to $`t=0.6`$), there is always a biplaronic state which is lower in energy than a pair of extended electrons. The phase diagram is shown fig.2. The first order transition lines are sharp and well defined for $`t`$ small when the bipolarons have a small spatial extension. When $`t`$ increases, the size of the bipolarons also increases and it becomes practically impossible to distinguish numerically the energy differences between bipolaronic states obtained by continuation from different states at the anticontinuous limit. This can be interpreted by the fact that when $`t`$ is large, such a model is well described with a continuous space variable $`x`$ instead of the discrete lattice site $`i`$. This continuous model exhibit only one bipolaronic solution which ”erases” all the transitions due to discreteness. We now look at the possible mobility of the obtained bipolarons. The precise calculation of the Peierls-Nabarow(PN) energy barrier requires too much numerical work while the calculation of the pinning modes of the bipolarons is much easier and brings nevertheless a similar information. The eigenvalues of the matrix of the second variation of the lattice potential energy $`F_{ad}(\{u_i\})`$ 11 are the squares of the phonon frequencies of the bipolaron calculated within the standard Born-Oppenheimer approximation (in unit $`\sqrt{\gamma }`$). Beside the initial flat optical branch of phonon at frequency $`1`$, there are two localized mode. One is spatially antisymmetric and correspond to the pinning mode of the bipolaron. The second one is spatially symmetric and is a breathing mode. There is a phonon softening at the first order transition between bipolaron (S0) and (S1) which becomes almost complete for $`t`$ large enough. Fig.3 shows for $`t=0.3`$, that the frequency of the bipolaron (S0) practically vanishes at the transition with (S1) which becomes almost second order. A strong softening of the pinning mode of the bipolaron is usually associated with a high mobility. This can be tested by integrating the classical dynamical equations of the lattice with the effective potential $`F_{ad}(\{u_i\})`$ taking as initial condition the bipolaron configuration with a small perturbation in the direction of the pinning mode (see a similar calculation for a moving breather in ref.). The moving bipolaron shown fig.4 corresponds to the dip of the pinning mode of fig.3. Although it is often believed that a high mobility for a localized object in discrete lattice can be achieved only when the size of this object is large compared to the lattice spacing, this moving bipolaron shown is localized on few sites only and also still strongly bounded (see inserts of fig.3). It is nevertheless highly mobile. During the motion, it can be viewed as exchanging from state (S0) to (S1) and vice-versa. Relatively small perturbation of the model parameters suffices to raise the frequency of the pinning mode and then, it can be checked that the quality of the bipolaron mobility diminishes and progresively disappears. ## 5 Bipolaronic Ground-states in Two dimensions In the 2D model, a smaller size (up to $`14\times 14`$ ) for the lattice turns out to be sufficient for accurate calculations because the polaronic states remains quite localized while they exist. The obtained phase diagram is shown fig.5. For $`t`$ large enough, the ground-state corresponds to extended electrons. For $`t`$ small, there is a first order transition as in the 1D model, between bipolarons (S0) and (S1) but also in addition, there is a small domain where the bipolaron (QS) which was initially unstable for $`t`$ small, recover its stability and even becomes the ground-state. The profiles of the three types of coexisting bipolarons (S0), (S1) and (QS) with a rather small spatial extension, are shown fig.6. Insert fig. 7 shows the energies of the bipolarons on several lines at constant $`t`$. Although, it becomes relatively small, the bipolaron binding energies remains non negligible in the vicinity of the triple point. As in 1D, the PN energy barrier is likely depressed in the vicinity of the first order transition line between (S0) and (S1) or (QS) and (S1) but difficult to calculate accurately. By contrast, the pinning and breathing frequencies of the bipolarons are much easier to calculate and shown fig.7 as a function of $`U`$ for several values of $`t`$. For (S0) and (QS) there are two degenerate pinning modes, one corresponding to the $`x`$ direction and the other one to the $`y`$ direction. Fig.7 shows that there is significant softening of both the pinning modes and the breathing mode essentially in the vicinity of the triple point. However, this softening is not sufficient for allowing the bipolaron mobility in that region (as confirmed by our tests). One may consider that a model involving onsites couplings does not favor the bipolaron mobility. Further studies on modified Holstein-Hubbard models and particularly models chosen in order to approach a better description of the cuprates layers $`Cu0_2`$, could perhaps produce classically mobile bipolarons in $`2D`$. ## 6 Discussion and Concluding Remarks Our results are not restricted to the original Holstein-Hubbard model. We already found that the three types of bipolarons (S0), (S1) and (QS) persist in appropriate domain of parameters for modified Holstein-Hubbard models for example when introducing nearest neighbour Hubbard interaction. We believe that few changes in the model which could be physically realistic, could favor the classical mobility. For example, a phonon dispersion chosen with an appropriate sign, should increases moderately the spatial extension of the bipolarons and thus favor its classical mobility. In any case, although the classical mobility of a bipolaron should favor its quantum mobility, this condition is not necessary. When the quantum lattice fluctuation are taken into account ( $`\gamma 0`$ in eq.4, the quantum correction to the bipolaronic solutions should involve at the lowest order a RPA term (corresponding to small local quantum fluctuations) and the most important term concerning its physical consequences, a tunnelling term raising the spatial degeneracy of the bipolarons. The bipolarons should thus form bands. Approximate methods for calculating these bands width analogous to those developped in ref. can be used. In the vicinity of the triple point of the phase diagram 6, the three types of bipolarons (S0), (S1), (QS) have the same energies and are resonant. There are four hybridized bands associated with a sharp increase of each bandwidth with relatively small gaps for moderately small $`\gamma `$ which makes globally a broad band while elsewhere the quantum mobility of the bipolarons sharply diminish. This current work shall be described in a forthcoming publication . Although we only considered two electrons in the lattice, we could expect reasonably that a high bipolaron mobility could persist at finite density provided it remains sufficiently far from the half filled band. Approaching this limit, the hard core interaction should prevent this mobility and produce a frozen structure of polaron (likely a SDW or an antiferromagnet). Before ending this paper, let us briefly check the physical relevance of our results. These preliminary studies indicates that although relatively weak, the binding energies of the bipolarons (for producing two extended polarons) in the vicinity of the triple point are about $`0.005`$ with our energy unit $`E_0`$. This chemical energy has to be measured in eV as well as the Hubbard term (which is about 8eV for the ion $`Cu^{++}`$). Thus it is not unreasonable that this binding energy range within several $`10^1`$ eV which is comparable to the critical temperature energies of the cuprates. In summary, we found that in the 1D model when $`U`$ increases from zero, the bipolarons are pairs of polarons at distance $`0`$ (onsite bipolaron (S0), at distance $`1`$ (two-sites bipolaron or Spin Resonant Bipolaron) (and further but with a binding energy going to zero very fast) and that there is first order transitions between these different configurations. In the 2D model, we find again the onsite bipolarons (S0) and the 2-site bipolaron (S1) but a new kind of bipolaron called quadrisinglet. There is a triple point in the phase diagram where the three kinds of bipolarons have the same energy. There is a significant phonon softening at the first order transition. In 1D, this softening can become almost complete at moderately large values of the transfer integral. Then, the classical bipolaron mobility has been effectively observed by direct simulation. By contrast, in 2D, the phonon softening has not been found sufficient to produce the classical mobility, but nevertheless one can expect a sharp effect of the bipolaron resonance at the triple point, when the quantum lattice fluctuations are taken into account. We also bring more arguments which maintain our early conjectures that the possible origin of high $`T_c`$ superconductors originate from an exceptional combination of circonstances consisting in a well-balanced competition between strong repulsive electron-electron interactions and a strong electron-phonon interactions.
warning/0004/quant-ph0004010.html
ar5iv
text
# Untitled Document Quantum computation with abelian anyons Seth Lloyd Department of Mechanical Engineering MIT 3-160, Cambridge, Mass. 02139 slloyd@mit.edu Abstract: A universal quantum computer can be constructed using abelian anyons. Two qubit quantum logic gates such as controlled-NOT operations are performed using topological effects. Single-anyon operations such as hopping from site to site on a lattice suffice to perform all quantum logic operations. Quantum computation using abelian anyons shares some but not all of the robustness of quantum computation using non-abelian anyons. A wide variety of methods can be used to construct quantum computers in principle (1-18). Essentially any interaction between two quantum degrees of freedom suffices to construct universal quantum logic gates (10-11). In particular, Kitaev (12) has shown that quantum computation can be effected using non-abelian anyons. The resulting quantum computation is intrinsically fault-tolerant. This paper shows that universal quantum computation can be effected using abelian anyons: two-qubit quantum logic gates, such as the controlled-NOT gate, are enacted topologically via the phase factor of $`e^{i\varphi }`$ that occurs when one anyon is moved around another. Quantum computation using abelian anyons shares some but not all of the robustness of quantum computation using non-abelian anyons. Possible realizations in terms of solid-state systems and interferometers are discussed, and issues of noise and decoherence are investigated. A quantum logic gate is an operation that transforms quantum-information bearing degrees of freedom. A set of quantum logic gates is universal if arbitrary quantum computations can be built up by repeatedly applying gates from the set to different qubits. To prove that abelian anyons are capable of universal quantum computation, we will show that single anyon quantum logic gates such as those that move an anyon from one site to another form a universal set: universal quantum computation can be effected simply by moving fermions around a lattice or network. Consider the case of anyons moving on a two-dimensional lattice. Each site $`j`$ of the lattice corresponds to a local mode that can be either occupied by an anyon, $`|+_j`$, or unoccupied $`|_j`$. For example, the anyons could be quantum-Hall effect excitations on a two-dimensional spatial lattice. Now consider operations that can be performed on anyons. Let $`b_j,b_j^{}`$ be the annihilation and creation operators for the $`j`$th mode. First, applying the Hamiltonian $`A_j=b_j^{}b_j`$ leaves $`|_j`$ unchanged and multiplies the state $`|+_j`$ by a phase. Second, if $`j`$ and $`k`$ are two adjacent sites or modes, applying the Hamiltonian $`B_{jk}=b_j^{}b_k+b_k^{}b_j`$ ‘swaps’ the states of the two modes. Clearly, the anyons on the lattice can be moved around at will by repeated swapping operations. Finally, when one anyon is moved around another anyon, its state acquires a phase of $`e^{i\varphi }`$. As will now be shown, this is all that is required to effect universal quantum computation. The trick to performing universal quantum computation using abelian anyons is to store a quantum bit on a single anyon at two sites. Associate two sites $`j,j^{}`$ with the $`j`$th qubit, and define the $`j`$th qubit by $`|0_j=|+_{jj^{}}`$ and $`|1_j=|+_{jj^{}}`$. The operations described above then map in a straightforward way onto the usual quantum logic operations on these qubits. The Hamiltonian $`A_j=|1_j1|=(\sigma _z^j+1)/2`$ corresponds to a rotation about the $`z`$-axis, where $`\sigma _z=|00|+|11|`$ Swapping $`j`$ and $`j^{}`$ then corresponds to a NOT operation, and a partial swap corresponds to a rotation $`e^{i\theta \sigma _x/2}`$, where $`\sigma _x=B_{jj^{}}=|01|+|10|`$. Since any single-qubit rotation can be built up out of rotations about $`x`$ and $`z`$ axes, the ability to apply $`A_j`$ and $`B_{jk}`$ translates into the ability to apply arbitrary single-qubit rotations. Note that although $`B_{jk}`$ operates on two modes or sites, it involves no direct interactions between anyons, as each of our two-site qubits contains exactly one anyons. To perform a two-qubit operation on two of our two-site qubits $`|x_j`$, $`|y_k`$, simply take whatever is in the first site of the $`j`$th qubit, and by repeated swaps, move it around the first site of the $`k`$th qubit. A convenient way to visualize such an operation is to think of time as a third dimension, so that moving the contents of one site around another is a braiding action on the time-lines of the sites. The exact path taken does not matter as long as it goes around no other qubit sites that might contain a anyon: the site is braided around one and only one other site. The overall state of the two qubits then acquires a phase of $`1`$ if and only if the first site of the $`j`$th qubit and the first site of the $`k`$th qubit originally contain anyons. Otherwise, no anyon is moved around another, and the state remains unchanged. That is, we have $$\begin{array}{cc}& |00|00\hfill \\ & |01|01\hfill \\ & |10|10\hfill \\ & |11e^{i\varphi }|11.\hfill \end{array}$$ $`(1)`$ But this is just a controlled-phase gate, closely related to a so-called controlled-NOT gate (indeed, for $`\varphi =\pi `$ as in the case of semions, the controlled phase gate can be turned into a controlled-NOT gate by application of $`\sigma _x`$ rotations to the second qubit). Single qubit rotations and controlled phase gates together form a universal set of quantum logic gates. This proves our basic result: a set of universal quantum logic gates can be constructed using abelian anyons. Note that all qubits are stored on single anyons, which can be kept an arbitrary distance from each other during the course of the quantum computation. An arbitrary quantum computation can be enacted as follows. First, map out a quantum circuit diagram for the computation in terms of elementary quantum logic gates. Program an array of two-site qubits with the proper initial states by moving an anyon to the proper site of each qubit. Enact one-qubit gates by phase shifts and partial swaps on the two sites corresponding to the qubit, and enact two-qubit operations by ‘braiding’ the contents of the first site of the first qubit about the first site of the second qubit. Read out the answer by determining the location of the fermions in the two-site output qubits. How might one realize such a quantum computer? Clearly, the above discussion suggests that a two-dimensional lattice of abelian anyons. The anyons of the quantum Hall effect will do ($`\varphi =2\pi /3`$). All that is required is the ability to perform accurate phase-shifts and swaps. These operations are local and act on single anyons. They could be enacted by applying localized potentials via, e.g., nanofabricated electrodes or scanning tunneling microscopes. Two-qubit operations are topological in nature, and hence are robust to local, anyon-number preserving errors, just as in anyonic quantum computation (12-13). Single qubit operations are not topological in nature and are less robust. If the anyons are massive and the modes in the lattice are spatially separated, then they will be subject to decoherence due to the environment effectively ‘detecting’ whether or not there is an in a particular site (20-21). If the anyons have an additional degree of freedom such as a magnetic moment, then a conceptually elegant, though technically difficult way of performing this type of anyonic quantum computation is to use interferometry in two dimensions. Here, rather than storing a qubit on two anyons, one can store it on the spin of a single anyon just as in NMR. Single qubit quantum logic operations can then be enacted by applying magnetic fields as in NMR quantum computation. The topological two-qubit gate can be enacted by applying a magnetic field gradient, as in a Stern-Gerlach apparatus. The gradient field diverts the $`j`$th fermion into one of two different modes, depending on whether or its spin is $`|1/2`$ or $`|+1/2`$. To perform the topological two-qubit phase shift gate, apply gradients to two qubits $`j`$ and $`k`$, braid the first mode of the $`j`$th qubit about the first mode of the $`k`$th qubit, then recombine the two modes of the $`j`$th qubit and the two modes of the $`k`$th qubit using gradient fields. That is, one creates two ‘braided’ Stern-Gerlach apparatuses by linking two of the four arms. The resulting ‘braided’ Stern-Gerlach apparatus performs the controlled phase shift. Accordingly, an anyonic quantum computer can in principle be constructed using interferometry alone. (In contrast, when one attempts to construct purely interferometric ‘bosonic’ quantum computers using photons, quantum computation can only be performed by using exponentially more resources than a conventional quantum computers (22-23).) Quantum computation with abelian anyons is likely to be quite difficult to accomplish in practice. Abelian anyons, however, though exotic, are less exotic than the non-abelian anyons of Kitaev. The reason that non-abelian effects are not required in this case is that the method of constructing quantum logic gates is not entirely topological: one-qubit gates are performed using ‘conventional’ quantum logic operations such as hopping. Only the two-qubit gates are topological in nature. As a result, although abelian anyons represent an experimentally more accessible path to anyonic quantum computation than that provided by non-abelian anyons, the resulting quantum computation is less fault-tolerant. It is to be hoped, however, that the universal quantum computation by abelian anyons will open up new avenues to understanding topological effects in quantum computation (16-17). Acknowledgements: The author would like to thank Michael Freedman, Alexei Kitaev, Eddie Farhi, and Jeffrey Goldstone for helpful discussions and David Divincenzo for pointing out the fatal flaw in the earlier version of this work. This work was supported by DARPA and by ARO. References (1) S. Lloyd, Science 261, 1569-1571, 1993. (2) D.P. DiVincenzo, Science 270, 255 (1995). (3) Q.A. Turchette, C.J. Hood, W. Lange, H. Mabuchi, H.J. Kimble, Phys. Rev. Lett. 75, 4710, (1995). (4) C. Monroe, D.M. Meekhof, B.E. King, W.M. Itano, D.J. Wineland, Phys. Rev. Lett. 75, 4714, (1995). (5) C.H. Bennett, Physics Today 48, 24-30 (1995). (6) D.G. Cory, A.F. Fahmy, T.F. Havel, in PhysComp96, Proceedings of the Fourth Workshop on Physics and Computation, T. Toffoli, M. Biafore, J. Leão, eds., New England Complex Systems Institute, 1996, pp. 87-91; Proc. Nat. Acad. Sci. 94, 1634 (1997); Physica D 120, 82 (1998). (8) N.A. Gershenfeld and I.L. Chuang, Science 275, 350-356 (1997). (9) J.E. Mooij, T.P. Orlando, L. Levitov, Lin Tian, Caspar H. van der Wal, and S. Lloyd, Science 285, 1036-1039 (1999). (10) S. Lloyd, Phys. Rev. Lett. 75, 346-349, 1995. (11) Deutsch, D., Barenco, A., Ekert, A., Proc. Roy. Soc. A 449, 669-677 (1995). (12) A. Yu. Kitaev, ”Fault-Tolerant Quantum Computation by Anyons” (1997), e-print quant-ph/9707021. (13) J. Preskill, ”Quantum Information and Physics: Some Future Directions” (1999), e-print quant-ph/9904022. (14) S. Lloyd, “Unconventional Quantum Computing Devices,” in Unconventional Models of Computation, C.S. Calude, J. Casti, M.J. Dinneen, eds., Springer, Singapore, 1998. (15) S.B. Bravyi, A.Yu. Kitaev, “Fermionic quantum computation,” quant-ph/0003137. (16) M.H. Freedman, A. Kitaev, Z. Wang, “Simulation of topological field theories by quantum computers,” quant-ph/0001071. (17) M.H. Freedman, M. Larsen, Z. Wang, “A modular functor which is universal for quantum computation,” quant-ph/0001108. (18) M.H. Freedman, “Quantum computation and the localization of Modular Functors,” quant-ph/0003128. (19) F. Wilczek, Fractional Statistics and Anyon Superconductivity, World Scientific, Singapore, (1990). (20) R. Landauer, Int. J. Theor. Phys. 21, 283 (1982); Found. Phys. 16, 551 (1986); Nature 335, 779 (1988); Nanostructure Physics and Fabrication. M.A. Reed and W.P. Kirk, eds. (Academic Press, Boston, 1989), pp. 17-29; Physics Today 42, 119 (October 1989); Proc. 3rd Int. Symp. Foundations of Quantum Mechanics, Tokyo, 407 (1989); Physica A 168, 75 (1990); Physics Today, 23 (May 1991); Proc. Workshop on Physics of Computation II, D. Matzke ed., 1 (IEEE Press, 1992). (21) W.H. Zurek, Physics Today 44, 1991. (22) N.J. Cerf, C. Adami, and P.G. Kwiat, Phys. Rev. A 57, 1477 (1998). (23) Lloyd, S., and Braunstein, S., Physical Review Letters, 82, 1784-1787, 1999.
warning/0004/math0004148.html
ar5iv
text
# Lagrangian and Hamiltonian Formalism for Constrained Variational Problems ## 1. Introduction The aim of this paper is to generalize to the context of constrained variational problems some classical results about the correspondence between Lagrangian and Hamiltonian formalisms (see for instance ). Particular cases of this theory are the sub-Riemannian geodesic problem (see for instance ), and the so called Vakonomic approach to the non holonomic mechanics (see for instance ). The constrained variational problem studied is modelled by the following setup: we consider an $`n`$-dimensional differentiable manifold $`M`$ endowed with a smooth distribution $`𝒟TM`$ of rank $`k`$; moreover, we assume that it is given a (possibly time-dependent) Lagrangian function $`L`$ on $`𝒟`$. In the non holonomic mechanics, $`M`$ represents the configuration space, $`𝒟`$ the constraint, and $`L`$ is typically the difference between the kinetic and a potential energy. In the sub-Riemannian geodesic problem, $`L`$ is simply the quadratic form corresponding to a positive definite metric on $`𝒟`$. The solutions of the constrained variational problem are given by curves $`\gamma :[a,b]M`$ that are critical points of the action functional $`(\gamma )=_a^bL(t,\gamma (t),\dot{\gamma }(t))dt`$ defined on the space: $$\mathrm{\Omega }_{PQ}([a,b],M,𝒟)=\{\gamma :[a,b]\stackrel{C^1}{}M:\gamma (a)P,\gamma (b)Q,\gamma ^{}(t)𝒟\text{for all}t\}$$ of horizontal curves of class $`C^1`$ in $`M`$ connecting two fixed submanifolds $`P,QM`$. It is well-known that the set $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ is in general not a submanifold of the Banach manifold of $`C^1`$ curves $`\gamma :[a,b]M`$; when $`P`$ and $`Q`$ are points, the singularities of $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ are known in the context of sub-Riemannian geometry as abnormals extremals (see ). Such singularities can be nicely described using the canonical symplectic structure of the cotangent bundle $`TM^{}`$ (see Corollary 3.7). In this paper we are interested in studying the action functional $``$ in the regular part of $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$. We remark that in several important cases the set $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ contains no singular curves (see, for instance, Corollary 3.8 and Remark 3.9). Recall from that when a Lagrangian function $`L:TMIR`$ is hyper-regular then the critical points of the corresponding (unconstrained) variational problem are given by the solutions of the Hamilton equations corresponding to a Hamiltonian $`H:TM^{}IR`$ which corresponds to $`L`$ by means of the Legendre transform. The Legendre transform described in can be generalized in a straightforward way to general vector bundles; namely, if $`L:\xi IR`$ is a smooth map on a vector bundle $`\xi `$ which is hyper-regular (in a suitable sense) then one can naturally associate to it a smooth map $`H:\xi ^{}IR`$ on the dual bundle $`\xi ^{}`$. At such level of generality, the Legendre transform does not seem to have a meaningful interpretation in the context of calculus of variations, as it does in the case $`\xi =TM`$. Our goal is to show that when $`\xi =𝒟`$ is a vector subbundle of a tangent bundle $`TM`$ (i.e., a distribution on $`M`$) then the Legendre transform for smooth maps on $`𝒟`$ has a nice application to the study of constrained variational problems. The key observation here is that, when passing to the dual bundles, the inclusion arrow $`𝒟TM`$ reverses and gives rise to a projection arrow $`TM^{}𝒟^{}`$; thus, while a constrained Lagrangian $`L:𝒟IR`$ has no canonical extension to a Lagrangian on $`TM`$, its Legendre transform $`H_0:𝒟^{}IR`$ naturally induces a map $`H:TM^{}IR`$ given by the composition of $`H_0`$ and the projection $`TM^{}𝒟^{}`$. Our main result (Theorem 4.1) is that the critical points of the constrained action functional $``$ are the solutions of the Hamilton equations of $`H`$ satisfying suitable boundary conditions. Observe that, unless $`𝒟=TM`$, the Hamiltonian $`H`$ is always degenerate and thus it cannot arise as the Legendre transform of a hyper-regular Lagrangian on the whole tangent bundle $`TM`$. In the particular case where $`P`$ and $`Q`$ are single points of $`M`$, $`𝒟`$ is endowed with a smoothly varying positive definite inner product $`g`$ and $`L`$ is given by $`L(t,q,\dot{q})=\frac{1}{2}g(\dot{q},\dot{q})`$, then the solutions of the corresponding Hamiltonian $`H`$ are known in the context of sub-Riemannian geometry as the normal extremals of $`(M,𝒟,g)`$. The critical points of the constraint defining $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ are the abnormal extremals. In particular, we obtain a variational proof of \[10, Theorem 1\]. By adding a potential energy term to $`L`$, the Hamilton equations of $`H`$ become the equations of motion for the Vakonomic mechanics (see ). Theorem 4.1 thus provides a unifying approach for the study of Lagrangian variational problems with linear constraints in the derivative; it also provides the appropriate setting for the study of the second variation of a constrained Lagrangian action functional and for the development of an index theory for such functional using the notion of Maslov index for a solution of a Hamiltonian (see ). The proof of Theorem 4.1 is based on the method of Lagrangian multipliers, which is used to pass from a constrained Lagrangian variational problem to a non constrained one. The main technical difficulty is the proof of the regularity of the Lagrangian multiplier (Lemma 4.9); such proof is based on a suitable version of Schwartz’s generalized functions calculus which is developed in Subsection 4.1. We give a brief description of the material presented in each section of the paper. In Subsection 2.1 we describe a general notion of Legendre transform. In Subsection 2.2 we recall some standard results concerning the correspondence between hyper-regular Lagrangians and Hamiltonians and in Section 3 we present some well-known facts about the manifold structure of the set of horizontal curves connecting two fixed submanifolds of a given manifold. In Section 4 we state the main result of the paper (Theorem 4.1), that establishes the correspondence between the critical points of the action functional of a hyper-regular constrained Lagrangian and the solutions of the corresponding degenerate Hamiltonian. The proof of Theorem 4.1 is given in Subsection 4.2. In Subsection 4.1 it is presented a suitable version of Schwartz’s generalized functions calculus, needed for technical reasons in the proof of Theorem 4.1. ## 2. The Legendre Transform. <br>Lagrangians and Hamiltonians on Manifolds In this section we recall some classical results from which are presented in a more general context needed for the statement and the proof of Theorem 4.1. In Subsection 2.1 we present a general version of the Legendre transform for vector spaces; we then apply it fiberwise to obtain a notion of Legendre transform for fiber bundles. In Subsection 2.2 we present the classical Hamiltonian formulation for the variational problem corresponding to a hyper-regular (non constrained) Lagrangian. The standard results from are proven in a slightly more general setup; namely, we consider curves with endpoints varying in submanifolds, time-dependent Lagrangians and rather weak regularity assumptions for the data. ### 2.1. The Legendre transform Let $`\xi _0`$ be a real finite-dimensional vector space, let $`\xi _0^{}`$ denote its dual, and let $`Z:UIR`$ be a function of class $`C^2`$ defined on an open subset $`U\xi _0`$. ###### Definition 2.1. Assume that the differential $`\mathrm{d}Z`$ is a diffeomorphism onto an open subset $`V\xi _0^{}`$. The Legendre transform of $`Z`$ is the $`C^1`$ map $`Z^{}:VIR`$ defined by: (2.1) $$Z^{}=E_Z(\mathrm{d}Z)^1,$$ where $`E_Z:UIR`$ is given by (2.2) $$E_Z(v)=\mathrm{d}Z(v)vZ(v),vU.$$ ###### Lemma 2.2. Using the canonical identification of $`\xi _0`$ and its bi-dual $`\xi _0^{}`$, the map $`\mathrm{d}Z^{}`$ is the inverse of $`\mathrm{d}Z`$. Therefore, $`Z^{}`$ is a map of class $`C^2`$. ###### Proof. Differentiating the equality $`Z^{}\mathrm{d}Z=E_Z`$ and (2.2), we obtain: $$\mathrm{d}Z^{}\left(\mathrm{d}Z(v)\right)\mathrm{d}^2Z(v)=\mathrm{d}E_Z(v),\mathrm{d}E_Z(v)=\widehat{v}\mathrm{d}^2Z(v),$$ where $`\widehat{v}\xi _0^{}`$ denotes evaluation at $`v`$. Since $`\mathrm{d}^2Z(v):\xi _0\xi _0^{}`$ is an isomorphism, the conclusion follows. ∎ ###### Corollary 2.3. $`Z^{}=Z`$. ###### Proof. By Lemma 2.2, we have: $$Z^{}=E_Z^{}(\mathrm{d}Z^{})^1=E_Z^{}\mathrm{d}Z.$$ Hence, by definition of $`E_Z^{}`$, we get: $$\begin{array}{cc}\hfill E_Z^{}\left(\mathrm{d}Z(v)\right)& =\mathrm{d}Z^{}\left(\mathrm{d}Z(v)\right)\mathrm{d}Z(v)Z^{}\left(\mathrm{d}Z(v)\right)=\hfill \\ & =\mathrm{d}Z(v)vE_Z(v)=Z(v).\mathit{}\hfill \end{array}$$ Let now $`M`$ be a smooth manifold and $`\pi :\xi M`$ be a smooth vector bundle over $`M`$; for $`mM`$, we denote by $`\xi _m`$ the fiber $`\pi ^1(m)`$. The dual bundle of $`\xi `$ will be denoted by $`\xi ^{}`$; the bi-dual $`\xi ^{}`$ is canonically identified with $`\xi `$. Let $`Z:U\xi IR`$ be a map such that, for every $`mM`$, $`U\xi _m`$ is open in $`\xi _m`$ and the restriction of $`Z`$ to $`U\xi _m`$ is of class $`C^2`$. ###### Definition 2.4. The fiber derivative $`𝔽Z:U\xi ^{}`$ is the map defined by: (2.3) $$𝔽Z(v)=\mathrm{d}(Z|_{U\xi _m})(v),vU,$$ where $`m=\pi (v)`$. Let $`V\xi ^{}`$ denote the image of $`𝔽Z`$. We say that $`Z`$ is regular if for each $`mM`$, the set $`V\xi _m`$ is open in $`\xi _m`$ and the restriction of $`𝔽Z`$ to $`U\xi _m`$ is a local diffeomorphism; $`Z`$ is said to be hyper-regular if for each $`m`$ such restriction is a diffeomorphism onto $`V\xi _m^{}`$. If $`Z`$ is hyper-regular, we define the Legendre transform of $`Z`$ as the map $`Z^{}:VIR`$ whose restriction to $`V\xi _m`$ is the Legendre transform of the restriction of $`Z`$ to $`U\xi _m`$. In analogy with (2.2) we also set: (2.4) $$E_Z(v)=𝔽Z(v)vZ(v),vU;$$ obviously $`Z^{}=E_Z𝔽Z^1`$. Applying Lemma 2.2 and Corollary 2.3 fiberwise, we obtain immediately the following: ###### Proposition 2.5. Assume that $`Z:U\xi IR`$ is hyper-regular. Then, for each $`mM`$, the restriction of $`Z^{}`$ to $`V\xi _m^{}`$ is of class $`C^2`$. Moreover, $`𝔽Z`$ and $`𝔽Z^{}`$ are mutually inverse bijections and $`Z^{}=Z`$.∎ ### 2.2. Time dependent Lagrangians and Hamiltonians on manifolds Let $`M`$ be a smooth $`n`$-dimensional manifold and let $`TM`$, $`TM^{}`$ denote respectively the tangent and the cotangent bundle of $`M`$; with a slight abuse of notation, we will denote both the projections of $`TM`$ and of $`TM^{}`$ by $`\pi `$. Consider the following vector bundles: $$\xi =IR\times TM\stackrel{\mathrm{Id}\times \pi }{}IR\times M,\xi ^{}=IR\times TM^{}\stackrel{\mathrm{Id}\times \pi }{}IR\times M.$$ Observe that the fiber $`\xi _{(t,m)}`$ is $`\{t\}\times T_mM`$ and that $`\xi _{(t,m)}^{}=\{t\}\times T_mM^{}`$. ###### Definition 2.6. A (time-dependent) Lagrangian on $`M`$ is a function $`L:U\xi IR`$ defined on an open set $`U\xi `$ and satisfying the following regularity conditions: 1. $`L`$ is continuous; 2. for each $`tIR`$, the map $`L(t,)`$ is of class $`C^1`$ on $`U\left(\{t\}\times TM\right)`$ and its differential is continuous on $`U`$; 3. for each $`tIR`$, the map $`𝔽L(t,):U\left(\{t\}\times TM\right)\{t\}\times TM^{}`$ is of class $`C^1`$. A (time-dependent) Hamiltonian on $`M`$ is a function $`H:V\xi ^{}IR`$ defined on an open set $`V\xi ^{}`$ and satisfying the following regularity conditions: 1. for all $`tIR`$, the map $`H(t,)`$ is of class $`C^1`$ on $`V\left(\{t\}\times TM^{}\right)`$; 2. for each $`(t,m)IR\times M`$, the restriction of $`H`$ to $`V\xi _{(t,m)}^{}`$ is of class $`C^2`$. We use the notions of regularity and hyper-regularity given in Definition 2.4 for Lagrangians and Hamiltonians on manifolds. Using the Legendre transform defined in Subsection 2.1 (Definition 2.4), given a hyper-regular Lagrangian $`L`$ on $`M`$, the map $`H=L^{}`$ is a hyper-regular Hamiltonian on $`M`$. Namely, the fact that $`H(t,)`$ is of class $`C^1`$ follows by applying the Inverse Function Theorem to the map $`𝔽L(t,)`$; moreover, the fact that $`V=𝔽L(U)`$ is open in $`\xi ^{}`$ follows from the Theorem of Invariance of Domain (see ) by observing that $`𝔽L`$ is continuous and injective<sup>1</sup><sup>1</sup>1As a matter of fact, this same argument shows that $`𝔽L:UV`$ is a homeomorphism and therefore the Hamiltonian $`H=L^{}`$ is continuous.. If $`H`$ is the hyper-regular Hamiltonian obtained by Legendre transform from the Lagrangian $`L`$, then by Proposition 2.5, we have that $`H^{}=L`$, and that $`𝔽H`$ and $`𝔽L`$ are mutually inverse bijections. In order to simplify the notation, in what follows we will write: $$𝔽L(t,v)=(t,𝔽L^{(2)}(t,v)),𝔽H(t,p)=(t,𝔽H^{(2)}(t,p)),$$ so that $`𝔽L^{(2)}`$ and $`𝔽H^{(2)}`$ are respectively a $`TM^{}`$-valued and a $`TM`$-valued map. Let $`L:UIR\times TMIR`$ be a Lagrangian on $`M`$ and $`\gamma :[a,b]M`$ be a curve of class $`C^1`$, with $`(t,\dot{\gamma }(t))U`$ for all $`t`$. The action $`(\gamma )`$ of $`L`$ on the curve $`\gamma `$ is given by the integral: (2.5) $$(\gamma )=_a^bL(t,\dot{\gamma }(t))dt.$$ $``$ defines a functional on the set: (2.6) $$\begin{array}{c}\mathrm{\Omega }_{PQ}([a,b],M;U)\hfill \\ \hfill =\{\gamma :[a,b]\stackrel{C^1}{}M:\gamma (a)P,\gamma (b)Q,(t,\dot{\gamma }(t))U,t[a,b]\},\end{array}$$ where $`P`$ and $`Q`$ are two smooth embedded submanifolds of $`M`$. It is well known that $`\mathrm{\Omega }_{PQ}([a,b],M;U)`$ has the structure of an infinite dimensional smooth Banach manifold (see for instance ), and $``$ is a functional of class $`C^1`$ on $`\mathrm{\Omega }_{PQ}([a,b],M;U)`$. We will call $``$ the action functional associated to the Lagrangian $`L`$. We have the following characterization of the critical points of $``$: ###### Proposition 2.7. A curve $`\gamma \mathrm{\Omega }_{PQ}([a,b],M;U)`$ is a critical point of $``$ if and only if the following three conditions are satisfied: 1. $`𝔽L^{(2)}(a,\dot{\gamma }(a))|_{T_{\gamma (a)P}}=0`$ and $`𝔽L^{(2)}(b,\dot{\gamma }(b))|_{T_{\gamma (b)Q}}=0`$; 2. $`t𝔽L(t,\dot{\gamma }(t))`$ is of class $`C^1`$; 3. for all $`[t_0,t_1][a,b]`$ and for any chart $`q=(q_1,\mathrm{},q_n)`$ on $`M`$ whose domain contains $`\gamma \left([t_0,t_1]\right)`$, the Euler–Lagrange equation is satisfied in $`[t_0,t_1]`$: (2.7) $$\frac{\mathrm{d}}{\mathrm{d}t}\frac{L}{\dot{q}}(t,q(t),\dot{q}(t))=\frac{L}{q}(t,q(t),\dot{q}(t)),$$ where $`L(t,q,\dot{q})`$ denotes the coordinate representation of $`L`$. ###### Proof. Let $`\gamma \mathrm{\Omega }_{PQ}([a,b],M;U)`$ be a critical point of $``$. Let $`[t_0,t_1][a,b]`$ be an interval and consider a chart $`q=(q_1,\mathrm{},q_n)`$ in $`M`$ whose domain contains $`\gamma \left([t_0,t_1]\right)`$. Choose an arbitrary $`vT_\gamma \mathrm{\Omega }_{PQ}([a,b],M;U)`$ with support contained in $`]t_0,t_1[`$; by standard computations it follows that: (2.8) $$_{t_0}^{t_1}\frac{L}{q}(t,q(t),\dot{q}(t))v(t)+\frac{L}{\dot{q}}(t,q(t),\dot{q}(t))\dot{v}(t)\mathrm{d}t=0.$$ The fact that the equality above holds for every smooth $`v`$ with support contained in $`]t_0,t_1[`$ implies that the term $`\frac{L}{\dot{q}}(t,q(t),\dot{q}(t))`$ is of class $`C^1`$; this will follow<sup>2</sup><sup>2</sup>2Alternatively, one could use integration by parts and the fact that $`_{t_0}^{t_1}\varphi \dot{v}=0`$ for all smooth $`v`$ with support in $`]t_0,t_1[`$ implies $`\varphi \text{constant}`$. from the generalized functions calculus developed in Subsection 4.1 (see Corollary 4.5). Integration by parts in (2.8) and the Fundamental Lemma of Calculus of Variations imply then that equation (2.7) is satisfied. Observe also that the coordinate representation of the map $`t𝔽L^{(2)}(t,\dot{\gamma }(t))`$ is given by $`t\frac{L}{\dot{q}}(t,q(t),\dot{q}(t))`$, so that condition (2) is satisfied. Condition (1) follows easily by integrating by parts (2.8) in intervals of the form $`[a,t_1]`$ and $`[t_0,b]`$. Conversely, if conditions (1), (2) and (3) are satisfied, equality (2.8) follows easily, which implies that $`\mathrm{d}_\gamma (v)=0`$ for all $`vT_\gamma \mathrm{\Omega }_{PQ}([a,b],M;U)`$ with small support. Since such $`v`$’s span $`T_\gamma \mathrm{\Omega }_{PQ}([a,b],M;U)`$, it follows that $`\gamma `$ is a critical point of $``$. ∎ We now pass to the study of the Hamiltonian formalism, and we consider the canonical symplectic form $`\omega `$ on $`TM^{}`$, given by $`\omega =\mathrm{d}\vartheta `$, where the canonical $`1`$-form $`\vartheta `$ on $`TM^{}`$ is defined by $`\vartheta _p(\zeta )=p\left(\mathrm{d}\pi _p(\zeta )\right)`$, for all $`pTM^{}`$, $`\zeta T_pTM^{}`$. If $`q=(q_1,\mathrm{},q_n)`$ is a chart on $`M`$ and $`(q,p)=(q_1,\mathrm{},q_n,p_1,\mathrm{},p_n)`$ is the corresponding chart on $`TM^{}`$, the forms $`\vartheta `$ and $`\omega `$ are given by: (2.9) $$\vartheta =\underset{i=1}{\overset{n}{}}p_i\mathrm{d}q_i,\omega =\underset{i=1}{\overset{n}{}}\mathrm{d}q_i\mathrm{d}p_i.$$ Given a Hamiltonian $`H`$ on $`M`$, we define its Hamiltonian vector field $`\stackrel{}{H}`$ to be the unique time-dependent vector field on $`TM^{}`$ satisfying: $$\omega (\stackrel{}{H},)=\mathrm{d}H_t,$$ where $`H_t=H(t,)`$. We say that a curve $`\gamma :[a,b]M`$ is a solution of the Hamiltonian $`H`$ if there exists a $`C^1`$-curve $`\mathrm{\Gamma }:[a,b]TM^{}`$ with $`\pi \mathrm{\Gamma }=\gamma `$ and such that: (2.10) $$\frac{\mathrm{d}}{\mathrm{d}t}\mathrm{\Gamma }(t)=\stackrel{}{H}(t,\mathrm{\Gamma }(t))$$ for all $`t`$. In this case, we say that $`\mathrm{\Gamma }`$ is a Hamiltonian lift of $`\gamma `$. In coordinates $`(q,p)`$, equation (2.10) is written as: (2.11) $$\{\begin{array}{cc}\hfill \frac{\mathrm{d}q}{\mathrm{d}t}=& \frac{H}{p}(t,q(t),p(t)),\hfill \\ \hfill \frac{\mathrm{d}p}{\mathrm{d}t}=& \frac{H}{q}(t,q(t),p(t)).\hfill \end{array}$$ These are called the Hamilton equations of $`H`$; observe that the first equation in (2.11) can be written intrinsically as: (2.12) $$\dot{\gamma }(t)=𝔽H^{(2)}(t,\mathrm{\Gamma }(t)).$$ ###### Theorem 2.8. Let $`L`$ be a hyper-regular Lagrangian on $`M`$ and let $`H=L^{}`$ be the corresponding hyper-regular Hamiltonian. Let $`P`$ and $`Q`$ be smooth submanifolds of $`M`$; a curve $`\gamma \mathrm{\Omega }_{PQ}([a,b],M;U)`$ is a critical point of $``$ if and only if $`\gamma `$ is a solution of the Hamiltonian $`H`$ which admits a Hamiltonian lift $`\mathrm{\Gamma }`$ such that (2.13) $$\mathrm{\Gamma }(a)|_{T_{\gamma (a)}P}=0,\mathrm{\Gamma }(b)|_{T_{\gamma (b)}Q}=0.$$ ###### Proof. Let $`\gamma \mathrm{\Omega }_{PQ}([a,b],M;U)`$ be a critical point of $``$; set $`\mathrm{\Gamma }(t)=𝔽L^{(2)}(t,\dot{\gamma }(t))`$. Since $`𝔽H`$ and $`𝔽L`$ are mutually inverse, equation (2.12) follows. Moreover, by Proposition 2.7, $`\mathrm{\Gamma }`$ is of class $`C^1`$ and (2.13) holds. We now prove that the second Hamilton equation holds, using a chart $`(q,p)`$ of $`TM^{}`$. To this aim, we differentiate with respect to $`q`$ the equality: $$H(t,q,\frac{L}{\dot{q}}(t,q,\dot{q}))=\frac{L}{\dot{q}}(t,q,\dot{q})\dot{q}L(t,q,\dot{q}),$$ obtaining: (2.14) $$\frac{H}{q}(t,q,p)+\frac{H}{p}(t,q,p)\frac{^2L}{q\dot{q}}(t,q,\dot{q})=\frac{^2L}{q\dot{q}}(t,q,\dot{q})\dot{q}\frac{L}{q}(t,q,\dot{q}),$$ where $`p=\frac{L}{\dot{q}}(t,q,\dot{q})`$. Using that $`𝔽H`$ and $`𝔽L`$ are mutually inverse, we get $`\frac{H}{p}(t,q,p)=\dot{q}`$; it follows from (2.14) that: (2.15) $$\frac{H}{q}(t,q,p)=\frac{L}{q}(t,q,\dot{q}).$$ The second Hamilton equation now follows from formula (2.15) and from the Euler–Lagrange equation (2.7). Conversely, suppose that $`\gamma `$ is a solution of the Hamiltonian $`H`$ which admits a Hamiltonian lift $`\mathrm{\Gamma }`$ satisfying (2.13). Since $`𝔽H`$ and $`𝔽L`$ are mutually inverse, from (2.12) it follows that $`\mathrm{\Gamma }(t)=𝔽L^{(2)}(t,\dot{\gamma }(t))`$. Finally, equality (2.15) and the second Hamilton equation imply the Euler–Lagrange equation (2.7), and the conclusion follows from Proposition 2.7. ∎ ## 3. The Space of Horizontal Curves and its Differentiable Structure In this section we recall some results concerning the manifold structure of the set of horizontal curves connecting two fixed submanifolds of a given manifold. Most of the material presented here is well-known in the context of sub-Riemannian geometry (see ). Detailed proofs can be found in . Actually, some minor adaptations of the proofs of have to be made due to the fact that deals with curves of Sobolev class $`H^1`$ while we have to deal here<sup>3</sup><sup>3</sup>3This is due to the fact that the sub-Riemannian energy functional studied in is smooth on the space of $`H^1`$ curves while the action functional of an arbitrary Lagrangian is not in general even well-defined on such space. with curves of class $`C^1`$. Throughout the section we consider fixed an $`n`$-dimensional differentiable manifold $`M`$ and a smooth distribution $`𝒟TM`$ on $`M`$ of rank $`kn`$. By a horizontal curve we mean a curve $`\gamma :[a,b]M`$ of class $`C^1`$ with $`\gamma ^{}(t)𝒟`$ for all $`t[a,b]`$. Given smooth embedded submanifolds $`P,QM`$ we consider the following spaces: $`\mathrm{\Omega }([a,b],M)`$ $`=\{\gamma :[a,b]M:\gamma \text{ is of class }C^1\};`$ $`\mathrm{\Omega }_P([a,b],M)`$ $`=\{\gamma \mathrm{\Omega }([a,b],M):\gamma (a)P\};`$ $`\mathrm{\Omega }_{PQ}([a,b],M)`$ $`=\{\gamma \mathrm{\Omega }([a,b],M):\gamma (a)P,\gamma (b)Q\};`$ $`\mathrm{\Omega }([a,b],M,𝒟)`$ $`=\{\gamma \mathrm{\Omega }([a,b],M):\gamma \text{ is horizontal}\};`$ $`\mathrm{\Omega }_P([a,b],M,𝒟)`$ $`=\mathrm{\Omega }_P([a,b],M)\mathrm{\Omega }([a,b],M,𝒟);`$ $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ $`=\mathrm{\Omega }_{PQ}([a,b],M)\mathrm{\Omega }([a,b],M,D).`$ It is well-known that $`\mathrm{\Omega }([a,b],M)`$ has a natural structure of a Banach manifold (see for instance ) and that $`\mathrm{\Omega }_P([a,b],M)`$ and $`\mathrm{\Omega }_{PQ}([a,b],M)`$ are embedded Banach submanifolds of $`\mathrm{\Omega }([a,b],M)`$. Also $`\mathrm{\Omega }_P([a,b],M,𝒟)`$ is an embedded Banach submanifold of $`\mathrm{\Omega }([a,b],M)`$. The proof of this fact is obtained by using a suitable atlas for $`\mathrm{\Omega }([a,b],M)`$ whose construction is described below. If $`\xi `$ is a vector bundle over $`M`$ then a time-dependent referential of $`\xi `$ over an open subset $`AIR\times M`$ is a family $`(X_i)_{i=1}^k`$ of smooth maps $`X_i:A\xi `$ such that $`\left(X_i(t,m)\right)_{i=1}^k`$ is a basis of the fiber $`\xi _m`$ for all $`(t,m)A`$. Given a time-dependent referential $`(X_i)_{i=1}^n`$ of the tangent bundle $`TM`$ over an open subset $`AIR\times M`$, we define a map: $$:\mathrm{\Omega }([a,b],M;\widehat{A})C^0([a,b],IR^n),$$ by $`(\gamma )=(h_1,\mathrm{},h_n)`$, where: $$\gamma ^{}(t)=\underset{i=1}{\overset{n}{}}h_i(t)X_i(t,\gamma (t)),$$ for all $`t[a,b]`$ and: (3.1) $`\widehat{A}`$ $`=\{(t,v)IR\times TM:(t,\pi (v))A\},`$ (3.2) $`\mathrm{\Omega }([a,b],M;\widehat{A})`$ $`=\{\gamma \mathrm{\Omega }([a,b],M):(t,\gamma ^{}(t))\widehat{A},\text{for all}t[a,b]\}.`$ ###### Lemma 3.1. If $`\varphi :UM\stackrel{~}{U}IR^n`$ is a local chart on $`M`$ and $``$ is defined as above then the map: (3.3) $$\{\gamma \mathrm{\Omega }([a,b],M;\widehat{A}):\gamma (a)U\}\gamma (\varphi (\gamma (a)),(\gamma ))IR^n\times C^0([a,b],IR^n),$$ is a local chart on the Banach manifold $`\mathrm{\Omega }([a,b],M)`$. ###### Proof. It is a simple application of the Inverse Function Theorem on Banach manifolds (see \[17, Corollary 4.2\] for details on a similar construction). ∎ The proposition below implies that the local charts defined on Lemma 3.1 form an atlas for $`\mathrm{\Omega }([a,b],M)`$. ###### Proposition 3.2. Let $`\xi `$ be a vector bundle over a differentiable manifold $`M`$. Given a continuous curve $`\gamma :[a,b]M`$, there exists a time-dependent referential $`(X_i)_{i=1}^k`$ of $`\xi `$ whose domain $`A`$ is an open neighborhood of the graph of $`\gamma `$ in $`IR\times M`$, i.e., $`(t,\gamma (t))A`$ for all $`t[a,b]`$. ###### Proof. See \[17, Lemma 2.3\]. ∎ Using the atlas constructed above we can prove easily that $`\mathrm{\Omega }_P([a,b],M,𝒟)`$ is a submanifold of $`\mathrm{\Omega }([a,b],M)`$. ###### Proposition 3.3. $`\mathrm{\Omega }_P([a,b],M,𝒟)`$ is an embedded Banach submanifold of $`\mathrm{\Omega }([a,b],M)`$. ###### Proof. Applying Proposition 3.2 to the vector bundle $`𝒟`$ and to a complementary vector bundle of $`𝒟`$ in $`TM`$ we obtain a time-dependent referential $`(X_i)_{i=1}^n`$ of $`TM`$ such that $`(X_i)_{i=1}^k`$ is a time-dependent referential for $`𝒟`$; moreover, we may choose $`(X_i)_{i=1}^n`$ so that its domain $`AIR\times M`$ contains the graph of any prescribed continuous curve in $`M`$. If $`\varphi `$ is a local chart of $`M`$ which sends $`P`$ to an open subset of $`IR^rIR^r\times \{0\}IR^n`$ then the corresponding chart (3.3) on $`\mathrm{\Omega }([a,b],M)`$ sends $`\mathrm{\Omega }_P([a,b],M,𝒟)`$ to an open subset of $`IR^r\times C^0([a,b],IR^k)`$. ∎ Given Banach manifolds $``$, $`𝒩`$, recall that a map $`f:𝒩`$ of class $`C^1`$ is said to be a submersion at a point $`x`$ if the differential $`\mathrm{d}f_x:T_xT_{f(x)}𝒩`$ is surjective and its Kernel $`\mathrm{Ker}(\mathrm{d}f_x)`$ is complemented in $`T_x`$, i.e., it admits a closed complementary subspace in $`T_x`$. When $`f`$ is a submersion at $`x`$, then the intersection of $`f^1\left(f(x)\right)`$ with some open neighborhood of $`x`$ in $``$ is a Banach submanifold of $``$ whose tangent space at $`x`$ is $`\mathrm{Ker}(\mathrm{d}f_x)`$. More generally, if $`𝒫𝒩`$ is a Banach submanifold of $`𝒩`$ and $`xf^1(𝒫)`$ then we say that $`f`$ is transverse to $`𝒫`$ at $`x`$ if the composition of $`\mathrm{d}f_x`$ with the quotient map $`T_{f(x)}𝒩T_{f(x)}𝒩/T_{f(x)}𝒫`$ is surjective and has complemented kernel in $`T_x`$; equivalently, $`f`$ is transverse to $`𝒫`$ at $`x`$ if $`\mathrm{Im}(\mathrm{d}f_x)+T_{f(x)}𝒫=T_{f(x)}𝒩`$ and $`\mathrm{d}f_x^1(T_{f(x)}𝒫)`$ is complemented in $`T_x`$. If $`f`$ is transverse to $`𝒫`$ at $`x`$ then the intersection of $`f^1(𝒫)`$ with some open neighborhood of $`x`$ in $``$ is a Banach submanifold of $``$ whose tangent space at $`x`$ is $`\mathrm{d}f_x^1(T_{f(x)}𝒫)`$. ###### Definition 3.4. A curve $`\gamma \mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ is called regular in $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ if the endpoint map: (3.4) $$\mathrm{\Omega }_P([a,b],M,𝒟)\mu \mu (b)M$$ is transverse to $`Q`$ at the point $`\gamma `$. When $`\gamma `$ is not regular in $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$, we say that $`\gamma `$ is singular in $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$. Since $`M`$ is finite-dimensional, a curve $`\gamma `$ is regular in $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ if and only if the image of the differential of (3.4) at $`\gamma `$ plus $`T_{\gamma (b)}Q`$ equals $`T_{\gamma (b)}M`$. Below we described an explicit method for computing the image of the differential of the endpoint map. ###### Definition 3.5. Denote by $`𝒟^\mathrm{o}TM^{}`$ the annihilator of $`𝒟`$. A curve $`\eta :[a,b]TM^{}`$ of class $`C^1`$ is called a characteristic for $`𝒟`$ if $`\eta \left([a,b]\right)𝒟^\mathrm{o}`$ and $`\eta ^{}(t)T_{\eta (t)}𝒟^\mathrm{o}`$ belongs to the kernel of the restriction of $`\omega _{\eta (t)}`$ to $`T_{\eta (t)}𝒟^\mathrm{o}`$ (recall (2.9)). ###### Proposition 3.6. The annihilator of the image of the differential of (3.4) at a curve $`\gamma `$ is the subspace of $`T_{\gamma (b)}M^{}`$ given by: $$\{\eta (b):\eta \text{ is a characteristic of }𝒟,\pi \eta =\gamma ,\eta (a)|_{T_{\gamma (a)}P}=0\}.$$ ###### Proof. The proof is a minor adaptation of the proof of \[17, Theorem 4.9\] where we consider the case that $`P`$ is a point and we use $`H^1`$ curves instead of $`C^1`$ curves. ∎ ###### Corollary 3.7. A curve $`\gamma \mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ is singular in $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ if and only if there exists a non zero characteristic $`\eta :[a,b]TM^{}`$ of $`𝒟`$ with $`\pi \eta =\gamma `$ and $`\eta (a)|_{T_{\gamma (a)}P}=0`$, $`\eta (b)|_{T_{\gamma (b)}Q}=0`$. ###### Proof. Follows from Proposition 3.6 observing that a characteristic $`\eta :[a,b]TM^{}`$ that vanishes at some $`t_0[a,b]`$ is identically zero (see \[17, Lemma 4.8\]). ∎ ###### Corollary 3.8. If either $`T_{\gamma (a)}P+𝒟_{\gamma (a)}=T_{\gamma (a)}M`$ or $`T_{\gamma (b)}Q+𝒟_{\gamma (b)}=T_{\gamma (b)}M`$ then $`\gamma `$ is regular in $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$.∎ ###### Remark 3.9. If the distribution $`𝒟`$ satisfies a strong non integrability condition (for instance, if $`𝒟`$ is a contact distribution) then the restriction of the symplectic form $`\omega `$ to the annihilator $`𝒟^\mathrm{o}`$ of $`𝒟`$ is nondegenerate outside the zero section and therefore all non zero characteristic curves of $`𝒟`$ are constant. In particular, every non constant curve in $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ is regular. So far we have looked at $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ as the set of curves $`\gamma `$ in the Banach manifold $`\mathrm{\Omega }_P([a,b],M,𝒟)`$ satisfying the constraint $`\gamma (b)Q`$. We could also think of $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ as the set of curves in the Banach manifold $`\mathrm{\Omega }_{PQ}([a,b],M)`$ satisfying the constraint $`\mathrm{Im}(\gamma ^{})𝒟`$. Actually, the latter point of view will be needed in the proof of our main theorem in Subsection 4.2. Our goal now is to show that both constraints have the same singularities. This fact was shown in in the context of curves of class $`H^1`$. However, in the case of curves of class $`C^1`$ the problem is a little harder due to the fact that not every closed subspace of a Banach space is complemented. We have thus decided to give all the details of the proof. The lemma below is a general principle that says that if a set is defined by two constraints then the singularities of the first in the space defined by the second constraint equals the singularities of the second in the space defined by the first. ###### Lemma 3.10. Let $``$, $`𝒩_1`$, $`𝒩_2`$ be Banach manifolds and $`𝒫_1𝒩_1`$, $`𝒫_2𝒩_2`$ be Banach submanifolds. Assume that we are given maps $`f_i:𝒩_i`$, $`i=1,2`$, of class $`C^1`$ and a point $`xf_1^1(𝒫_1)f_2^1(𝒫_2)`$ such that $`f_i`$ is transverse to $`𝒫_i`$ at $`x`$, $`i=1,2`$. Then the restriction $`f_1|_{f_2^1(𝒫_2)}`$ is transverse to $`𝒫_1`$ at $`x`$ if and only if the restriction $`f_2|_{f_1^1(𝒫_1)}`$ is transverse to $`𝒫_2`$ at $`x`$. ###### Proof. Consider the Banach spaces $`X=T_x`$, $`Y_i=T_{f_i(x)}𝒩_i/T_{f_i(x)}𝒫_i`$, $`i=1,2`$, and the continuous linear maps $`L_i:XY_i`$, $`i=1,2`$, given by composition of $`\mathrm{d}f_i(x)`$ with the quotient map $`T_{f_i(x)}𝒩_iT_{f_i(x)}𝒩_i/T_{f_i(x)}𝒫_i`$. We know that both $`L_1`$ and $`L_2`$ are surjective and have complemented kernel. We have to show that $`L_1|_{\mathrm{Ker}(L_2)}`$ is surjective with complemented kernel if and only if $`L_2|_{\mathrm{Ker}(L_1)}`$ is surjective with complemented kernel. To this aim, observe first that $`L_1|_{\mathrm{Ker}(L_2)}`$ is surjective if and only if $`\mathrm{Ker}(L_1)+\mathrm{Ker}(L_2)=X`$ and the latter condition is symmetric in $`L_1`$ and $`L_2`$. Finally, to complete the proof we show that, given $`i=1,2`$, then $`\mathrm{Ker}(L_1)\mathrm{Ker}(L_2)`$ is complemented in $`\mathrm{Ker}(L_i)`$ if and only if it is complemented in $`X`$. If $`\mathrm{Ker}(L_1)\mathrm{Ker}(L_2)`$ is complemented in $`X`$ then by intersecting a closed complement of $`\mathrm{Ker}(L_1)\mathrm{Ker}(L_2)`$ in $`X`$ with $`\mathrm{Ker}(L_i)`$ we obtain a closed complement of $`\mathrm{Ker}(L_1)\mathrm{Ker}(L_2)`$ in $`\mathrm{Ker}(L_i)`$. Conversely, if $`Z`$ is a closed complement of $`\mathrm{Ker}(L_1)\mathrm{Ker}(L_2)`$ in $`\mathrm{Ker}(L_i)`$ and $`Z^{}`$ is a closed complement of $`\mathrm{Ker}(L_i)`$ in $`X`$ then $`ZZ^{}`$ is a closed complement of $`\mathrm{Ker}(L_1)\mathrm{Ker}(L_2)`$ in $`X`$ because $`X=\mathrm{Ker}(L_i)Z^{}`$ has the product topology of $`\mathrm{Ker}(L_i)`$ and $`Z^{}`$. ∎ We can now prove the following: ###### Proposition 3.11. Let $`(\theta _i)_{i=1}^{nk}`$ be a time-dependent referential of $`𝒟^\mathrm{o}`$ defined over an open subset $`AIR\times M`$; set $`\theta =(\theta _1,\mathrm{},\theta _{nk})`$, so that $`\theta _{(t,m)}:T_mMIR^{nk}`$ is a surjective linear map with $`\mathrm{Ker}(\theta _{(t,m)})=𝒟_m`$ for all $`(t,m)A`$. Consider the map: $$\mathrm{\Theta }:\mathrm{\Omega }_{PQ}([a,b],M;\widehat{A})C^0([a,b],IR^{nk})$$ defined by: $$\mathrm{\Theta }(\gamma )(t)=\theta \left(\gamma ^{}(t)\right),t[a,b].$$ Then $`\gamma `$ is regular in $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ (in the sense of Definition 3.4) if and only if $`\mathrm{\Theta }`$ is a submersion at $`\gamma `$. ###### Proof. Let $`\overline{\mathrm{\Theta }}`$ denote the extension of $`\mathrm{\Theta }`$ to $$\mathrm{\Omega }_P([a,b],M;\widehat{A})=\mathrm{\Omega }_P([a,b],M)\mathrm{\Omega }([a,b],M;\widehat{A})$$ which is again defined by $`\overline{\mathrm{\Theta }}(\gamma )(t)=\theta \left(\gamma ^{}(t)\right)`$. The conclusion will follow by applying Lemma 3.10 with $`=\mathrm{\Omega }_P([a,b],M;\widehat{A})`$, $`𝒩_1=C^0([a,b],IR^{nk})`$, $`𝒫_1=\{0\}`$, $`𝒩_2=M`$, $`𝒫_2=Q`$, $`f_1=\overline{\mathrm{\Theta }}`$ and $`f_2:𝒩_2`$ equal to the endpoint map $`\mu \mu (b)`$. Since $`f_2`$ is obviously a submersion, we only need to show that $`f_1=\overline{\mathrm{\Theta }}`$ is a submersion. Choose a distribution $`𝒟^{}TM`$ with $`TM=𝒟𝒟^{}`$ and let $`(X_i)_{i=1}^{nk}`$ be the time-dependent referential of $`𝒟^{}`$ over $`A`$ which is dual to $`(\theta _i)_{i=1}^{nk}`$, i.e., $`\theta _i(X_j)=1`$ for $`i=j`$ and $`\theta _i(X_j)=0`$ for $`ij`$. Choose a time-dependent referential $`(X_i)_{i=nk+1}^n`$ of $`𝒟`$ over an open neighborhood of the graph of $`\gamma `$. The coordinate representation of $`\mathrm{\Theta }`$ in the chart (3.3) corresponding to $`(X_i)_{i=1}^n`$ is the natural projection of $`IR^nC^0([a,b],IR^n)`$ onto $`C^0([a,b],IR^{nk})`$. This shows that $`\overline{\mathrm{\Theta }}`$ is a submersion and concludes the proof. ∎ ## 4. Lagrangians with linear constraints and degenerate Hamiltonians Let $`M`$ be an $`n`$-dimensional manifold and $`𝒟TM`$ be a smooth distribution of rank $`k`$. We consider $`𝒟`$ as a vector bundle over $`M`$ with projection $`\pi :𝒟M`$. We apply the theory of Subsection 2.1 to the vector bundle $`\xi =IR\times 𝒟`$ over the manifold $`IR\times M`$, with projection $`\mathrm{Id}\times \pi `$. The fiber $`\xi _{(t,m)}`$ is given by $`\{t\}\times 𝒟_m`$. Let $`L:U\xi IR`$ be a map of class $`C^2`$ defined in an open set $`U\xi `$; we assume that $`L`$ is hyper-regular in the sense of Definition 2.4, so that (by the Inverse Function Theorem) the fiber derivative $`𝔽L:UV`$ is a $`C^1`$ diffeomorphism onto an open subset $`V\xi ^{}`$. Let $`H_0=L^{}`$ be the Legendre transform of $`L`$. Then $`H_0:VIR`$ is a map of class $`C^1`$ whose restriction to each fiber of $`\xi ^{}`$ is of class $`C^2`$; moreover, the fiber derivative $`𝔽H_0:VU`$ is the inverse of $`𝔽L`$ (see Proposition 2.5). For every $`pTM^{}`$ we denote by $`p|_𝒟`$ the restriction of $`pT_mM^{}`$ to $`𝒟_m`$. Observe that the restriction map $`TM^{}pp|_𝒟𝒟^{}`$ is the transpose of the vector bundle inclusion $`𝒟TM`$. By composing $`H_0`$ with the restriction map $`TM^{}𝒟^{}`$ we obtain a map $`H:\stackrel{~}{V}IR`$ given by: (4.1) $$H(t,p)=H_0(t,p|_𝒟),(t,p)\stackrel{~}{V},$$ where: $$\stackrel{~}{V}=\{(t,p)IR\times TM^{}:(t,p|_𝒟)V\}.$$ Observe that $`H`$ is a Hamiltonian on $`M`$ (see Definition 2.6) of class $`C^1`$ defined in the open set $`\stackrel{~}{V}IR\times TM^{}`$. We will call $`L`$ a constrained Lagrangian on $`M`$, and $`H`$ the corresponding degenerate Hamiltonian (observe indeed that $`H`$ cannot be regular unless $`𝒟=TM`$). Given any two submanifolds $`P`$ and $`Q`$ of $`M`$ then a constrained Lagrangian $`L`$ on $`M`$ defines an action functional $``$ on $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟;U)`$ by formula (2.5). Our goal is to determine the critical points of $``$. The following is the main result of the paper and its proof is given in Subsection 4.2: ###### Theorem 4.1. Let $`M`$ be an $`n`$-dimensional manifold, $`𝒟TM`$ be a smooth distribution of rank $`k`$ and $`L:UIR\times 𝒟IR`$ be a hyper-regular constrained Lagrangian of class $`C^2`$. Let $`H_0=L^{}`$ be the Legendre transform of $`L`$ and let $`H`$ be the corresponding degenerate Hamiltonian as in (4.1). Fix two submanifolds $`P`$ and $`Q`$ of $`M`$ and let $``$ be the action functional of $`L`$ defined in the space $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟;U)=\mathrm{\Omega }_{PQ}([a,b],M,𝒟)\mathrm{\Omega }_{PQ}([a,b],M;U)`$, given by (2.5). Let $`\gamma \mathrm{\Omega }_{PQ}([a,b],M,𝒟;U)`$ be a regular curve. Then, $`\gamma `$ is a critical point of $``$ if and only if it is a solution of $`H`$ that admits a Hamiltonian lift $`\mathrm{\Gamma }:[a,b]TM^{}`$ with $`\mathrm{\Gamma }(a)|_{T_{\gamma (a)}P}=0`$ and $`\mathrm{\Gamma }(b)|_{T_{\gamma (b)}Q}=0`$. The classical example of a constrained hyper-regular Lagrangian function $`L`$ is given by: (4.2) $$L(t,v)=\frac{1}{2}g(v,v)V\left(\pi (v)\right),$$ where $`g`$ is a sub-Riemannian metric on $`(M,𝒟)`$ (i.e., a smooth Riemannian structure on the vector bundle $`𝒟`$) and $`V:MIR`$ is a map of class $`C^2`$. The fiber derivative $`𝔽L`$ of (4.2) is given by: $$𝔽L(t,v)=g(v,)𝒟^{},$$ so that $`L`$ is indeed hyper-regular. Recalling (2.4), we compute as follows: $$E_L(t,v)=\frac{1}{2}g(v,v)+V\left(\pi (v)\right),v𝒟,$$ $$H_0(t,\rho )=\frac{1}{2}g^1(\rho ,\rho )+V\left(\pi (\rho )\right),\rho 𝒟^{},$$ where $`g^1`$ denotes the induced Riemannian structure on the dual bundle $`𝒟^{}`$. The degenerate Hamiltonian $`H`$ corresponding to (4.2) is thus given by: $$H(t,p)=\frac{1}{2}g^1(p|_𝒟,p|_𝒟)+V\left(\pi (p)\right),pTM^{}.$$ Theorem 4.1 implies that the critical points of the action functional $``$ corresponding to (4.2) on the space $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$ are the solutions of $`H`$ that admit a Hamiltonian lift $`\mathrm{\Gamma }:[a,b]TM^{}`$ satisfying the boundary conditions $`\mathrm{\Gamma }(a)|_{T_{\gamma (a)}P}=0`$ and $`\mathrm{\Gamma }(b)|_{T_{\gamma (b)}Q}=0`$. Observe that (in the case when $`P`$ and $`Q`$ are points) we obtain the equations for the trajectories of the Vakonomic mechanics given in ; when $`V=0`$ we obtain the equations for the normal geodesics of the sub-Riemannian manifold $`(M,𝒟,g)`$ (see ). ###### Remark 4.2. We emphasize that, in general, a minimum of the action functional $``$ may not be a regular curve in $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟)`$, and in this situation it may not satisfy the Hamilton equations of $`H`$. Examples of this phenomenon are given in in the sub-Riemannian case $`L(t,v)=\frac{1}{2}g(v,v)`$. Hence, one can only conclude that a minimum of $``$ is either a solution of the Hamilton equations or the projection of a non null characteristic of $`𝒟`$. ### 4.1. Generalized functions calculus For the proof of Theorem 4.1 we will occasionally have to consider derivatives of functions that are in principle only continuous<sup>4</sup><sup>4</sup>4This situation already occurred in the proof of Proposition 2.7. In that case the difficulty could also be circumvented by a simpler technique.. These derivatives should be understood in the sense of Schwartz’s generalized functions calculus. However, the usual definition of the generalized functions space as the dual of the space of smooth compactly supported maps only allows products of generalized functions by smooth maps. To overcome this difficulty, we introduce a calculus for generalized functions of stronger regularity, that are elements of the dual of a space of functions with weaker regularity. Let $`V`$ be a real finite dimensional vector space. For $`k0`$, we define $`C_0^k([a,b],V)`$ to be the Banach space of $`V`$-valued $`C^k`$ maps on $`[a,b]`$ whose first $`k`$ derivatives vanish at $`a`$ and at $`b`$; we endow it with the standard $`C^k`$-norm. We denote by $`D^k([a,b],V)`$ the dual Banach space of $`C_0^k([a,b],V^{})`$ (dual spaces will always be meant in the topological sense). Denoting by $`L^p([a,b],V)`$ the Banach space of $`V`$-valued measurable functions on $`[a,b]`$ whose $`p`$-th power is Lebesgue integrable, we have an inclusion: (4.3) $$L^1([a,b],V)D^k([a,b],V)$$ defined by $$f,\alpha =_a^b\alpha (t)f(t)dt,fL^1([a,b],V),\alpha C_0^k([a,b],V^{});$$ in the formula above we have denoted by $`f,\alpha `$ the evaluation at $`\alpha `$ of the linear functional which is the image of $`f`$ by (4.3). In what follows we will always identify a function $`fL^1([a,b],V)`$ with its image by (4.3); moreover, the evaluation of $`fD^k([a,b],V)`$ at $`\alpha C_0^k([a,b],V^{})`$ will always be denoted by $`f,\alpha `$. Observe that we have inclusions $`D^kD^{k+1}`$ defined by restriction of the functionals, i.e., $`D^kD^{k+1}`$ is simply the transpose of the inclusion of $`C_0^{k+1}([a,b],V^{})`$ in $`C_0^k([a,b],V^{})`$. We summarize the observations above by the following diagram: $$\mathrm{}C^1C^0L^1D^0D^1\mathrm{}$$ An element $`f`$ of any space $`D^k([a,b],V)`$ is called a generalized function. In what follows, we will occasionally write simply $`C^k`$, $`C_0^k`$, $`D^k`$, $`L^p`$ instead of $`C^k([a,b],V)`$, $`C_0^k([a,b],V)`$, $`D^k([a,b],V)`$, $`L^p([a,b],V)`$. In addition to the standard vector space operations in $`D^k`$, we define the following: * derivative operation: for $`fD^k([a,b],V)`$, we define the derivative of $`f`$ to be the generalized function $`f^{}D^{k+1}([a,b],V)`$ defined by: $$f^{},\alpha =f,\alpha ^{},$$ for all $`\alpha C_0^{k+1}([a,b],V^{})`$; * product operation: for $`fD^k([a,b],V)`$, $`gC^k([a,b],W)`$ and a fixed bilinear map $`V\times WU`$, we define the product $`fgD^k([a,b],U)`$ as follows. The bilinear map $`V\times WU`$ induces a bilinear map $`W\times U^{}V^{}`$ defined by $`(wu^{})(v)=u^{}(vw)`$; we set: $$fg,\alpha =f,g\alpha ,$$ for all $`\alpha C_0^k([a,b],U^{})`$; * restriction operation: for $`fD^k([a,b],V)`$ and $`[c,d][a,b]`$, we set: $$f|_{[c,d]},\alpha =f,\overline{\alpha },$$ for all $`\alpha C_0^k([c,d],V^{})`$, where $`\overline{\alpha }C_0^k([a,b],V^{})`$ is the extension to zero of $`\alpha `$ outside $`[c,d]`$. It is easily seen that when we apply the above operations to elements of $`D^k`$ which correspond to functions then we obtain the standard operations on functions. Moreover, the standard Leibnitz rule for derivatives of products holds for generalized functions, i.e.: $$(fg)^{}=f^{}g+fg^{},$$ for all $`fD^k`$ and $`gC^{k+1}`$. In order to prove some regularity results we present the following elementary lemmas. ###### Lemma 4.3. Let $`fD^k([a,b],V)`$ be such that $`f^{}=0`$. Then $`f`$ is a constant function. ###### Proof. We first consider the case $`V=IR`$. If $`f^{}=0`$, then $`f,\alpha ^{}=0`$ for all $`\alpha C_0^{k+1}([a,b],IR)`$, hence $`f,\beta =0`$ for all $`\beta C_0^k([a,b],IR)`$ with $`_a^b\beta =0`$. Choose $`\beta _0C_0^k([a,b],IR)`$ with $`_a^b\beta _0=1`$; set $`c=f,\beta _0`$. It is easily seen that $`fc`$. For the general case, observe that for all $`\lambda V^{}`$, the product $`\lambda fD^k([a,b],IR)`$ has vanishing derivative, and hence it is constant. Since $`\lambda `$ is arbitrary, it follows that $`f`$ is constant. ∎ ###### Lemma 4.4. Let $`fD^k([a,b],V)`$, $`k1`$; there exists an element $`FD^{k1}([a,b],V)`$ with $`F^{}=f`$. If $`fD^0([a,b],V)`$, there exists $`FL^2([a,b],V)`$ with $`F^{}=f`$. ###### Proof. Consider the map $`\mathrm{d}:C_0^{k+1}C_0^k`$ given by $`\mathrm{d}(\alpha )=\alpha ^{}`$. It is easily seen that $`\mathrm{d}`$ is injective with closed image. It follows that the transpose map $`\mathrm{d}^{}:D^kD^{k+1}`$ is surjective; clearly, the derivative operator for generalized functions is $`\mathrm{d}^{}`$, which proves the first part of the thesis. For the case $`k=0`$, let $`H_0^1`$ denote the Sobolev space of absolutely continuous functions $`\alpha :[a,b]V^{}`$ having square integrable derivative, and such that $`\alpha (a)=\alpha (b)=0`$. Again, the derivation map $`\mathrm{d}:H_0^1L^2`$ is injective and has closed image. Therefore, given $`fD^0`$, we can find $`FL_{}^{2}{}_{}{}^{}L^2`$ with $`\mathrm{d}^{}F=f|_{H_0^1}`$. It follows that $`F^{}=f`$. ∎ ###### Corollary 4.5 (Bootstrap lemma). Let $`f`$ be a generalized function. 1. If $`f^{}D^0`$ then $`fL^2`$; 2. If $`f^{}L^2`$ then $`fC^0`$; 3. If $`f^{}C^0`$ then $`fC^1`$. ###### Proof. We prove, for example, the first item. By Lemma 4.4, we can find $`FL^2`$ with $`F^{}=f^{}`$. By Lemma 4.3, it follows that $`Ff`$ is constant, hence $`fL^2`$. The other items are proven similarly. ∎ We now give a result that shows that regularity of a generalized function is a local property: ###### Lemma 4.6. Let $`\lambda `$ be a generalized function on $`[a,b]`$. Suppose that for all $`t[a,b]`$ there exists $`\epsilon >0`$ such that the restriction $`\lambda |_{[t\epsilon ,t+\epsilon ][a,b]}`$ is of class $`C^k`$, $`k0`$. Then $`\lambda `$ is of class $`C^k`$. ###### Proof. Consider a partition $`a=t_0<t_1<\mathrm{}<t_r=b`$ such that $`f_i=\lambda |_{[t_i,t_{i+2}]}`$ is of class $`C^k`$ for all $`i=0,\mathrm{},r2`$. Since the operation of restriction for generalized functions gives the standard operation of restriction for functions, it follows that: $$f_i|_{[t_{i+1},t_{i+2}]}=\lambda |_{[t_{i+1},t_{i+2}]}=f_{i+1}|_{[t_{i+1},t_{i+2}]},$$ for $`i=0,\mathrm{},r3`$. Hence there exists a $`C^k`$ map $`f`$ on $`[a,b]`$ such that $`f|_{[t_i,t_{i+2}]}=f_i`$ for all $`i=0,\mathrm{},r2`$. We know that $`f,\alpha =\lambda ,\alpha `$ if $`\alpha `$ has support contained in some interval $`]t_i,t_{i+2}[`$; but such $`\alpha `$’s span a dense subspace of the domain of the linear functional $`\lambda `$ and therefore $`\lambda =f`$. ∎ Finally, we need the following result that relates the dual spaces of $`C^0`$ and $`C_0^0`$. For $`t[a,b]`$ and $`\sigma V`$, we denote by $`\delta _t^\sigma C^0([a,b],V^{})^{}`$ the Dirac’s delta, defined by: $$\delta _t^\sigma ,\alpha =\alpha (t)\sigma ,\alpha C^0([a,b],V^{}).$$ ###### Lemma 4.7. If $`\lambda C^0([a,b],V^{})^{}`$ vanishes identically on $`C_0^0([a,b],V^{})`$ then there exist $`\sigma _a`$ and $`\sigma _b`$ in $`V`$ such that: (4.4) $$\lambda =\delta _a^{\sigma _a}+\delta _b^{\sigma _b}.$$ ###### Proof. If $`𝒜`$ denotes the subspace of $`C^0([a,b],V^{})`$ consisting of affine maps $`\alpha (t)=Pt+Q`$ then obviously: $$C^0([a,b],V^{})=C_0^0([a,b],V^{})𝒜.$$ It is easy to see that we can find $`\sigma _a,\sigma _bV`$ such that both sides of (4.4) agree on $`𝒜`$. Since both sides of (4.4) vanish on $`C_0^0([a,b],V^{})`$, the conclusion follows. ∎ ### 4.2. Proof of Theorem 4.1 The proof of Theorem 4.1 is based on the method of Lagrange multipliers, and we start with the precise statement of the result needed for our purposes. ###### Proposition 4.8. Let $``$ be a Banach manifold, $`E`$ a Banach space and $`F:IR`$, $`g:E`$ maps of class $`C^1`$. Let $`pg^1(0)`$ be such that $`g`$ is a submersion at $`p`$. Then, $`p`$ is a critical point for $`f|_{g^1(0)}`$ if and only if there exists $`\lambda E^{}`$ such that $`p`$ is a critical point for the functional $`f_\lambda =f\lambda g`$ in $``$. ###### Proof. The point $`p`$ is critical for $`f|_{g^1(0)}`$ if and only if $`\mathrm{d}f(p)`$ vanishes on $`T_pg^1(0)=\mathrm{Ker}\left(\mathrm{d}g(p)\right)`$. The conclusion follows from elementary functional analysis arguments. ∎ The linear functional $`\lambda E^{}`$ of Proposition 4.8 is called the Lagrange multiplier of the constrained critical point $`p`$; it is easily seen that such $`\lambda `$ is unique. We can now prove the main result of the section. In the argument we will need a regularity result for a Lagrangian multiplier; such proof is postponed to Lemma 4.9. ###### Proof of Theorem 4.1. We start by choosing an arbitrary complementary distribution $`𝒟^{}`$ to $`𝒟`$, i.e., a smooth distribution of rank $`nk`$ in $`M`$ such that $`T_mM=𝒟_m𝒟_m^{}`$ for all $`mM`$; moreover, we fix an arbitrary smooth Riemannian structure $`g`$ on the vector bundle $`𝒟^{}`$. Let $`\pi _𝒟:TM𝒟`$ and $`\pi _𝒟^{}:TM𝒟^{}`$ denote the projections and define an extension $`\stackrel{~}{L}:\stackrel{~}{U}IR\times MIR`$ of $`L`$ by: (4.5) $$\stackrel{~}{L}(t,v)=L(t,\pi _𝒟(v))+\frac{1}{2}g(\pi _𝒟^{}(v),\pi _𝒟^{}(v)),$$ where $$\stackrel{~}{U}=\{(t,v)IR\times TM:(t,\pi _𝒟(v))U\}.$$ Then $`\stackrel{~}{U}`$ is open in $`IR\times TM`$ and $`\stackrel{~}{L}`$ is a Lagrangian on $`M`$ as in Definition 2.6; we denote by $`\stackrel{~}{}`$ the corresponding action functional in $`\mathrm{\Omega }_{PQ}([a,b],M;\stackrel{~}{U})`$, defined as in (2.5). Let $`\theta `$, $`\mathrm{\Theta }`$, $`A`$ and $`\widehat{A}`$ be as in the statement of Proposition 3.11 (recall also (3.1) and (3.2)). Then, since $`\gamma `$ is regular, the map $`\mathrm{\Theta }`$ is a submersion at $`\gamma `$; moreover, $`\gamma `$ is a critical point of $``$ in $`\mathrm{\Omega }_{PQ}([a,b],M,𝒟;U)`$ if and only if it is a critical point of $`\stackrel{~}{}|_{\mathrm{\Theta }^1(0)}`$. By the method of Lagrange multipliers (Proposition 4.8), this is equivalent to the existence of $`\lambda C^0([a,b],IR^{nk})^{}`$ such that $`\gamma `$ is a critical point of $`\stackrel{~}{}_\lambda =\stackrel{~}{}\lambda \mathrm{\Theta }`$ in $`\mathrm{\Omega }_{PQ}([a,b],M;\widehat{A}\stackrel{~}{U})`$. We will prove in Lemma 4.9 below that the Lagrange multiplier $`\lambda `$ is of class $`C^1`$, i.e., that it is given by: (4.6) $$\lambda (\alpha )=_a^b\lambda _0(t)\alpha (t)dt,\alpha C^0([a,b],IR^{nk}),$$ for some $`C^1`$ map $`\lambda _0:[a,b](IR^{nk})^{}`$. Therefore, $`\stackrel{~}{}_\lambda `$ is the action functional corresponding to the Lagrangian $`\stackrel{~}{L}_\lambda `$ in $`M`$ defined by: (4.7) $$\stackrel{~}{L}_\lambda (t,v)=\stackrel{~}{L}(t,v)\lambda _0(t)\theta _{(t,m)}(v),(t,v)\widehat{A}\stackrel{~}{U},$$ where $`m=\pi (v)`$. We now prove that $`\stackrel{~}{L}`$ and $`\stackrel{~}{L}_\lambda `$ are hyper-regular and we compute their Legendre transforms. The fiber derivatives $`𝔽\stackrel{~}{L}`$ and $`𝔽\stackrel{~}{L}_\lambda `$ are easily computed as: (4.8) $$𝔽\stackrel{~}{L}(t,v)=𝔽L(t,\pi _𝒟(v))\pi _𝒟+g(\pi _𝒟^{}(v),\pi _𝒟^{}())T_mM^{},$$ (4.9) $$𝔽\stackrel{~}{L}_\lambda (t,v)=𝔽\stackrel{~}{L}(t,v)\lambda _0(t)\theta _{(t,m)}T_mM^{},$$ where $`m=\pi (v)`$. The hyper-regularity is proven by exhibiting explicit inverses: (4.10) $$\begin{array}{cc}& 𝔽\stackrel{~}{L}^1(t,p)=𝔽L^1(t,p|_𝒟)+g^1(p|_𝒟^{}),\hfill \\ & 𝔽\stackrel{~}{L}_\lambda ^1(t,p)=𝔽\stackrel{~}{L}^1(t,p+\lambda _0(t)\theta _{(t,m)});\hfill \end{array}$$ by $`g^1`$ in the above formula we mean the inverse of $`g`$ seen as a linear map from $`𝒟_m`$ to $`𝒟_m^{}`$. We now compute the Legendre transforms $`\stackrel{~}{H}`$ and $`\stackrel{~}{H}_\lambda `$ of $`\stackrel{~}{L}`$ and $`\stackrel{~}{L}_\lambda `$ respectively. Using Definition 2.1 and equations (4.8), (4.9), we compute easily: (4.11) $$E_{\stackrel{~}{L}_\lambda }(t,v)=E_{\stackrel{~}{L}}(t,v)=E_L(t,\pi _𝒟(v))+\frac{1}{2}g(\pi _𝒟^{}(v),\pi _𝒟^{}(v));$$ and, using (4.10), we therefore obtain: $`\stackrel{~}{H}(t,p)=H(t,p)+\frac{1}{2}g^1(p|_𝒟^{},p|_𝒟^{}),`$ $`\begin{array}{cc}\hfill \stackrel{~}{H}_\lambda (t,p)=& \stackrel{~}{H}(t,p+\lambda _0(t)\theta _{(t,m)})\hfill \\ \hfill =& H(t,p)+\frac{1}{2}g^1((p+\lambda _0(t)\theta _{(t,m)})|_𝒟^{},(p+\lambda _0(t)\theta _{(t,m)})|_𝒟^{}).\hfill \end{array}`$ We now compute the Hamilton equations of the Hamiltonian $`\stackrel{~}{H}_\lambda `$ with the help of local coordinates $`(q_1,\mathrm{},q_n,p_1,\mathrm{},p_n)`$ in $`TM^{}`$ and of a local $`g`$-orthonormal referential $`X_1,\mathrm{},X_{nk}`$ of $`𝒟^{}`$. We write: (4.12) $$\stackrel{~}{H}_\lambda (t,p)=H(t,p)+\frac{1}{2}\underset{i=1}{\overset{nk}{}}\left(p+\lambda _0(t)\theta _{(t,m)}\right)(X_i)^2,$$ and, using (2.11), the Hamilton equations of $`\stackrel{~}{H}_\lambda `$ are given by: (4.13) $$\{\begin{array}{cc}\hfill \frac{\mathrm{d}q}{\mathrm{d}t}& =\frac{H}{p}+\underset{i=1}{\overset{nk}{}}(p+\lambda _0\theta )(X_i)X_i,\hfill \\ \hfill \frac{\mathrm{d}p}{\mathrm{d}t}& =\frac{H}{q}\underset{i=1}{\overset{nk}{}}(p+\lambda _0\theta )(X_i)\left[\lambda _0\frac{\theta }{q}(X_i)+(p+\lambda _0\theta )\left(\frac{X_i}{q}\right)\right].\hfill \end{array}$$ By Theorem 2.8, $`\gamma `$ is a critical point of $`\stackrel{~}{}_\lambda `$ if and only if it admits a lift $`\mathrm{\Gamma }:[a,b]TM^{}`$ satisfying (4.13) with $`\mathrm{\Gamma }(a)T_{\gamma (a)}P^o`$ and $`\mathrm{\Gamma }(b)T_{\gamma (b)}Q^o`$. Now, it follows easily from (4.1) that $`\frac{H}{p}`$ is in $`𝒟`$; since $`\gamma `$ is horizontal, i.e., $`\frac{\mathrm{d}q}{\mathrm{d}t}𝒟`$, from the first equation of (4.13) it follows that $`(p+\lambda _0\theta )(X_i)=0`$ for all $`i=1,\mathrm{},nk`$. Setting $`(p+\lambda _0\theta )(X_i)=0`$ in (4.13) we obtain the Hamilton equations of $`H`$, which concludes the proof. ∎ We are left with the proof of the regularity of the Lagrange multiplier $`\lambda `$. We will use the generalized functions calculus developed in Subsection 4.1. ###### Lemma 4.9. Under the assumptions of Theorem 4.1, using the notations adopted in its proof, if $`\gamma `$ is horizontal and if, for some $`\lambda C^0([a,b],IR^{nk})^{}`$, it is a critical point of $`\stackrel{~}{}\lambda \mathrm{\Theta }`$, then there exists a $`C^1`$ map $`\lambda _0:[a,b](IR^{nk})^{}`$ such that (4.6) holds. ###### Proof. We set $$\lambda _0=\lambda |_{C_0^0([a,b],IR^{nk})}D^0([a,b],(IR^{nk})^{});$$ we first prove the regularity of the generalized function $`\lambda _0`$. To this aim, we localize the problem by considering variational vector fields along $`\gamma `$ having support in the domain of a local chart $`q=(q_1,\mathrm{},q_n)`$ in $`M`$. Let $`[c,d][a,b]`$ be such that $`\gamma \left([c,d]\right)`$ is contained in the domain of the local chart; we still denote by $`\lambda _0`$ the restriction of $`\lambda _0`$ to $`[c,d]`$. Since $`\gamma `$ is a critical point of $`\stackrel{~}{}\lambda \mathrm{\Theta }`$, by standard computations it follows that the following equality holds: (4.14) $$\begin{array}{c}_c^d\frac{\stackrel{~}{L}}{q}(t,q(t),\dot{q}(t))v(t)+\frac{\stackrel{~}{L}}{\dot{q}}(t,q(t),\dot{q}(t))\dot{v}(t)\mathrm{d}t\hfill \\ \hfill \lambda _0,\frac{\theta }{q}|_{(t,q(t))}(v(t),\dot{q}(t))+\theta _{(t,q(t))}\dot{v}(t)=0,\end{array}$$ for every vector field $`v`$ of class $`C^1`$ along $`\gamma `$ having support in $`]c,d[`$; in the formula above we have regarded the derivative $`\frac{\theta }{q}|_{(t,q(t))}`$ as an $`IR^{nk}`$-valued bilinear map in $`IR^n`$. In terms of the local coordinates, the maps $`\theta `$, $`\frac{\theta }{q}(,\dot{q})`$, $`\frac{\stackrel{~}{L}}{q}`$ and $`\frac{\stackrel{~}{L}}{\dot{q}}`$ evaluated along $`\gamma `$ will be interpreted as follows: * $`\theta C^1([c,d],\mathrm{Lin}(IR^n,IR^{nk}))`$; * $`{\displaystyle \frac{\theta }{q}}(,\dot{q})C^0([c,d],\mathrm{Lin}(IR^n,IR^{nk}))`$; * $`{\displaystyle \frac{\stackrel{~}{L}}{q}},{\displaystyle \frac{\stackrel{~}{L}}{\dot{q}}}C^0([c,d],IR_{}^{n}{}_{}{}^{})`$, where $`\mathrm{Lin}(,)`$ denotes the space of linear maps between two given vector spaces. Using the definition of derivative for generalized functions, from (4.14) we get: (4.15) $$\left(\frac{\stackrel{~}{L}}{\dot{q}}\right)^{}+\frac{\stackrel{~}{L}}{q}\lambda _0\frac{\theta }{q}(,\dot{q})+(\lambda _0\theta )^{},v=0,$$ for every $`C^1`$ map $`v:[c,d]IR^n`$ having support in $`]c,d[`$, and, by density, for every $`vC_0^1([c,d],IR^n)`$. It follows: (4.16) $$\left(\frac{\stackrel{~}{L}}{\dot{q}}\right)^{}+\frac{\stackrel{~}{L}}{q}\lambda _0\frac{\theta }{q}(,\dot{q})+\lambda _0^{}\theta +\lambda _0\theta ^{}=0.$$ Let $`X_1,\mathrm{},X_{nk}`$ be a referential of $`𝒟^{}`$ along $`\gamma `$; in terms of the local coordinates the $`X_i`$’s will be thought as elements of $`C^1([c,d],IR^n)`$. Moreover, we set $$X=(X_1,\mathrm{},X_{nk})C^1([c,d],\mathrm{Lin}(IR^{nk},IR^n)),$$ where the $`(nk)`$-tuple $`(X_1(t),\mathrm{},X_{nk}(t))`$ is identified with the linear map that takes the $`i`$-th vector of the canonical basis of $`IR^{nk}`$ to $`X_i(t)`$. Composing (4.16) with $`X`$, we obtain: (4.17) $$\lambda _0^{}\theta (X)+\lambda _0\theta ^{}(X)\lambda _0\frac{\theta }{q}(X,\dot{q})+\frac{\stackrel{~}{L}}{q}X\left(\frac{\stackrel{~}{L}}{\dot{q}}\right)^{}X=0.$$ Evaluating (4.8) at $`X_i`$ with $`v=\gamma ^{}`$ and using the horizontality of $`\gamma `$ we get: (4.18) $$\frac{\stackrel{~}{L}}{\dot{q}}X_i=0,i=1,\mathrm{},nk;$$ hence: (4.19) $$\left(\frac{\stackrel{~}{L}}{\dot{q}}\right)^{}X=\frac{\stackrel{~}{L}}{\dot{q}}X^{}C^0([c,d],(IR^{nk})^{}).$$ Now, considering that $`\theta (X)\mathrm{Lin}(IR^{nk},IR^{nk})`$ is invertible, by (4.19) we can write (4.16) in the form: (4.20) $$\lambda _0^{}=\lambda _0h_1+h_2,$$ with $`h_1C^0([c,d],\mathrm{Lin}(IR^{nk},IR^{nk}))`$ and $`h_2C^0([c,d],(IR^{nk})^{})`$. Applying three times Corollary 4.5, from (4.20) we conclude that $`\lambda _0`$ belongs to the space $`C^1([c,d],(IR^{nk})^{})`$; now Lemma 4.6 implies that $`\lambda _0C^1([a,b],(IR^{nk})^{})`$. By Lemma 4.7, there exist $`\sigma _a,\sigma _b(IR^{nk})^{}`$ such that: (4.21) $$\lambda (\alpha )=_a^b\lambda _0\alpha dt+\sigma _a\alpha (a)+\sigma _b\alpha (b),\alpha C^0([a,b],IR^{nk}).$$ To conclude the proof we show that $`\sigma _a=\sigma _b=0`$. Let’s show for instance that $`\sigma _a=0`$; the proof of the equality $`\sigma _b=0`$ is analogous. Using local charts around $`\gamma \left([a,d]\right)`$, for $`d`$ close to $`a`$, we consider variational vector fields $`v`$ of class $`C^1`$ supported in $`[a,d[`$, with $`v(a)T_{\gamma (a)}P`$. Arguing as in the deduction of formula (4.14), we get the following equality: (4.22) $$\begin{array}{c}_a^d\frac{\stackrel{~}{L}}{q}(t,q(t),\dot{q}(t))v(t)+\frac{\stackrel{~}{L}}{\dot{q}}(t,q(t),\dot{q}(t))\dot{v}(t)\mathrm{d}t\hfill \\ \hfill _a^d\lambda _0(t)\left[\frac{\theta }{q}|_{(t,q(t))}(v(t),\dot{q}(t))+\theta _{(t,q(t))}\dot{v}(t)\right]dt\\ \hfill \sigma _a\left[\frac{\theta }{q}|_{(a,q(a))}(v(a),\dot{q}(a))+\theta _{(a,q(a))}\dot{v}(a)\right]=0.\end{array}$$ From Corollary 4.5 and formula (4.16) it follows that $`\frac{\stackrel{~}{L}}{\dot{q}}`$ is of class $`C^1`$, and we can thus use integration by parts in (4.22) to obtain an equality of the form: (4.23) $$_a^du(t)v(t)dt+\sigma _a\theta _{(a,q(a))}\dot{v}(a)=0,$$ for some $`uC^0([a,d],IR_{}^{n}{}_{}{}^{})`$, whenever $`v`$ is chosen with $`v(a)=0`$. By considering arbitrary $`v`$ supported in $`]a,d[`$, from (4.23) we obtain that $`u0`$ in $`[a,d]`$, so that the integral in (4.23) vanishes for all $`v`$. Now, we can choose $`v`$ with $`v(a)=0`$ and $`\dot{v}(a)`$ arbitrary, so that (4.23) implies that $`\sigma _a=0`$, because $`\theta _{(a,q(a))}`$ is surjective. This concludes the proof. ∎
warning/0004/cond-mat0004364.html
ar5iv
text
# Continuous weak measurement of the macroscopic quantum coherent oscillations ## I Introduction Quantum superposition of macroscopically distinct states is one of the most characteristic features of quantum behavior at the macroscopic level. In the “mesoscopic” regime, when the states involved in the superposition correspond to a collective motion of a number of particles that is larger than one but not quite macroscopic, the superposition of states has been demonstrated for photons in a high-quality microwave cavity , and for the center-of-mass motion of large molecules . Among the most basic dynamic manifestations of quantum superposition of states are quantum coherent oscillations between the two basis states of a two-state system. However, while the macroscopic quantum phenomena brought about by the incoherent quantum tunneling are by now commonly found in a variety of systems ranging from mesoscopic tunnel junctions in the regimes of flux and charge dynamics, to molecular and nano-magnets , the situation with experimental observation of macroscopic quantum coherent (MQC) oscillations remains much more uncertain. Claim of the observation of the MQC oscillations in molecular magnets remain highly controversial . Remarkable experimental demonstration of the MQC oscillations in the charge-dynamics regime of a small Josephson junction is open to criticism that the two charge states of the observed quantum superposition differ in charge only by the the charge of one Cooper pair, and the oscillations between these two states can not be interpreted as macroscopic. Although this criticism is not fully justified, since the charge dynamics of a small Josephson junction is just another representation of its flux dynamics, which is the paradigm of the “macroscopic” quantum dynamics , it does not allow to consider the question of MQC oscillations to be completely settled. Recently, macroscopic quantum dynamics has attracted renewed attention as the possible basis for development of scalable quantum logic circuits for quantum computation. In this context, macroscopic quantum two-state system plays the role of a qubit, an elementary building block of a quantum computer. Several variants of qubits and quantum logic gates have been proposed that are based on the macroscopic quantum dynamics of Josephson junctions. Many characteristics of the macroscopic qubits compare favorably with those of the microscopic qubits: they are insensitive to disorder at the microscopic level and offer much larger freedom in design and fabrication of complex systems of qubits. The price of these advantages is the problem of the environment-induced decoherence, which is typically much more serious for the macroscopic than microscopic quantum systems. Suppression of decoherence to the acceptable level requires thorough isolation of the qubit from its environment, the condition that typically limits the ability to control the qubit dynamics. This trade-off between the external control and decoherence in a qubit has a fundamental aspect related to measurement. Even in the case of a perfect set-up, the measurement necessary to observe the MQC oscillations perturbs the system by projecting its state on the eigenstates of the measured observable, and therefore presents an unavoidable source of decoherence. The intensity of this measurement-induced decoherence increases with increasing coupling strength between the detector and the oscillations, and more efficient measurement leads to stronger decoherence. First approaches to the problem of measurement of the MQC oscillations in an individual two-states system suggested for oscillations of magnetic flux in SQUIDs, considered only the conventional limit of strong or “projective” quantum measurements, in which the detector-oscillation coupling is strong. In this case, the measurement leads to rapid localization of the measured observable (flux) in one of its eigenstates, and suppresses the oscillation. This means that the time evolution of oscillations can be studied with strong measurements only if the detector can be switched on and off on the time scale shorter that the oscillations period, and only in the “ensemble” of measurements, i.e. when the experiment is repeated many times with the same initial conditions. The measurement cycle consists then of preparation of the initial state of the system followed by its free evolution and the subsequent measurement. The information about dynamics of the oscillation is contained in the probability distribution of the measurement outcomes. Since the oscillation frequency is limited from below by several factors including the decoherence rate and temperature, the need to switch the detector on and off rapidly in this approach presents at the very least a serious technical challenge. The goal of this work is to study quantitatively a new approach to measurement of the MQC oscillations that is based on weak quantum measurements , in which the dynamic interaction between the detector and the measured system is weak and does not establish perfect correlation between their states. Such a weak measurement provides only limited information about the system but, in contrast to strong measurements, perturbs the system only slightly and can be performed continuously. In this work, the process of continuous weak measurement of the MQC oscillations in an individual two-state system is considered quantitatively. Recent results for continuous measurements of electron oscillations in coupled quantum dots by a quantum point-contact are reformulated within a generic detector model and applied explicitly to the MQC oscillations of flux measured by a dc SQUID and oscillations of charge measured with a Cooper-pair electrometer. It is shown that the signal-to-noise ratio of the measurement, defined as the ratio of the amplitude of the oscillation line in the output spectrum of the detector to the background noise, is fundamentally limited by the trade-off between the acquisition of information and dephasing due to detector backaction on the oscillations. This limitation is the least restrictive for a symmetric detector, for which the signal-to-noise ratio can be expressed as $`(\mathrm{}/ϵ)^2`$, where $`ϵ`$ is the detector energy sensitivity. Since the energy sensitivity $`ϵ`$ is limited for regular (non-QND) quantum measurements by $`\mathrm{}/2`$, the signal-to-noise ratio of the continuous weak measurement of the quantum coherent oscillations is limited by 4. This limit reflects the fundamental tendency of quantum measurement to localize the system in one of the eigenstates of the measured observable. As a spin-off, the established relation between the signal-to-noise ratio and energy sensitivity is used to demonstrate that the quantum point contact is the quantum-limited detector with energy sensitivity $`ϵ=\mathrm{}/2`$. ## II Continuos measurement of the MQC oscillations with a linear detector We consider the MQC oscillations in an individual two-states system, with the two basis states separated in energy by $`\epsilon `$ and coupled by the tunneling amplitude $`\mathrm{\Delta }/2`$. The basis states are chosen to coincide with the eigenstates of an oscillating variable $`x`$, for instance, magnetic flux in a SQUID loop. In this basis, $`x=x_0\sigma _z/2`$, where $`x_0`$ is the difference between the values of $`x`$ in the two states of the system, and $`\sigma _z`$ is the Pauli matrix. In the simplest measurement scheme considered in this work, the detector measures directly the oscillating variable $`x`$, and therefore is coupled to $`x`$. This means that the Hamiltonian describing the measurement set-up (Fig. 1) is: $$H=\frac{1}{2}(\epsilon \sigma _z+\mathrm{\Delta }\sigma _x+\sigma _zf)+H_0,$$ (1) where $`H_0`$ is the Hamiltonian of the detector, and $`f`$ is the detector operator that couples it to the oscillations. For convenience, the amplitude $`x_0`$ of the oscillations and the coupling strength are included in $`f`$. We assume that the characteristic response time of the detector is short, e.g., much shorter than the period of measured oscillations, and that the detector operates in the linear regime. These two assumptions mean that the detector output $`o(t)`$ can be written as $$o(t)=q(t)+\frac{\lambda }{2}\sigma _z(t),$$ (2) where $`q(t)`$ is the noise part of the output, and $`\lambda `$ is a response coefficient. The linearity assumption implies that the coupling $`\sigma _zf/2`$ of the detector to the oscillations is sufficiently weak so that the detector response can be described in the linear-response approximation: $$\lambda =i_0^{\mathrm{}}𝑑\tau e^{i\omega \tau }[q(\tau ),f]_0.$$ (3) The average $`\mathrm{}_0`$ in eq. (3) is taken over the stationary density matrix of the detector. Since the response coefficient $`\lambda `$ is independent of frequency $`\omega `$, eq. (3) can be written in terms of the $`q`$$`f`$ correlator as $$q(t+\tau )f(t)_0=2\pi S_{qf}\delta (\tau 0),S_{qf}\frac{ai\lambda }{4\pi },$$ (4) where the infinitesimal shift in the argument of the $`\delta `$-function represents small but finite response time of the detector, and is needed to resolve the ambiguity in eq. (3). Parameter $`a`$ in eq. (4) is introduced to represent the real part of the $`q`$$`f`$ correlator that is not determined by the response coefficient $`\lambda `$. The role of detector in a quantum measurement is to convert the quantum input signal, in our case, the oscillations $`x(t)=x_0\sigma _z(t)/2`$, into the output signal that is already classical and can be dealt with (e.g., monitored or recordered) without “fundamental” problems. Condition of the classical behavior of the detector output requires the output spectral density to be much larger than the spectral density of the zero-point fluctuations in the relevant range of low frequencies of the input signal. For a detector with a short internal time scale this means that the noise $`q(t)`$ is $`\delta `$-correlated on the time scale of the input signal: $$q(t+\tau )q(t)_0=2\pi S_q\delta (\tau ).$$ (5) Here $`S_q`$ is the constant low-frequency part of the spectral density $`S_q(\omega )`$ of the detector output noise. For a quantum-limited detector at small temperature $`T0`$, eqs. (4) and (5) impose a constraint on the spectral density $`S_f`$ of the coupling operator $`f`$. Indeed, eqs. (4) and (5) can be written explicitly in the basis of the energy eigenstates $`|k`$ of the detector and give the following expression for the spectral density $`S_q`$: $$S_q=𝑑\epsilon _k𝑑\epsilon _k^{}\nu (\epsilon _k)\nu (\epsilon _k^{})\rho _kk|q|k^{}k^{}|q|k\delta (\epsilon _k\epsilon _k^{}\omega ).$$ (6) Expression for the correlation amplitude $`S_{qf}`$ is similar, with the matrix element $`k^{}|q|k`$ replaced by the matrix element of $`f`$. In these expressions, $`\epsilon _k`$ is the energy of the state $`|k`$, $`\rho _k`$ is the probability to be in this state, and $`\nu `$ is the state energy density. Since $`S_q`$ is independent of the frequency $`\omega `$, eq. (6) is satisfied when the matrix elements of $`q`$ and the density of states are constant, and eq. (6) can then be written as $`S_q=|q|^2\nu ^2.`$ It should be noted that the constant matrix elements $`q`$ and $`f`$ are off-diagonal in the $`k`$-basis and can be imaginary. Following the same steps for $`S_f`$ we express this spectral density in terms of $`S_q`$ and $`S_{qf}`$: $$S_f=|f|^2\nu ^2=|S_{qf}|^2/S_q.$$ (7) Equations (7) and (4) relate the backaction noise of the detector determined by the spectral density $`S_f`$ to its response coefficient and the output noise. Conceptually, such a relation resembles the fluctuation-dissipation theorem that links response of the system to its equilibrium fluctuations, but it does not have the status of a “theorem”. It is obvious from the derivation above that eq. (7) is not necessarily valid for an arbitrary system playing the role of detector in a quantum measurement. Nevertheless, it holds for several of the “standard” detectors: quantum point contact, resistively-shunted dc SQUID, and Cooper-pair electrometer, considered later in this work. Making use of $`q`$– and $`f`$– correlators, we can calculate the spectral density of the detector output $`o(t)`$ in the process of continuos measurement of the MQC oscillations. From eq. (2), the correlation function of $`o(t)`$ is: $$K_o(\tau )=2\pi S_q\delta (\tau )+\frac{\lambda ^2}{4}\text{Tr}\{\rho \sigma _z\sigma _z(\tau )\},$$ (8) where $`\rho `$ is the stationary density matrix of the two-state system established as a result of the interaction with the detector. Averaging the Heisenberg equation of motion of the operator $`\sigma _z(\tau )`$ over the $`\delta `$-correlated backaction noise $`f`$ of the detector, we get the set of equations for the time evolution of the matrix elements $`\sigma _{ij}`$ of $`\sigma _z(\tau )`$: $$\dot{\sigma }_{11}=\mathrm{\Delta }\text{Im}\sigma _{12},\dot{\sigma }_{12}=(i\epsilon \mathrm{\Gamma })\sigma _{12}i\mathrm{\Delta }\sigma _{11},$$ (9) and $`\sigma _{22}=\sigma _{11}`$, with the rate $$\mathrm{\Gamma }=\pi S_f$$ (10) describing the backaction dephasing of the oscillations by the detector. The density matrix $`\rho `$ of the two-state system satisfies the same set of equations (9), except for the normalization, $`\rho _{11}+\rho _{22}=1`$, and its stationary value is $`\rho =1/2`$. Solving eqs. (9) with the initial condition $`\sigma _z(0)=\sigma _z`$ and averaging $`\sigma _z\sigma _z(\tau )`$ over $`\rho =1/2`$ we find the spectral density $`S_o(\omega )=(1/2\pi )_{\mathrm{}}^{\mathrm{}}K_o(\tau )e^{i\omega \tau }𝑑\tau `$. Under the conditions of “resonance”, $`\epsilon =0`$, when the oscillation amplitude is maximum, we get: $$S_o(\omega )=S_q+\frac{\mathrm{\Gamma }\lambda ^2}{4\pi }\frac{\mathrm{\Delta }^2}{(\omega ^2\mathrm{\Delta }^2)^2+\mathrm{\Gamma }^2\omega ^2}.$$ (11) When $`\epsilon 0`$, it is convenient to calculate the spectrum numerically from eq. (9). The spectrum in this case is plotted in Fig. 2 for several values of $`\epsilon `$ and the dephasing rate $`\mathrm{\Gamma }`$. For weak dephasing, $`\mathrm{\Gamma }\mathrm{\Delta }`$, the spectrum consists of a zero-frequency Lorentzian that vanishes at $`\epsilon =0`$ and grows with increasing $`|\epsilon |`$, and a peak at the oscillation frequency $`\mathrm{\Omega }=(\mathrm{\Delta }^2+\epsilon ^2)^{1/2}`$. The peak at zero frequency reflects the incoherent transitions with a small rate of order $`\mathrm{\Gamma }`$ between the states of the two-state system. The high-frequency peak of the MQC oscillations also has the width $`\mathrm{\Gamma }`$. While this width can be small for sufficiently weak dot-contact coupling, the height of the oscillation peak cannot be arbitrarily large in comparison to the background noise spectral density $`S_q`$. At $`\epsilon =0`$, when the amplitude of the oscillations is maximum, the peak height is $`S_{max}=\lambda ^2/4\pi \mathrm{\Gamma }`$. Even in this case, the ratio of the peak height to the background is limited: $$\frac{S_{max}}{S_q}=\frac{\lambda ^2}{4\pi ^2S_fS_q}=\frac{4\lambda ^2}{\lambda ^2+a^2}4.$$ (12) This limitation is universal, e.g., independent of the coupling strength between the detector and oscillations, and reflects quantitatively the interplay between measurement of the MQC oscillations and their backaction dephasing. The fact that the height of the spectral line of the oscillations can not be much larger than the noise background means that, in the time domain, the oscillations are drowned in the shot noise. When the backaction dephasing rate $`\mathrm{\Gamma }`$ increases, the oscillation line broadens towards the lower frequencies, and eventually turns into the growing spectral peak at zero frequency associated with the incoherent jumps between the two basis states of the two-state system. At large $`\mathrm{\Gamma }`$, when the coherent oscillations are suppressed, the rate of incoherent tunneling decreases with increasing $`\mathrm{\Gamma }`$. For instance, at $`\mathrm{\Gamma }\mathrm{\Omega }`$, the tunneling rate is $`\gamma =\mathrm{\Delta }^2/2\mathrm{\Gamma }`$ , and the spectral density of the detector response at low frequencies $`\omega \gamma `$ has the standard Lorentzian form, $`S_o(\omega )S_q=2\gamma \lambda ^2/4\pi (4\gamma ^2+\omega ^2)`$. Suppression of the tunneling rate $`\gamma `$ with increasing dephasing rate $`\mathrm{\Gamma }`$ is an example of the generic “Quantum Zeno Effect” in which quantum measurement suppresses the decay rate of a metastable state. In the context of search for the macroscopic quantum coherent oscillations, the Lorentzian spectral density has been observed and used for measuring the tunneling rate of incoherent quantum flux tunneling in SQUIDs . The maximum signal-to-noise ratio $`S_{max}/S_q`$ (12) is attained if the fundamental backaction of the detector is the only mechanism of dephasing of the coherent oscillations. We now discuss briefly the effect of a weak additional dephasing and energy-relaxation on the spectral density of the oscillations. The efect of such a weak relaxation is noticeable if the backaction dephasing is also weak, $`\mathrm{\Gamma }\mathrm{\Delta }`$. Energy relaxation arises typically due to interaction with some external system (“reservoir”) that is in equilibrium at temperature $`T`$. The interaction term in the Hamiltonian can be written similarly to the interaction with the detector (1) as $$H_c=\sigma _zf_r,$$ (13) where $`f_r`$ is the reservoir force acting on the system. Under the assumption of the frequency-independent relaxation rate, the standard free equilibrium correlator of this force is (see, e.g., ): $$f_r(t)f_r(t+\tau )=\alpha \frac{d\omega }{\pi }\frac{\omega e^{i\omega \tau }}{1e^{\omega /T}},$$ (14) where parameter $`\alpha `$ characterizes the relaxation strength. Comparison of this correlator with the $`\delta `$-correlated backaction noise of the detector shows that the detector (1) is acting effectively as a reservoir with temperature that is much larger that the energies of the two-state system. Energy relaxation makes the stationary average value of $`\sigma _z`$ non-vanishing, and the output correlation function should now be calculated as $$K_o(\tau )=K_q(\tau )+\frac{\lambda ^2}{8}[\sigma _z\sigma _z(\tau )+\sigma _z(\tau )\sigma _z2\sigma _z^2].$$ (15) For weak coupling, it is convenient to find the time evolution of $`\sigma _z(\tau )`$ in the basis of eigenstates of the two-state Hamiltonian. In this basis, the Hamiltonian including interaction with the reservoir (and omitting temporarily the detector) is : $$H=\frac{1}{2}\mathrm{\Omega }\sigma _z\frac{1}{\mathrm{\Omega }}(\epsilon \sigma _z\mathrm{\Delta }\sigma _x)f_r+H_r.$$ (16) $`H_r`$ is the Hamiltonian of the reservoir. Heisenberg equations of motion that follow from the Hamiltonian (16) are: $`\dot{f}_r=i[H_r,f_r],\dot{H}_r={\displaystyle \frac{i}{\mathrm{\Omega }}}(\mathrm{\Delta }\sigma _x\epsilon \sigma _z)[f_r,H_r],`$ $`\dot{\sigma }_z=2f_r\mathrm{\Delta }\sigma _y,\dot{\sigma }_\pm =i(\mathrm{\Omega }+2\epsilon f_r/\mathrm{\Omega })\sigma _\pm i\mathrm{\Delta }\sigma _z/\mathrm{\Omega },`$ where $`\sigma _\pm (\sigma _x\pm i\sigma _y)/2`$. Integration of the first two of these equations gives in the first order in coupling (13) to the two-state system: $`f_r(t)=f_r^{(0)}(t)i\mathrm{\Delta }{\displaystyle ^t}𝑑\tau {\displaystyle ^\tau }𝑑\tau ^{}[f_r^{(0)}(\tau ),f_r^{(0)}(\tau ^{})]\sigma _y(\tau ^{}),`$ where $`f_r^{(0)}`$ is the free part of the fluctuating reservoir force in absence of coupling. Solving the second pair of the Heisenberg equations up to the second order in coupling, making the rotating-wave approximation, and tracing out the reservoir degrees of freedom with the help of the correlator (14), we get a set of equations for the evolution of the matrix elements $`s_{ij}`$ of the operator of the oscillating variable (given by $`\sigma _z(\tau )`$ in the original “position” basis) in the eigenstate basis: $`\dot{s}_{jj}(\tau )=\mathrm{\Gamma }_e[{\displaystyle \frac{\epsilon }{\mathrm{\Omega }}}\mathrm{coth}\{{\displaystyle \frac{\mathrm{\Omega }}{2T}}\}s_{jj}]+(1)^j{\displaystyle \frac{\mathrm{\Gamma }\mathrm{\Delta }^2}{2\mathrm{\Omega }^2}}(s_{11}s_{22}),`$ $$\dot{s}_{12}(\tau )=(i\epsilon \mathrm{\Gamma }_0)s_{12}.$$ (17) Initial conditions for these equations are: $`s_{11}=s_{22}=\epsilon /\mathrm{\Omega }`$, and $`s_{12}=\mathrm{\Delta }/\mathrm{\Omega }`$. The characteristic energy-relaxation rate in eq. (17) is $`\mathrm{\Gamma }_e=2\alpha \mathrm{\Delta }^2/\mathrm{\Omega }`$, and the total dephasing rate is $`\mathrm{\Gamma }_0={\displaystyle \frac{1}{\mathrm{\Omega }^2}}[\alpha (\mathrm{\Delta }^2\mathrm{\Omega }\mathrm{coth}\{{\displaystyle \frac{\mathrm{\Omega }}{2T}}\}+4\epsilon ^2T)+\mathrm{\Gamma }(\epsilon ^2+{\displaystyle \frac{\mathrm{\Delta }^2}{2}})].`$ In eqs. (17), we also added the detector dephasing terms from eq. (9) “rotated” from the position basis into the eigenstate basis. The density matrix $`r`$ of the two-state system in the basis of eigenstates satisfies similar equations, and the stationary values of its matrix elements are $`r_{12}=0`$ and $`r_{11}=(\mathrm{\Gamma }_t+\mathrm{\Gamma }_e)/2\mathrm{\Gamma }_t`$, where $`\mathrm{\Gamma }_t\mathrm{\Gamma }_e\mathrm{coth}(\mathrm{\Omega }/2T)+\mathrm{\Gamma }\mathrm{\Delta }^2/\mathrm{\Omega }^2.`$ Using these relations, the definition (15), and the evolution equations (17) we find the spectral density: $`S_o(\omega )=S_q+{\displaystyle \frac{\lambda ^2}{4\pi \mathrm{\Omega }^2}}\times `$ (19) $`\left([1({\displaystyle \frac{\mathrm{\Gamma }_e}{\mathrm{\Gamma }_t}})^2]{\displaystyle \frac{2\epsilon ^2\mathrm{\Gamma }_t}{\omega ^2+\mathrm{\Gamma }_t^2}}+{\displaystyle \underset{\pm }{}}{\displaystyle \frac{\mathrm{\Delta }^2\mathrm{\Gamma }_0}{(\omega \pm \mathrm{\Omega })^2+\mathrm{\Gamma }_0^2}}\right).`$ As in the case without energy relaxation, the spectral density consists of a zero-frequency Lorentzian of width $`\mathrm{\Gamma }_t`$ and peaks at $`\pm \mathrm{\Omega }`$ of width $`\mathrm{\Gamma }_0`$ due to coherent oscillations. For weak relaxation, the incoherent slow transitions giving rise to the low-frequency noise are the transitions between the two energy eigenstates of the system. The height of the oscillation peak is suppressed in presence of additional energy relaxation that contributes to the dephasing rate $`\mathrm{\Gamma }_e`$, and the relative magnitude of the peak, $`S_{max}/S_q`$ is smaller than its value without the relaxation. ## III Relation to energy sensitivity The detector characteristics for measurement of the quantum coherent oscillations in a two-state system considered in the previous Section are related to another detector characteristic that is used for measurements of the harmonic signals – see, e.g., examples in , and sometimes is loosely referred to as “energy sensitivity”. It is defined by considering the detector measuring a harmonic oscillator with a frequency $`\omega _0`$ and a small relaxation rate $`\gamma \omega _0`$. In this Section, we establish the quantitative relation between the signal-to-noise ratio $`S_{max}/S_q`$ for measurements of the two-state systems discussed above with the energy sensitivity used in the literature for measurements of harmonic signals. The Hamiltonian of the damped harmonic oscillator attached to a detector is obtained by replacing the two-state part of eqs. (1) and (13) with the corresponding oscillator terms: $$H=\frac{M}{2}(\dot{x}^2+\omega _0^2x^2)x(f+f_r)+H_0+H_r,$$ (20) where $`M`$ is the mass of the oscillator and $`x`$ is the oscillating coordinate. Due to linearity of the system, the Heisenberg equation of motion for $`x`$ that follow from the Hamiltonian (20) (see, e.g., ) coincides with the classical equation of motion of the damped oscillator: $$\ddot{x}\gamma \dot{x}+\omega _0^2x=\frac{1}{M}(f_r(t)+f(t)).$$ (21) As in the previous Section, the random forces $`f_r(t)`$ and $`f(t)`$ are produced, respectively, by the reservoir responsible for the energy relaxation of the oscillator and by the detector. (Although the operators $`f_r`$ and $`f`$ in eqs. (20) and (21), as well as the transfer coefficient $`\lambda `$ in eq. (22) below, differ from the corresponding quantities used in Section 2 by a normalization factor, this distinction is not made explicit. This should not lead to any confusion, since the normalization factor drops out of all expression which are compared between the two Sections.) In general, the right-hand-side of eq. (21) should also contain an external perturbation that creates a “signal” component of the oscillations $`x(t)`$. However, for the discussion of the detector sensitivity, it is appropriate to treat the equilibrium fluctuations of the oscillator driven by the reservoir noise $`f_r(t)`$ (e.g., the zero-point oscillations at vanishing temperature $`T`$) as part of the signal. This allows us not to include additional signal terms in eq. (21). The sensitivity of the detector is characterized by the detector noise contribution to the spectral density of the output $`S_o`$ reduced to the detector input. Similarly to eq. (2) the output of the detector measuring harmonic oscillator is: $$o(t)=q(t)+\lambda x(t).$$ (22) This equation implies that the detector contribution to the output noise comes from the two sources: direct output noise $`q(t)`$, and effect of the backaction noise $`f(t)`$ on the oscillator coordinate $`x(t)`$. Introducing the dynamic response function $`G(t)`$ of the oscillator: $$x(t)=_0^+\mathrm{}𝑑\tau G(\tau )(f_r(t\tau )+f(t\tau )),$$ (23) with $`G(\tau )=0`$ for $`\tau <0`$, we see from eq. (21) that $$G(\omega )=𝑑\tau e^{i\omega \tau }G(\tau )=\frac{1}{M}\frac{1}{\omega _0^2\omega ^2+i\gamma \omega }.$$ (24) In terms of the response function, $`x(\omega )=G(\omega )(f_r(\omega )+f(\omega ))`$. Since the detector noise $`f(t)`$ is uncorrelated with the reservoir force $`f_r(t)`$, or in general, any other signal component of $`x(t)`$, we see from eqs. (22), (23), and (24) that the spectral density of the detector output consists of the two additive components: signal, i.e., equilibrium spectral density of the harmonic oscillations transformed to the output, and the detector noise $`S_N(\omega )`$: $`S_o(\omega )=\lambda ^2|G(\omega )|^2S_r(\omega )+S_N(\omega ),`$ $$S_N(\omega )=S_q+\lambda ^2|G(\omega )|^2S_f+2\lambda \text{Re}G(\omega )\overline{S}_{qf}.$$ (25) Here $`\overline{S}_{qf}`$ is the symmetrized correlator of the detector output and input noises, $`\overline{S}_{qf}=(S_{qf}+S_{fq})/2=a/4\pi `$ (with $`a`$ defined in eq. (4)), and $`S_r(\omega )`$ is the spectral density of the reservoir force $`f_r`$: $$S_r(\omega )=(\gamma \omega M/2\pi )\mathrm{coth}(\mathrm{}\omega /2T).$$ (26) The noise properties of the detector are better characterized if its contribution to the output noise is reduced to the input, i.e. instead of $`S_N(\omega )`$ (25) we consider the quantity $`F(\omega )S_N(\omega )/\lambda ^2|G(\omega )|^2`$. For weak damping, $`\gamma \omega _0`$, when $`|G(\omega )|^2{\displaystyle \frac{1}{4\omega _0^2M^2((\omega \omega _0)^2+\gamma ^2/4)}},`$ we get $`F(\omega )=S_f+S_q(2\omega _0M/\lambda )^2((\omega \omega _0)^2+\gamma ^2/4)`$ $`\overline{S}_{qf}(4\omega _0M/\lambda )(\omega \omega _0).`$ The three terms in this equation scale differently with the strength of the detector-oscillator coupling, since $`\lambda `$ and $`\overline{S}_{qf}`$ are proportional to the first power, while $`S_f`$ is proportional to the second power of the coupling strength. $`F(\omega )`$ can be minimized with respect to the coupling strength and also with respect to the small detuning $`\omega \omega _0`$ between the signal frequency and the oscillator frequency. The minimum is reached when $`\omega \omega _0={\displaystyle \frac{\lambda \overline{S}_{qf}}{2\omega _0MS_q}},`$ and $`\lambda ={\displaystyle \frac{\gamma \omega _0MS_q}{S_0}},S_0(S_qS_f\overline{S}_{qf}^2)^{1/2},`$ and is equal to $$F_{min}=2\gamma \omega _0M\frac{S_0}{\lambda }.$$ (27) It is convenient to normalize the reduced noise $`F`$ in such a way that it can be directly compared to the equilibrium fluctuations in the oscillator driven by $`S_r`$. Since at this stage we can already neglect small difference between the oscillator frequency and signal frequency, both the minimum noise $`F_{min}`$ (27) and the equilibrium spectral density $`S_r`$ (26) are proportional to the oscillator parameters $`\gamma \omega _0M`$. This means that with appropriate normalization we can define $`F_{min}`$ directly in terms of the number of quanta added to the signal at frequency $`\omega `$ by the detector. This is achieved by introducing the energy sensitivity as $$ϵ\frac{\pi }{\gamma \omega _0M}F_{min}=\frac{2\pi }{\lambda }(S_qS_f\overline{S}_{qf}^2)^{1/2}.$$ (28) Equations (7) and (4) of the previous Section show that for a quantum-limited detector $$ϵ=\mathrm{}/2.$$ (29) (Note that in this Chapter, $`\mathrm{}`$ is shown only in some of the final results.) Equation (29) agrees with the conclusion of a general theory of quantum linear amplifiers, according to which a phase-insensitive linear amplifier adds at least half-a-quantum of noise to the amplified signal – see, e.g., . This result is related to eq. (ref39) since detector in the measurement process plays the role of an amplifier transforming weak quantum input signal into the classical output. Comparing eq. (28) for the energy sensitivity with eq. (12) for the signal-to-noise ratio of the measurement of the quantum coherent oscillations we see finally that these two quantities are closely related. When the output and input noises of the detector are uncorrelated, $`\overline{S}_{qf}=0`$, the relation is simple: $$\frac{S_{max}}{S_q}=(\mathrm{}/ϵ)^2.$$ (30) As will be clear from the examples of specific detector considered below, the situation with $`\overline{S}_{qf}=0`$ can be reasonably referred to as “symmetric detector”. As follows from eq. (ref40), the largest signal-to-noise ratio of 4 is obtained for such a symmetric detector in the quantum-limited regime with $`\epsilon =\mathrm{}/2`$. When the input-output correlation is non-vanishing, the signal-to-noise ratio is smaller that the value given by eq. (30), while the energy sensitivity $`ϵ`$ for measurement of the harmonic signal can still be made equal to $`\mathrm{}/2`$ by optimizing the detuning between the signal and the oscillator. Another difference between the measurement of harmonic oscillator defining $`ϵ`$ and measurement of the two-state system, is that the minimum noise (27) in the oscillator measurement is reached only for optimum detector-oscillator coupling, while the maximum signal-to-noise ratio (12) for the quantum oscillation measurement is independent of the coupling strength to the detector as long as the coupling is weak. ## IV Energy sensitivity of a quantum point contact One of the applications of the results obtained in the previous Section is the demonstration that a quantum point contact that is frequently used as the detector of electric charge or voltage can reach the quantum-limited regime with ultimate energy sensitivity (29). The mechanism of operation of a quantum point contact as a detector utilizes modifications of the electron transmission properties of the contact by the measured voltage . When the contact is biased with a large voltage $`V`$, changes in the electron transmission probability lead to changes in the current $`I`$ flowing through the contact which serve as the measurement output. Fluctuations of the electric potential in the contact region due to the current flow produce the backaction dephasing of the measured object by the point contact . This dephasing was calculated for symmetric contacts within different approaches in . It is known that in the case of measurement of a two-state system, when the coupling to the system is symmetric, quantum point contact is an ideal detector of the quantum coherent oscillations. Such a detector causes the minimum dephasing of the oscillations that is consistent with the information acquisition by the measurement. For asymmetric coupling, the dephasing by the point contact is larger that the fundamental minimum . To calculate the energy sensitivity of the point contact detector, we start with the standard Hamiltonian of a single-mode point contact. Including a weak additional scattering potential $`U(x)`$ for the point contact electrons which is the input signal of the measurement we can write the Hamiltonian as $$H=\underset{ik}{}\epsilon _ka_{ik}^{}a_{ik}+U,U=\underset{ij}{}U_{ij}\underset{kp}{}a_{ik}^{}a_{jp}.$$ (31) The operators $`a_{ik}`$ in this Hamiltonian represent point-contact electrons in the two scattering states $`i=1,2`$ (incident from the two contact electrodes) with momentum $`k`$, and $`U_{ij}=𝑑x\psi _i^{}(x)U(x)\psi _j(x)`$ are the matrix elements of the potential $`U(x)`$ in the basis of the scattering states. Here $`\psi _i(x)`$ is the wavefunction of the scattering state, and $`x`$ is the coordinate along the point contact. Several assumptions are made about the contact. The bias energy $`eV`$ is assumed to be much larger than temperature $`T`$, but much smaller than both the Fermi energy in the point contact and the inverse traversal time of the contact. This allows us to linearize the energy spectrum of the point-contact electrons: $`\epsilon _k=v_Fk`$, where $`v_F`$ is the Fermi velocity, and neglect the momentum dependence of the matrix elements $`U_{ij}`$. The potential $`U(x)`$ is also assumed to be sufficiently weak and can be treated as perturbation. In this regime, the point contact operates as a linear detector, and the current response to the perturbation $`U`$ can be calculated in the linear-response approximation. The last assumption is that the frequencies of the input signal are much smaller than $`eV`$, the fact that allows to treat $`U`$ as the static perturbation. At frequencies much lower that both $`eV`$ and inverse traversal time of the contact, the current is constant throughout the contact and the contact response can be calculated at any point $`x`$. We choose the origin of the coordinate $`x`$ in such a way that the unperturbed scattering potential is effectively symmetric, i.e., the reflection amplitudes for both scattering states are the same, and then take $`x`$ to lie in the asymptotic region of the scattering states. In this case, the standard expression for the current in terms of the electron operators $`\mathrm{\Psi }(x)`$: $`I={\displaystyle \frac{ie\mathrm{}}{2m}}(\mathrm{\Psi }^{}{\displaystyle \frac{\mathrm{\Psi }}{x}}{\displaystyle \frac{\mathrm{\Psi }^{}}{x}}\mathrm{\Psi }),\mathrm{\Psi }(x)={\displaystyle \underset{ik}{}}\psi _{ik}(x)a_{ik},`$ gives for the current operator at $`x`$: $`I={\displaystyle \frac{ev_F}{L}}{\displaystyle \underset{kp}{}}[D(a_{1k}^{}a_{1p}a_{2k}^{}a_{2p})+`$ (32) $`i(DR)^{1/2}e^{i(kp)|x|}(a_{1k}^{}a_{2p}a_{2k}^{}a_{1p})].`$ (33) Here $`D`$ and $`R=1D`$ are the transmission and reflection probabilities of the point contact, $`L`$ is a normalization length, and the variation of the momentum near the Fermi points (i.e., the difference between $`k`$ and $`p`$) was neglected everywhere besides the phase factor in the second term. The reason for keeping this factor will become clear later. In the linear-response regime, the current response of the point contact is driven by the part of the perturbation $`U`$ causing transitions between the two scattering states $`\psi _{1,2}`$. As shown in the Appendix, the real part of the transition matrix element $`U_{12}`$ is related to the change $`\delta D`$ of the transmission probability of the contact: $$U_{12}=\frac{v_F}{L}\frac{\delta D+iu}{2(DR)^{1/2}},U_{21}=U_{12}^{}.$$ (34) The imaginary part of $`U_{12}`$, expressed through a dimensionless parameter $`u`$ in eq. (4), does not affect the current $`I`$. Qualitatively, it characterizes the degree of asymmetry in the coupling of the quantum dots to the point contact; $`u=0`$ if the perturbation potential $`U(x)`$ is applied symmetrically with respect to the main scattering potential of the point contact. In the measurement process, the perturbation $`U`$ represents the coupling operator between the contact and the measured system, whereas the current $`I`$ is the measurement output. As follows from the discussion in the previous Section, the energy sensitivity of the contact is determined by the correlators of $`U`$ and $`I`$, and by the response coefficient $`\lambda `$. Using eqs. (31), (33), and (34) we can evaluate the correlators directly. In the limit of large voltages, $`eVT`$, when the contact noise properties are dominated by the shot noise, we get: $`U(t)U(t+\tau )_0={\displaystyle \frac{eV}{4\pi }}{\displaystyle \frac{(\delta D)^2+u^2}{DR}}\delta (\tau ),`$ (35) $`I(t+\tau )I(t)_0={\displaystyle \frac{e^3VDR}{\pi }}\delta (\tau ),`$ (36) $`U(t)I(t+\tau )_0={\displaystyle \frac{e^2V}{2\pi }}(i\delta D+u)\delta (\tau \eta ).`$ (37) The time delay $`\eta |x|/v_F`$ in the last of eqs. (36) comes from the phase factor $`e^{i(kp)|x|}`$ kept in eq. (33), and is infinitesimally small for small traversal time of the contact. It is nevertheless important for correct calculation of the contact response $`\lambda `$. From the $`U`$$`I`$ correlator and the standard expression for the linear response (3) we confirm that $`\lambda `$ is equal to the change of current through the contact due to change $`\delta D`$ of the transmission coefficient, $`\lambda =e^2V\delta D/\pi `$. The correlators (36) satisfy the general relations (7) and (4), and therefore the energy sensitivity of the quantum point contact reaches the fundamental quantum limit (29). It should be noted, however, that this conclusion is strictly valid only in the large-voltage limit $`eVT`$. At finite temperature $`T`$, scattering of the point-contact electrons within the same direction of propagation (described by the terms $`U_{11}`$ and $`U_{22}`$ of the perturbation $`U`$) creates additional contribution to the backaction noise and degrades the energy sensitivity. The magnitude of this effect depends on the magnitude of the “forward” scattering matrix elements $`U_{11}`$ and $`U_{22}`$ relative to the backscattering matrix element $`U_{12}`$, increasing with intensity of the forward scattering but decreasing with $`T/eV`$ ratio. ## V Flux and charge MQC oscillations This Section provides specific examples of measurements of the macroscopic quantum coherent oscillations of magnetic flux and electric charge. It is shown that the typical detectors for the flux and charge measurements, dc SQUID and Cooper-pair electrometer, satisfy the general equations of Sections 2 and 3, and should be capable of reaching the fundamental limit of the signal-to-noise ratio for the continuous weak measurement of the MQC oscillations. ### A Flux oscillations measured with a dc SQUID Typical set-up of a measurement of the MQC oscillations of flux with a dc SQUID consists of a two-state flux system (rf SQUID with half of a magnetic flux quantum $`\mathrm{\Phi }_0=\pi \mathrm{}/e`$ induced in it by an external magnetic field) coupled inductively to a dc SQUID biased with an external current $`I_0`$ and shunted by a resistor $`R`$ (Fig. 3). When the inductance of dc SQUID loop is small, the difference between the two Josephson phases $`\phi _{1,2}`$ across the two junctions of the SQUID is directly linked to the flux $`\mathrm{\Phi }`$ induced in the dc SQUID by the flux oscillations: $`\phi _1\phi _2=2\pi \mathrm{\Phi }/\mathrm{\Phi }_0\mathrm{\Theta }.`$ In this case, the SQUID is equivalent to a single Josephson junction, with the supercurrent in this junction modulated by the flux $`\mathrm{\Phi }`$. The total amplitude of Cooper-pair tunneling in the SQUID is equal to a sum of the two individual amplitudes of tunneling in the two SQUID junction, and can be written as $$E_J/2=I^{(+)}(\mathrm{\Theta })/4e,I^{(+)}(\mathrm{\Theta })=I_1e^{i\mathrm{\Theta }/2}+I_2e^{i\mathrm{\Theta }/2},$$ (38) where $`I_{1,2}`$ are the critical currents of the two junctions. Coherent sum of the two tunneling amplitudes in (38) leads to modulation of the total supercurrent of the SQUID. We consider the simplest regime of the dc SQUID dynamics when the bias current $`I_0`$ and associated average voltage $`V_0=RI_0`$ across the dc SQUID are sufficiently large, and the dc component of the Josephson current through it is small in comparison to $`I_0`$. In this regime, the Cooper-pair tunneling through the SQUID is adequately described by perturbation theory in the tunneling amplitude (38) and can be qualitatively interpreted as incoherent jumps of individual Cooper pairs. The resistor $`R`$ provides the dissipation mechanism that transforms reversible dissipationless Cooper-pair oscillations between the two electrodes of the SQUID into incoherent tunneling. Using the known results for the incoherent Cooper-pair tunneling , we can calculate the rate of this tunneling and find all the detector characteristics of the dc SQUID. The SQUID is coupled to the oscillations by the operator of the current $`I_{}`$ circulating in the SQUID loop multiplied by the change $`\delta \mathrm{\Phi }`$ of the flux in the loop induced by the flux oscillations. Changes in the flux through the dc SQUID change the total current $`I_+`$ through both SQUID junctions and create deviations $`V`$ of the voltage across the SQUID from $`V_0`$, $`V=RI_+`$, that serve as the measurement output. As follows from the general discussion in Sec. 2, SQUID parameters important for measurement are the coefficient $`\lambda `$ of the transformation of the oscillating flux into the voltage $`V`$, the spectral density $`S_I`$ of the circulating current $`I_{}`$ that is responsible for backaction dephasing by the SQUID, the spectral density of the output voltage $`S_V`$, and the correlator $`S_{VI}`$ between $`V`$ and $`I_{}`$. These parameters can be found quantitatively starting from the tunneling part $`H_T`$ of the SQUID Hamiltonian that can be written as $$H_T=\frac{E_J}{2}e^{i(2eV_0t+\phi (t))}+h.c.,$$ (39) with $`\phi (t)`$ being the random Josephson phase across the dc SQUID accumulated due to equilibrium voltage fluctuations produced by the resistor $`R`$. It is characterized by the correlator $$\phi (t)\phi =\rho \frac{d\omega }{\omega }g(\omega )\frac{e^{i\omega t}}{1e^{\omega /T}},$$ (40) where $`\rho =R/R_Q`$ is the resistance $`R`$ in units of the quantum resistance $`R_Q=\pi \mathrm{}/4e^2`$, the average $`\mathrm{}`$ is taken over equilibrium density matrix of the resistor $`R`$, and $`g(\omega )`$ describes the cut-off of the dissipation provided by $`R`$ at some large frequency $`\omega _c`$ associated with either finite inductance of the SQUID or finite capacitance of its junctions, while $`g(\omega )=1`$ at $`\omega \omega _c`$. The operators of the two currents $`I_\pm `$ that determine the SQUID parameters are: $$I_\pm =\frac{i}{2}[I^{(\pm )}(\mathrm{\Theta })e^{i(2eV_0t+\phi (t))}h.c.],$$ (41) $`I^{()}(I_1e^{i\mathrm{\Theta }/2}I_2e^{i\mathrm{\Theta }/2})/2.`$ In the regime of the incoherent Cooper-pair tunneling the average dc current $`I_+`$ can be found treating the tunneling $`H_T`$ as perturbation: $$I_+=i_0^{\mathrm{}}𝑑t[I_+,H_T(t)]=\pi |I^{(+)}(\mathrm{\Theta })|^2\tau /e,$$ (42) where $$\tau \frac{1}{4\pi }\text{Re}_0^{\mathrm{}}𝑑te^{i2eV_0t}[e^{i\phi (t)},e^{i\phi }].$$ (43) For example, for vanishing temperature $`T`$, and small bias voltages, $`2eV_0\omega _c`$, the time $`\tau `$ defined in (43) can be found from eqs. (43) and (40) to be (see, e.g., ): $$\tau =(1/4\omega _c\mathrm{\Gamma }(\rho ))(2eV_0/\omega _c)^{\rho 1}.$$ (44) When the resistance $`R`$ is small, $`RR_Q`$, $`\tau `$ becomes independent of $`\omega _c`$, $`\tau =eR/2\pi V`$. In this case, all the SQUID characteristics, including the average current $`I_+`$ (42) can be obtained by direct time averaging of the classical Josephson oscillations in the SQUID. The noise spectral densities of the two currents, $`I_\pm `$, are obtained by directly taking the average over equilibrium density matrix of the resistor $`R`$. They vary with frequency on the scale of the Josephson frequency $`2eV_0`$, and are constant at $`\omega 2eV_0`$. In this frequency range, $`S_I={\displaystyle \frac{1}{2\pi }}{\displaystyle }dtI_{}(t)I_{}={\displaystyle \frac{1}{8\pi }}|I^{()}(\mathrm{\Theta })|^2\times `$ $$𝑑te^{i2eV_0t}[e^{i\phi (t)},e^{i\phi }]_+.$$ (45) Fluctuation-dissipation theorem relates the anticommutator $`[\mathrm{}]_+`$ in this equation to the commutator in eq. (43) and gives: $$S_I=|I^{()}(\mathrm{\Theta })|^2\tau ^{},S_V=R^2|I^{(+)}(\mathrm{\Theta })|^2\tau ^{},$$ (46) where $`\tau ^{}\tau \mathrm{coth}(eV_0/T)`$. The correlation function $`S_{VI}`$ is found similarly: $$S_{VI}=R[I^{(+)}(\mathrm{\Theta })]^{}I^{()}(\mathrm{\Theta })\tau ^{}.$$ (47) Comparison of the spectral density $`S_V`$ (46) and the average current (42) shows that $`S_{I_+}=S_V/R^2=(eI_+/\pi )\mathrm{coth}(eV_0/T)`$, i.e. the noise of the current $`I_+`$ can indeed be interpreted as resulting from uncorrelated transitions of individual Cooper pairs. In particular, at $`TeV_0`$, the noise is the shot noise of Cooper pairs. Finally, eq. (42) gives the response coefficient of the SQUID $$\lambda V/\mathrm{\Phi }=2\pi R(|I^{(+)}(\mathrm{\Theta })|^2/\mathrm{\Theta })\tau .$$ (48) (Note that as in eq. (46) for the backaction noise $`S_I`$ and also in eq. (47) for the correlator $`S_{VI}`$, the factor $`\delta \mathrm{\Phi }`$ is omitted from the definition of $`\lambda `$.) For temperatures negligible on the scale of $`eV_0`$, $`\tau ^{}`$ in the noise spectral densities is equal to $`\tau `$ and eqs. (46) through (48) show that the spectral densities $`S_V`$, $`S_I`$, and $`S_{VI}`$ satisfy the general relation (7), and since $`|I^{(+)}(\mathrm{\Theta })|^2/\mathrm{\Theta }=2\text{Im}\{[I^{(+)}(\mathrm{\Theta })]^{}I^{()}(\mathrm{\Theta })\}`$, the correlator $`S_{VI}`$ is also related to the response coefficient $`\lambda `$ by the expression identical to eq. (4). Moreover, since $`[I^{(+)}(\mathrm{\Theta })]^{}I^{()}(\mathrm{\Theta })={\displaystyle \frac{1}{2}}(I_1^2I_2^2)+iI_1I_2\mathrm{sin}\mathrm{\Theta },`$ we see that the real part of the correlator $`S_{VI}`$, which increases backaction dephasing produced by the SQUID, is indeed associated with the SQUID asymmetry. When $`I_1=I_2`$, the real part vanishes and the SQUID as detector reaches quantum-limited optimum for measurement of the quantum coherent oscillations. For such a symmetric SQUID the signal-to-noise ratio of the oscillation measurement is given by eq. (30) (with the intensity of background output noise given by $`S_V`$). If the temperature $`T`$ of the symmetric SQUID is negligible, eqs. (46) through (48), and (28) show that in this regime the SQUID is the quantum-limited detector with the energy sensitivity $`ϵ=\mathrm{}/2`$, and the signal-to-noise ratio for measurement of the quantum flux oscillations is $`S_{max}/S_V=4`$. When $`T`$ becomes non-vanishing, both the output and backaction noise increase, $`\tau ^{}>\tau `$, and eq. (30) describes the gradual suppression of the signal-to-noise ratio with increasing temperature. ### B Charge oscillations measured with a Cooper-pair electrometer Coherent oscillations of charge take place in Josephson junctions which are sufficiently small for the charging energy $`E_C`$ of an individual Cooper pair, $`E_C=(2e)^2/2C`$ to be larger than temperature $`T`$ and Josephson coupling energy $`E_J`$. The supercurrent flow through the junction is “discretized” in this regime into the transfer of individual Cooper pairs by strong Coulomb repulsion. Quantitatively, if the charging energy is smaller that the superconducting energy gap $`\mathrm{\Delta }`$ so that the dissipative quasiparticle tunneling is suppressed, the junction dynamics is governed by the simple Hamiltonian: $$H=E_C(nq)^2\frac{E_J}{2}(|nn+1|+|n+1n|),$$ (49) where $`n`$ is the number of Coper-pairs charging the junction, and, here and below, $`q`$ is the charge (in units of 2$`e`$) injected into the junction from external circuit. Eigenstates of the Hamiltonian (49) form energy bands as functions of the injected charge $`q`$, which can be varied continuously. Variations of the injected charge $`q`$ within these bands leads to the possibility of controlling the tunneling of individual Cooper pairs . The best way of injecting the charge $`q`$ in a junction is provided by the “Cooper-pair box” system in which the junction is attached to external bias voltage $`V_g`$ through a capacitor. If $`q`$ is fixed at half of a Cooper-pair charge, $`q1/2`$, and the tunneling amplitude $`E_J/2`$ is much less than $`E_C`$, the two states of the Hamiltonian (49): $`n=0`$ and $`n=1`$ are nearly-degenerate and separated by the large energy gaps from all other states. In this regime, the junction dynamics is equivalent to that of a regular quantum two-state system with the two basis state that correspond to a Coooper pair being on the left or on the right electrode of the junction. Coherent superposition of charge states in such a two-state system is observed indirectly by measuring either the width of the transition region between the two charge states or the energy gap between the eigenstates as functions of the induced charge $`q`$. Quantum coherent oscillations between the two charge states were also observed directly in the time-dependent measurement . The experiment was effectively based on the strong measurement of charge oscillations, when each measurement suppresses the oscillations, and they are observed as oscillations of probability in an ensemble of measurements. Continuous weak measurement of the quantum-coherent charge oscillations similar to the measurement of the flux oscillations with a dc SQUID discussed above would provide a less intrusive way of studying these oscillations. One of the detectors appropriate for such a measurement is a Cooper-pair electrometer : two small Josephson junctions with Josephson coupling energies $`E_{1,2}`$ and capacitances $`C_{1,2}`$ connected in series and shunted with a resistor $`R`$ (Fig. 4). As in the case of dc SQUID, we consider dynamics of the Cooper pair transfer through the electrometer in the regime of incoherent tunneling. The main difference with the SQUID case is that now the amplitude of Cooper pair tunneling is modulated through the modulation of energy of the intermediate state in the process of the two-step transfer of Cooper pairs in the two junctions. At small bias voltages $`V_0E_C/e`$, where from now on $`E_C=2e^2/(C_1+C_2)`$ is the charging energy of the central electrode of the electrometer, the intermediate state is virtual, and the average current $`I`$ through the electrometer is determined by the same eq. (42) with $`I^{(+)}(\mathrm{\Theta })`$ replaced with the amplitude $`I(q)`$ of Cooper-pair transfer through both junctions (defined below). The charge $`q`$ injected into the central electrode controls the energy of the intermediate states in the process of the Cooper-pair transfer and modulates the tunneling amplitude. At small Josephson coupling energies $`E_{1,2}E_C`$, and away from the resonance points $`q=\pm 1/2`$, Cooper-pair tunneling can be treated as perturbation. Then the instantaneous value of the current $`I`$ for a fixed value of the Josephson phase $`\phi `$ across the electrometer (see, e.g., description of the two-junction system in ) is: $$I=I(q)\mathrm{sin}(2eV_0t+\phi (t)),$$ (50) $`I(q){\displaystyle \frac{eE_1E_2}{E_C}}({\displaystyle \frac{1}{12q}}+{\displaystyle \frac{1}{1+2q}}).`$ The two terms in the second equation in (50) correspond to the two intermediate states with different charges $`n=\pm 1`$ on the central electrode of the electrometer in the Cooper-pair transfer process. Averaging eq. (50) over the equilibrium quantum fluctuations of $`\phi `$ we get expression for the average value of the current $`I`$ that is equivalent to eq. (42). Since the tunneling current $`I`$ through the electrometer is the measurement output, this expression determines the response coefficient $`\lambda `$ of the electrometer: $$\lambda (I/q)=\pi ([I(q)]^2/q)\tau /e,$$ (51) where $`\tau `$ is given by eqs. (43) and (44). Similarly, the output noise of the current $`I`$ can be obtained as $$S_I=[I(q)]^2\tau ^{}.$$ (52) The backaction noise of the Cooper-pair electrometer is created by fluctuations of the charge on its central electrode in the process of the Cooper-pair tunneling. This fluctuations lead to fluctuations of electric potential of this electrode. In the same regime as for eq. (50), the magnitude of this fluctuations is determined by the magnitude of the instantaneous value of the potential at a fixed Josephson phase difference $`\phi `$ across the electrometer: $$V=U(q)\mathrm{cos}(2eV_0t+\phi (t)),$$ (53) where $`U(q){\displaystyle \frac{E_1E_2}{2eE_C}}({\displaystyle \frac{1}{(12q)^2}}{\displaystyle \frac{1}{(1+2q)^2}}).`$ As before, averaging over the fluctuations of $`\phi (t)`$ we get the spectral density of the backaction noise and its correlation with the output noise: $$S_V=[U(q)]^2\tau ^{},S_{VI}=iU(q)I(q)\tau ^{}.$$ (54) Equations (51), (52), and (54) show that at vanishing temperature, when $`\tau ^{}=\tau `$, the noise characteristics of the Cooper-pair electrometer satisfy the general relations (7) and (4) of a quantum-limited detector. Moreover, the electrometer is “symmetric” detector in a sense that the input-output correlator $`S_{VI}`$ is purely imaginary. This means that both its energy sensitivity $`ϵ`$ and the signal-to-noise ratio (30) for measurement of the two-state system reach fundamental limits. In summary, the examples of specific detectors considered in this Section show explicitly that many standard detectors should be capable of reaching the fundamental limits of sensitivity for measurements of electric charge and magnetic flux. In this regime, they are characterized by the fundamental signal-to-noise ration of 4 for continuous weak measurement of the macroscopic quantum coherent oscillations. The limitation on the signal-to-noise ratio of such a measurement has the same origin as the quantum limitation on the operation of an ideal linear phase-insensitive amplifier that adds a minimum of half-a-quantum of noise to the amplified signal. The author would like to acknowledge discussions with M.H. Devoret, J.R. Friedman, A.N. Korotkov, K.K. Likharev, J.E. Lukens, Yu.V. Nazarov, R.J. Schoelkopf, G. Schön, and A.B. Zorin. This work was supported in part by AFOSR. Appendix In this Appendix, we derive eq. (34) that relates the matrix elements of the perturbation of the scattering potential to the transmission properties of a point contact. Relation (34) can be established considering the stationary states of an electron confined to move on the interval $`x[L/2,L/2]`$ with the main scattering potential located at the center of the interval, $`x0`$. The scattering matrix $`S`$ for the symmetric scattering potential can be written as $`S=e^{i\mathrm{\Theta }}\left(\begin{array}{cc}i\mathrm{sin}\nu ,& \mathrm{cos}\nu \\ \mathrm{cos}\nu ,& i\mathrm{sin}\nu \end{array}\right),`$ where $`\mathrm{\Theta }`$ is the phase of the transmission amplitude, and $`\nu `$ parametrizes transmission probability $`D`$: $$D=\mathrm{cos}^2\nu .$$ (55) Diagonalizing the scattering matrix $`S`$, we find the two phase shifts $`\phi _{1,2}`$ associated with it: $`\phi _1=\mathrm{\Theta }+\nu `$, $`\phi _2=\mathrm{\Theta }+\pi \nu `$. The variations $`\delta \phi _j`$ of the two phase shifts due to perturbation of the scattering potential lead to changes $`\delta \epsilon _j`$ in energies of the two stationary states, $$\delta \epsilon _j=v_F\delta \phi _j/L$$ (56) For symmetric main scattering potential, the two stationary states $`\chi _j(x)`$ are given by the even and odd combinations of the scattering states $`\psi _j(x)`$. In the basis of the states $`\chi _j`$, the perturbation matrix $`U`$ introduced in eq. (31) is: $$U=\left(\begin{array}{cc}(U_{11}+U_{22})/2+\text{Re}U_{12},& (U_{11}U_{22})/2i\text{Im}U_{12}\\ (U_{11}U_{22})/2+i\text{Im}U_{12},& (U_{11}+U_{22})/2\text{Re}U_{12}\end{array}\right).$$ (57) The diagonal elements of this matrix give the first-order corrections to the energies of the stationary states. Comparing expressions for the energy corrections given by eq. (57) to expressions for the phase shifts combined with eq. (56), we see that $`\text{Re}U_{12}=v_F\delta \nu /L.`$ This equation, together with the relation (55) between $`\nu `$ and transmission probability $`D`$, gives eq. (34) of the main text. Since the matrix (57) should be symmetric for the perturbation of the scattering potential that is symmetric with respect to the main part of the potential, we also see that $`\text{Im}U_{12}=0`$ in the symmetric case. This means that the nonvanishing imaginary part of $`U_{12}`$ can be viewed as a measure of the asymmetry of coupling between the point contact and a source of the perturbation.
warning/0004/astro-ph0004132.html
ar5iv
text
# Merging history as a function of halo environment ## 1. INTRODUCTION A significant fraction of mass in the universe is believed to be in the form of dark matter (DM). According to the standard theoretical paradigm of structure formation, small-mass DM perturbations collapse first and the resulting objects then merge to form increasingly larger DM halos. Baryonic matter (gas) is assumed to follow the gravitationally dominant dark matter. Galaxies, thus, could have been formed within dense DM halos when the infalling gas reaches sufficiently high overdensities to cool, condense, and form stars. The most convincing observational evidence for substantial amounts of dark matter even in the very inner regions of galaxies comes from HI studies of dwarf and low surface brightness galaxies. The gravitational domination of DM on the scale of galaxy virial radius implies that collisionless simulations can be used to study the formation of the DM component of galaxies. Interactions between halos, such as mergers, collisions, and tidal stripping, are thought to play a crucial role in the evolution of galaxies. In particular, there is a substantial evidence that elliptical galaxies may have formed by mergers of disk systems (e.g., Barnes 1999). Observations of faint distant systems indicate that interaction rate rapidly increases with redshift (e.g., Abraham 1999). Intuitively, one could expect that the merging rate of galaxies should depend on environment (in particular, on the local density and velocity dispersion). For example, Makino & Hut (1997) found under some simplifying assumptions that the merging rate in clusters is proportional to $`n^2\sigma ^3`$ where $`n`$ is the number density of galaxies in the cluster and $`\sigma `$ is their one dimensional velocity dispersion of galactic velocities. Since the environment changes with time one could also expect dramatic changes in the evolution of the merging rate. In order to study the evolution of the merging rate and its dependence on environment one must follow the evolution of halos in a representative cosmological volume. Moreover, the simulation must have sufficiently high mass and force resolutions. Insufficient resolutions leads to structureless virialized halos instead of systems similar to observed groups and clusters of galaxies with wealth of substructure. This effect is well known as the overmerging problem (e.g., Moore et al. 1996, Frenk et al. 1996, Klypin et al. 1999). Cosmological scenarios with cold dark matter (CDM) alone cannot explain the structure formation both on small and very large scales. Variants of the CDM model with a non-zero cosmological constant, $`\mathrm{\Lambda }`$, have proven to be very successful in describing most of the observational data at both low and high redshifts. Moreover, from a recent analysis of 42 high-redshift supernovae Perlmutter et al. (1999) found direct evidence for $`\mathrm{\Omega }_\mathrm{\Lambda }=0.72`$, if a flat cosmology is assumed. Also, from the recent BOOMERANG data Melchiorri et al. (1999) found strong evidence against an open universe with $`\mathrm{\Lambda }=0`$. For our study we have chosen a spatially flat cosmological model with a cosmological constant $`\mathrm{\Omega }_\mathrm{\Lambda }=0.7`$ and the present-day Hubble constant of $`H_0=70`$ km/s/Mpc. The goal of this study is to determine the distribution with redshift of merging events for halos which exist at $`z=0`$. We study this merging rate and its dependence on environment of halos at $`z=0`$. The paper is organized as follows. In the next section we define the merging events studied in this paper. In § 3 we describe the cosmological model and the numerical simulation. We briefly describe our halo finding algorithm, the definition of environment and the detection of progenitors of halos. The technical details of these procedures are presented in the Appendix. We use the extended Press-Schechter formalism to test our procedure. In § 4 we discuss the merging of halos found in the simulation and compare our results with observations. In § 5 we summarize our results and briefly discuss their implications. ## 2. MERGING OF HALOS According to the hierarchical scenario, galaxies and the dark matter halos associated with them have been formed in a process of merging with other halos and accretion of small objects. Here merging denotes the coalescence of two objects with comparable masses whereas accretion means the infall of objects with masses much smaller than the mass of the accreting object. Obviously, there is no sharp distinction between the two processes. During the formation of every halo there are events (let us name them major mergers), when mass of the halo increases substantially over a short period of time. Such events are very important because they can lead to dramatic changes in the structure of dark matter halos and galaxies they harbor. For example, the increase in mass leads to the change of potential and, likely, density structure of the dark matter halo. One may expect even more dramatic changes for the baryonic component. Infalling objects may, for example, damage or even destroy stellar disk. The inflow of material may also serve as a source of fresh gas and may therefore induce increase in star formation rate. At the same time, collisions between halos may result in shock heating of the gas, which would tend to delay or prevent star formation for some period of time. Impact of a major merger on the structure of a halo or a galaxy likely depends on a particular configuration of the merging event: one massive infalling object or accretion of many small-mass objects, gas-rich or gas-poor mergers, etc. Nevertheless, it is reasonable to expect that accretion of many small halos is as damaging as merging with one massive satellite of the same total mass. Thus, it is logical to define major merger event as an accretion event in which mass of a halo increases substantionally (say, by more than 20% – 30%) over a short period of time (e.g., one dynamical time of the halo), as opposed to a single merger with a massive halo. There is another issue related to the major merger definition and statistics. One can consider mergers in a population of all halos present at a given redshift. In this case, one counts all merging events and divides the count by the total number of halos. This gives an estimate of the merging rate. The procedure should be used, for example, if one compares the frequency of close pairs with theoretical predictions. In this paper we consider a different statistics: we study merging history of present day halos. Namely, we ask the following question: what is the probability that the most massive progenitor of a $`z=0`$ halo identified at redshift $`z`$ had a major merger at this redshift? At relatively small redshifts ($`z<1`$), the merging rate of galaxy-size halos is low and the differences between the merger rates defined in these two different ways are small. At higher redshifts the differences may become substantial. We are also interested in the effect of the environment of halos on their merging rate. The environment of a given halo changes with time. An isolated halo may fall into a group or cluster and groups, in their turn, may then be accreted onto clusters. In this paper, we study the differences between merging histories of halos residing in different environments at $`z=0`$. The ultimate outcome of merging or accretion event depends on both the time interval and on the fractional mass increase within this time interval. Observationally, the time scale of merging is the time interval for which traces of the event can be observed. Physically, this time scale is of order of the dynamical time of accreting halo. A reasonable lower limit on the merging time scale is the crossing time of the halo defined as the ratio of the halo radius, $`R`$, to the typical accretion velocity, $`V`$: $$t_{\mathrm{cross}}1\mathrm{Gyr}\left(\frac{R}{200\mathrm{kpc}}\right)\left(\frac{V}{200\mathrm{km}\mathrm{s}^1}\right)^1.$$ (1) The crossing time is approximately equal to one gigayear for a wide range of halo masses. We conclude therefore that it is reasonable to consider time interval as large as 500 Myrs in analyzing the simulations. For our analysis we have used the simulation outputs at 25 time moments. Although the time intervals between two stored moments differ slightly, the mean interval is about 0.5 Gyr. In § 3.1 we will use the extended Press-Schechter formalism to test the effects of the different choices for the time interval. We find that our results do not change significantly if we vary the time interval from 0.1 to 0.5 Gyr. We have chosen a minimum fractional mass increase of 25% to define a major merger. The resulting merging rate depends slightly on the choice of this value as discussed in § 4. To summarize, in the remainder of the paper the major mergers are defined as accretion events in which mass $`M_1`$ of the most massive progenitor of a present-day halo increases by more than 25%: $`(M_2M_1)/M_2>0.25`$, where $`M_2`$ is halo mass after merging. In general, during the evolution of a halo in the simulation its mass increases due to accretion and merging. However, interacting halos may also exchange and lose mass. In particular, tidal stripping (and thus mass loss) becomes important in the dense environment of clusters (Klypin et al. 1999; Gottlöber et al. 1999a). Depending on the environment of the halo at $`z=0`$, we have divided our sample into three subsamples: isolated halos, halos in groups, and halos in clusters. Merging rates estimated using the extended Press-Schechter (EPS; Bond et al. 1991; Lacey & Cole 1993) formalism are expected to be close to those in numerical simulations. We use the EPS formalism to test robustness of our assumptions and parameters used in the analysis (e.g., the time interval of the merging). It should be noted that halos are treated differently in simulations and in the EPS formalism. By definition, in the EPS formalism the halos are isolated and their mass can only increase with time. Substructure (subhalos inside a larger halo) is not considered by the EPS. Let us consider a merger of two isolated halos with substructure. From the point of view of the EPS formalism this is a single merging event: the mass of the resulting merger product is considerably higher than the mass of the individual merging systems. In the simulation we would detect the mass growth for the most massive progenitor of the merged halo, but not for the individual subhalos present as substructure. In fact, the subhalos may even lose some mass due to the tidal stripping. To summarize, an accretion event that should be classified as a merger on mass-scale of a group or cluster, may not have a corresponding merger on the mass-scale of the galaxy-size halos belonging to the group. This example shows that somewhat different results must be expected for the merging rates of halos in the EPS formalism and in the simulation. Nevertheless, the general behavior is expected to be the same and the EPS formalism is a powerful method to test the assumptions made when studying the merging rate of halos in a numerical simulation. ## 3. COSMOLOGICAL MODEL AND NUMERICAL SIMULATION We study a flat CDM cosmological model with a non-zero cosmological constant ($`\mathrm{\Lambda }`$CDM). The model has following parameters: $`\mathrm{\Omega }_0=1\mathrm{\Omega }_\mathrm{\Lambda }=0.3`$; $`\sigma _8=1.0`$; $`H_0=70`$ km/s/Mpc. It is normalized in accord with the four year COBE DMR observations (Bunn & White 1997) and observed abundance of galaxy clusters (Viana & Liddle 1996). The age of the universe in this model is $`13.5`$ Gyrs. The main goal of this study is the evolution of both isolated halos and halos located inside virial radii of larger group- and cluster-size systems. This requires high force and mass resolution of the simulation. The force and mass resolution required for a simulated halo to survive in the high-density environments typical of groups and clusters is $`13`$ kpc and $`10^9\mathrm{M}_{}`$, respectively (Klypin et al. 1999). We use the Adaptive Refinement Tree (ART) $`N`$-body code (Kravtsov, Klypin & Khokhlov 1997) to follow the evolution of $`256^3`$ dark matter particles with the range in spatial resolution of $`32,000`$. With the required resolution we can simulate the formation of halos in a box of $`60h^1\mathrm{Mpc}`$. With $`256^3`$ dark matter particles the particle mass is $`1.1\times 10^9h^1\mathrm{M}_{}`$. We reach the force resolution of $`2h^1\mathrm{kpc}`$. In the box of this size there are sufficiently large number of halos in different environments. This allows us to study the merging rate of halos in different environments. The evolution of an isolated halo (galaxy-, group-, or cluster-size), whose mass grows due to accretion and merging, is relatively simple. This mass growth is well described by the extended Press-Schechter model. However, in this analysis we are interested not only in the evolution of isolated halos but also in the evolution of subhalos located within groups or clusters, i.e. in the evolution of substructures of bigger isolated objects. In fact, isolated halos and subhalos in groups or clusters evolve differently (cf. Gottlöber et al. 1999a). For example, the latter may loose mass due to tidal interaction when falling into the group or cluster or merge with the host halo if their orbit decays due to dynamical friction. The bottom panel of Fig. 1 shows a typical medium-size cluster of mass $`1.5\times 10^{14}h^1\mathrm{M}_{}`$ and diameter of about $`3h^1\mathrm{Mpc}`$ (extent of the shown particle distribution) at $`z=0`$. The figure shows that the final halo contains many subhalos. The figure shows all DM particles of the cluster which are linked using the friends-of-friends (FOF) algorithm with the linking length of 0.2 times the mean interparticle separation. (This linking length approximately corresponds to the virial overdensity). The particles are colored on a gray scale according to the logarithm of the local density at particle position smoothed over a sphere of the comoving radius of $`10h^1\mathrm{kpc}`$. This cluster has formed through a merger of two massive groups; at $`z=1`$ (top of figure), the merger is still in progress. The two largest halos apparent at $`z=1`$ merged into one central object by $`z=0`$. The galaxy size halos that can be seen within this cluster were formed well before the cluster formation in a region of high density. At $`z<1`$, the cluster grows further through relatively mild accretion of dark matter and DM halos: mass increased by a factor 1.6 from $`z=1`$ to $`z=0`$. With the high resolution of the simulation we can follow the evolution of each halo (with mass above a certain threshold determined by the mass resolution of the simulation) from the moment of its formation until $`z=0`$. Fig. 2 shows the extreme case when a large halo does not have substantial substructure within virial radius. In the bottom panel of Fig. 2 we show an isolated halo of mass $`1.5\times 10^{13}h^1\mathrm{M}_{}`$. The extended dark matter halo has a diameter of about $`1h^1\mathrm{Mpc}`$. At redshift $`z=1`$ the progenitor of this halo is a group of small-mass halos with the total mass of $`9.4\times 10^{12}h^1\mathrm{M}_{}`$ and a size of about $`1.3h^1\mathrm{Mpc}`$. This is an interesting but a rare case: in the simulation we found 20 halos ($`>10^{12}h^1\mathrm{M}_{}`$) that were classified as isolated at $`z=0`$ (no subhalos of $`v_{circ}>100`$ km/s), but whose $`z=1`$ progenitors were groups of four to seven members with $`v_{circ}>100`$ km/s. Observed counterparts of such merged halos could be the massive isolated ellipticals with group-like X-ray halos (see Mulchaey & Zabludoff 1999; Vikhlinin et al. 1999). ### 3.1. Finding halos at different redshifts Identification of halos in dense environments and reconstruction of their evolution is a challenge. The most widely used halo-finding algorithms, the friends-of-friends (FOF) and the spherical overdensity, both discard “halos inside halos”, i.e. , satellite halos located within the virial radius of larger halos. In order to cure this, we have developed and used two algorithms to find halos: the hierarchical friends-of-friends (HFOF) and the bound density maxima (BDM) algorithms (Klypin et al. 1999). The HFOF algorithm uses a set of different linking lengths in order to identify the substructures of large DM halos as “halos inside halos”. The BDM algorithm does the same by identification of all local density maxima and following the density profiles starting at these points. The HFOF and BDM algorithms are complementary. Both of them find essentially the same halos above a reasonable mass threshold ($`30`$ particles). Therefore, we believe that each of them is a stable algorithm which finds in a given dark matter distribution the DM halos. The advantage of the HFOF algorithm is that it can handle halos of arbitrary, not only spherically symmetric, shape. The advantage of the BDM algorithm is that it describes better the physical properties of the halos because it separates background unbound particles from the particles gravitationally bound to the halo and constructs density and velocity profiles for each halo. It is difficult and usually ambiguous to find mass of a halo located within a larger halo. The formal virial radius of such a halo is equal to the bound system’s virial radius. If necessary, we define halo mass as mass within its tidal or truncation radius defined as a radius where the halo density profile starts to flatten. We try to avoid the problem of mass determination by assigning not only the mass to a halo, but finding also its maximum circular velocity $`v_{circ}=\sqrt{GM(<R)/R}|_{max}`$. Numerically, $`v_{circ}`$ can be measured more easily and more accurately then the mass. Beside of being more stable numerically, the circular velocity is also more meaningful observationally. For our analysis we need a halo sample which is as complete as possible but does not contain any fake halos. Recently, we have shown that the halo samples constructed from the simulation used here do not depend on the numerical parameters of the halo finder for halos with $`v_{circ}\genfrac{}{}{0pt}{}{_>}{^{}}100`$ km/s (Gottlöber et al. 1999a). At $`z=0`$, we decided to be even more restrictive and limit analyses to halos with a circular velocity of $`v_{circ}>120`$ km/s which contain more than 100 bound particles within a radius of $`100h^1\mathrm{kpc}`$. To avoid misidentifications we have required that each halo of our sample has a unique progenitor at the last five time steps (see Appendix for details). At $`z=0`$ our halo sample consists of 4193 halos. The halo number density, $`0.019h^3\mathrm{Mpc}^3`$, roughly corresponds to the number density of galaxies with $`M\genfrac{}{}{0pt}{}{_<}{^{}}18.5`$ in the Las Campanas redshift survey (Lin et al. 1996). ### 3.2. Definition of environment As mentioned above, the goal of this paper is to study the merging history of the halos as a function of their environment at present ($`z=0`$). To characterize the environment we have run friend-of-friend analysis of the simulation outputs with a linking length of 0.2 times the mean interparticle distance. The FOF algorithms, thus, identifies clusters of DM particles with average overdensity of about 200. The virial overdensity in the $`\mathrm{\Lambda }`$CDM model under consideration is about 330 which corresponds to a linking length of about 0.17. Therefore, the objects which we find have a slightly larger extent than the objects at virial overdensity. We have increased the linking length to account (at least partially) for the halos gravitationally bound to the cluster halo but located just outside at the epoch of identification. For each of the identified halos, we find a host halo if such host exists (see Appendix). We call the halo isolated, if it does not belong to any higher-mass host. We call the halo a cluster if it belongs to a particle cluster with a total mass larger than $`10^{14}h^1\mathrm{M}_{}`$. Finally, we identify group halos as halos consisting of 3 or more subhalos and having masses of $`10^{14}h^1\mathrm{M}_{}`$. Two halos located within a common cluster of overdensity 200, are considered to be pairs. In the subsequent analysis, we have omitted all pairs due to the following reasons. A massive halo with a single subhalo of much smaller mass should probably be considered isolated. Subhalos of even smaller mass could have been unresolved or missed due to the limited mass resolution so that such pair, depending on mass, should have been a small group rather than an isolated halo. To avoid this kinds of confusing identifications of the environment, we have omitted all pairs. The analyzed halos, therefore, are classified as isolated, cluster, or group halos. The procedure described above results in identification at $`z=0`$ in our $`60h^1`$ Mpc box of 401 cluster halos (10 %) in 18 clusters, 743 halos in groups (18 %) and 2545 isolated halos (60 %). The remaining 504 halos are found in pairs (12 %). The first cluster has formed between $`z=2.5`$ and $`z=2`$. The fraction of galaxies in clusters increases with time, whereas the fraction of isolated galaxies in the considered mass range decreases, and the fraction of pairs remains approximately constant (Gottlöber et al. 1999c). ### 3.3. Progenitors of halos For each of the halos in our $`z=0`$ sample we have constructed a complete evolution tree over 25 epochs ($`z=0`$, 0.05, 0.1, 0.15, 0.2, 0.25, 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, 0.9, 1.0, 1.2, 1.5, 1.8, 2.0, 2.5, 3.0, 4.0, 5.0, 6.1, 7.2, 10.0). Procedure of progenitor identification is based on the comparison of lists of particles belonging to the halos at different redshifts both back and forward in time (for details see the Appendix). Our algorithm of tracing halo histories identifies the correct “ancestor-descendant” relationships rather accurately, with estimated ancestor-descendant misidentifications $`\genfrac{}{}{0pt}{}{_<}{^{}}2\%`$ of the cases (Gottlöber et al. 1999a). These misidentifications happen usually with small-mass halos consisting of only a few tens of particles (below the mass limit of halos in our sample), i.e. the halos in the mass range of our sample are not significantly affected. Now we define the halo detection epoch as the epoch at which the halo has reached a threshold when it can be expected to host a galaxy. The threshold is taken to be $`v_{circ}>50`$ km/s for the first time. Our halo-finder assumes a minimum circular velocity, $`v_{circ}>50`$ km/s, and a minimum number of bound particles of 40, for the progenitors of our halos at $`z>0`$. In Fig. 3 we show how many cluster and isolated halos were detected per redshift interval as a function of redshift. In this figure we show only cluster halos; halos in groups show similar behavior. Due to the construction procedure of our sample (§ 3.1), all present-day halos exist already at $`z=0.3`$. Fig. 3 shows that in general cluster halos form earlier than isolated halos. The maximum formation rate is reached at $`z3`$ whereas the maximum of the formation rate of isolated halos is reached later ($`z<2.5`$). Cluster halos form in regions of higher overdensity and therefore reach the detection threshold earlier. The integral over the curves of Fig. 3. gives us the completeness of the progenitor samples as a function of redshift. It reaches 90 % (50 %) at $`z=2`$ ($`z=4`$) for halos in clusters and at $`z=1.5`$ ($`z=3`$) for isolated halos. Let us consider a subsample of halos with $`150>v_{circ}>200`$ km/s (Fig. 4, thick lines). Cluster halos of this sample form earlier ($`z4`$) than isolated halos. We see the same tendency for the subsample of the most massive halos with $`v_{circ}>300`$ km/s (thin lines), however due to poor statistics we do not see the maximum of formation rate of these massive cluster halos. The progenitor of the central halo of the most massive cluster (which corresponds to a massive central cD galaxy) can be identified at $`z=15`$ as an object of $`3\times 10^{10}h^1\mathrm{M}_{}`$. ### 3.4. Extended Press-Schechter formalism To test how sensitive the merger rate estimates are to our assumptions, we use the synthetic merger histories generated using the extended Press-Schechter formalism (hereafter EPS, Bond et al. 1991; Bower 1991; Lacey & Cole 1993). Specifically, we use the “$`N`$-branch trees with accretion” method of Somerville & Kolatt (1999) to construct merger histories for the host. In this method the progenitors of a halo of mass $`M_h`$ at a given epoch and in mass range $`[M_p^{min},M_h]`$ are determined through a series of Monte-Carlo picks using probability for a halo to have accreted mass $`\mathrm{\Delta }M`$ during time $`\mathrm{\Delta }t`$ (see eq. 2.29; of Lacey & Cole 1993). $`M_p^{min}`$ is the minimum progenitor mass that is kept track of, specific to particular intended use of the merger histories. The method was slightly modified; at each step we require the number of progenitors in the mass range of interest to be close to the expected average. This modification significantly improves agreement of the progenitor mass function generated by the method with the analytical prediction (Kravtsov et al. 2000). We refer reader to Somerville & Kolatt (1999) for further details of the method. For each step in time, a merger history contains information about the mass of the host at the current epoch and masses of its progenitors at the previous epoch. The most massive progenitor in the list is assumed to represent the host halo a time step back in time, while the other progenitors are considered to be the halos accreted during the step. The time steps were chosen as prescribed by Somerville & Kolatt (1999). Using the EPS method we have generated a sample of 1350 merging histories of halos of mass $`10^{12}M_{}`$ at $`z=0`$. This is a typical mass for halos in our numerical catalogs. We followed merger histories back in time until the mass of the most massive progenitor falls below $`4\times 10^{10}M_{}`$, the same minimum mass that was used in analysis of the simulation. With a mean $`\mathrm{\Delta }z`$ of about 0.003 the shortest history (425 steps) starts at $`z=1.4`$, whereas the longest (3802) starts at $`z=11`$. Our goal is to use the high temporal resolution of the EPS merging histories to test the effect of the coarse temporal resolution of the merger histories constructed using simulation. As discussed in § 2, we define a major merger as an event in which the mass of a halo increases by more than a certain threshold in a given time interval. In Fig. 5 we show five of the original EPS merger histories (solid lines) and the major merger events detected (solid triangles) using the original time steps of the history and a high threshold of 0.35. The dashed lines show the evolution of the same halos tracked only at 25 time moments used in the simulation (see § 3.3) where the open triangles denote the major merger events detected for this history using the same definition of the major merger as before. One can see that at $`z2`$ each solid triangle is accompanied by an open one. At higher $`z`$ there are exceptions. In one of the histories two subsequent merger events detected in the original merger history are detected as only one event in the coarse-time history. In another history (rightmost curve) two successive minor mergers in the original history (at $`z5`$), that are not classified as “major” add up to a major merger when the time resolution is degraded. The fact that two (or more) successive minor mergers occuring within less than 0.5 Gyrs are treated as one major merger is not necessarily wrong. The physical effect of large mass accretion within a short time interval may be the same regardless of whether this accretion was through a single major or several minor mergers. However, the fact that coarse temporal resolution may masquerade several major mergers as one, may lead to an underestimate of the merger rate at $`z2`$. In the following we want to test whether the estimates of the merging rate are influenced by the relatively coarse time intervals used in simulation analysis. To this end, we tracked the original EPS merger histories in equally spaced time intervals of 500 Myr, 250 Myr, and 100 Myr. We have then counted the major mergers as events in which mass increases by more than 25% in a given time interval. Finally, we have calculated the merging rate of the EPS halos in the same way as for the halos of the simulation: the number of major merger events per halo per Gigayear. The result is shown in the three panels of Fig. 6. The solid line is a power law fit ($`[1+z]^\alpha `$) for $`z<2`$ epochs, with $`\alpha =2.40`$ for $`\mathrm{\Delta }t=500`$ Myr, $`\alpha =2.38`$ for $`\mathrm{\Delta }t=250`$ Myr, and $`\alpha =2.27`$ for $`\mathrm{\Delta }t=100`$ Myr. There is a good agreement of the merging rates for different time intervals. The somewhat lower merging rate in case of $`\mathrm{\Delta }t=100`$ Myr is a result of the poor statistics at low $`z`$. Obviously, the merging rate increases with redshift, but it cannot increase to infinity due to the adopted major merger definition. Clearly, if one assumes that all halos underwent a major merger in the given time interval $`\mathrm{\Delta }t`$ (the algorithm by design cannot detect multiple major merger events in a particular time interval, see above), the merging rate reaches its maximum value $`1/\mathrm{\Delta }t`$, i.e. 2, 5, or 10 in the three panels of Fig. 6. One can see that the curves in this figure indeed flatten before reaching this value, but at different epochs. This flattening is due to the limited mass resolution of the simulation. At a given time moment and with a given resolution, some of the halos fall below the detection threshold at the earlier time moment, i.e. for these halos major merging events cannot be detected. With larger $`\mathrm{\Delta }t`$ this probability increases and reduces therefore the rate and epoch at which this effect sets in. However, Fig. 6 shows that for the interval of $`\mathrm{\Delta }t=0.5`$ Gyr (the interval used in our simulation analysis), the results are robust for $`z2`$. In Fig. 7 we compare two different detection schemes of major merger events. Open circles denote all major merger events detected by the condition $`(M_2M_1)/M_2>0.25`$. These events include events due to the merger of two massive halos and multiple mergers and/or rapid accretion. The filled circles denote the subsample of binary mergers of two massive progenitors. Note that in the upper panel almost all open circles coincide with filled circles at $`M>2\times 10^{12}h^1\mathrm{M}_{}`$, i.e. the major merger events detected through the mass growth are in fact due to binary mergers. With decreasing mass and increasing redshift the number of major mergers with only one massive progenitor rapidly increases. This is simple due to the fact that with decreasing mass of the halo the probability that its second massive progenitor is already below the detection threshold increases. Although in the EPS formalism one could easily extend the tree to lower masses, in simulations one is limited by the mass resolution. In the lower panel of Fig. 7 we plot the merging rates detected for all major mergers (open circles, the same as the top panel of Fig. 6) and for binary major mergers (filled circles). There is no differences at $`z2`$. We conclude that there is no difference between the two definitions at redshifts $`z2`$. It is more reasonable to analyze the simulation using the mass growth, $`(M_2M_1)/M_2`$, as a major merger definition, because this procedure provides informations to somewhat larger redshifts (see Fig. 7, lower panel). More investigations are necessary to understand whether the bend at higher $`z`$ is real. ## 4. RESULTS AND DISCUSSION As was noted above, to identify major mergers we calculate the relative mass growth $`(M_2M_1)/M_2`$ during the time interval $`t_2t_1`$. Due to the selection of simulation output redshift, the time intervals (see Sect. 3.3) have somewhat variable length. As a reminder, we assume a major merger to have occured if the relative mass growth is larger than 0.25. As discussed in the previous section, we calculate the total change of mass, not only the contribution of merging with another massive halo. In the previous section we have shown that variable time intervals do not change the result if this merger rate is normalized by time interval. Therefore, we calculate the merger rate as the number of major-merging events of the progenitors of our halo sample per gigayear normalized to the number of progenitors at a given moment. In Fig. 8 we show the evolution of the merger rate in this definition. The estimated merger rate is obtained by averaging over three subsequent time intervals in order to reduce the scatter. The error bars are $`\sqrt{N}`$ errors for the number of events detected. For redshifts $`z2`$ our sample is 90 % complete. For these redshifts the merger rate can be fitted by a simple power law $`(1+z)^{3.0}`$. There is an indication of flattening of the merging rate at higher redshifts. As shown in Fig. 3 at $`z>2`$ we are rapidly loosing the halo progenitors due to mass resolution. Moreover, Fig. 6 also shows that this effect is likely to be due to the limited mass resolution. However, as we have argued above, prediction for the evolution at $`z2`$ is reliable. Recently, Le Fèvre et al. (1999) published the first direct observational measurement of the merger fraction at redshifts $`z>0.5`$. They have used visual merger identifications as well as statistics of close pairs of galaxies. To transform the observed merger fraction into a merger rate requires knowledge of the lifetime of a merger, i.e. the time for which the traces of the merging event can be observed. An upper limit of 0.4 to 1 Gyr, as assumed by Le Fèvre et al. (1999), approximately corresponds to the time step which we have used to derive the merging rate. Le Fèvre et al. (1999) have derived a merger rate varying with redshift as $`(1+z)^{3.2\pm 0.6}`$. This result is in good agreement with our theoretical prediction. In principle, the evolution of the merger rate is a very important observable for testing cosmological models. In practice, both the theoretical and observational estimates of merging rates depend on a number of assumptions. In our model estimates, the result depends mainly on the definition (mass threshold) of major merger events. Increasing the threshold 0.25 to 0.4, corresponds to a faster evolution described by $`(1+z)^{3.7}`$. Lowering the threshold to 0.20, results in a slightly slower evolution of the merger rate. Observationally, merger rate evolution is determined by studies of the evolution of the correlation function, the evolution of pair numbers or by morphological studies (e.g., Abraham 1999). In the first two cases the merging rate is not directly measured and the conversion of the measured quantities to the merging rate is not straightforward. The third approach, i.e. observing mergers in progress, is the most direct one (cf. Le Fèvre et al. 1999). From an excess of power in the observed two-point angular correlation function at angular scales $`2^{\prime \prime }\theta 6^{\prime \prime }`$ Infante et al. (1996) determined a merging rate of galaxies $`(1+z)^{2.2\pm 0.5}`$. According to Burkey et al. (1994) the pair fraction in deep HST images grows with redshift as $`(1+z)^{3.5\pm 0.5}`$ from which they deduce a merger rate $`(1+z)^{2.5\pm 0.5}`$. On the contrary, Carlberg et al. (1994) conclude from the evolution of the close pair ($`\theta 6^{\prime \prime }`$) fraction that the merger rate–redshift relation is $`(1+z)^{3.4\pm 1.0}`$. Finally, Yee & Ellingson (1995) found for the close pair (projected distance less than $`20h^1\mathrm{kpc}`$) evolution rate a somewhat steeper redshift dependence $`(1+z)^{4.0\pm 1.5}`$. Different authors used different relationships between the evolution of pair fraction and the evolution of the merger rate. Nevertheless, the derived merger rates are in general agreement and also in agreement with our theoretical predictions. It is very interesting to see whether the merger rate evolution depends on the halo environment. In Fig. 9 we show the merger rate of cluster, group and isolated halos normalized to the merger rate of all halos shown in Fig. 8. The higher rate of major mergers at early epochs for cluster and group halos is due to the higher density in regions, where cluster and groups have been forming. Note that the clusters have not yet existed at $`z23`$. As clusters with large internal velocities form, merging rate quickly decreases (Makino and Hut 1997; Kravtsov & Klypin 1999; Mamon 2000). There are almost no major merger events of cluster halos in the recent past. Those 8 events at $`z=0.1,\mathrm{\hspace{0.33em}0.25}`$, and 0.5 have probably happened just outside the clusters before the halos were accreted by cluster or, alternatively, they might have occured within the surviving low internal velocity substructute within cluster. Note also that the massive central halo of the cluster (which should correspond to the observed brightest cluster galaxies) may accrete other halos after the cluster has been formed (Dubinski 1998; Mamon 1999). If the accreted halo is massive enough a major merging event would be detected. The low merger rate in simulated clusters has been noted also by Ghigna et al. (1998), although mergers within accreted substructure have been observed (Springel et al. 2000). ## 5. CONCLUSIONS We have analyzed a high-resolution collisionless simulation of the evolution of structure in a $`\mathrm{\Lambda }`$CDM model. We have followed the formation and evolution of DM halos in different cosmological environments and estimated the evolution of the major merger rate of dark matter halos. We have found that regardless of their present-day mass, halos that end in clusters form earlier than isolated halos of the same mass (Figs. 3 and 4). We find that at redshifts $`z2`$ major merger rate evolves as $`(1+z)^{3.0}`$, in good agreement with observations. Finally, we have calculated the merging rate evolution as a function of halo environment at $`z=0`$ (Fig. 9). The merger rate of halos located in clusters or groups at present increases faster back in time than that of isolated halos. The cluster and group halos are therefore predicted to have a higher rate of major merger events in the past. At $`z1`$, the merger rate of cluster and group halos drops very quickly, while numerous major merger events for isolated halos have been detected down to $`z=0`$. This implies possible systematic differences between cluster and field ellipticals. Evidence for such differences was found by de Carvalho & Djorgovski (1992), while Bernardi et al. 1998 detected close similarity between cluster and field early type galaxies. The agreement between theoretical predictions and observations are encouraging and supports the validity of the hierarchical structure formation scenario. Future, higher resolution simulations should extend the predictions presented here to higher redshift and, in case of gasdynamics simulations, provide a more straightforward connection to observations. On the observational side, the ever increasing size of the high-redshift galaxy samples should also allow estimates of the merger rate at high redshifts in the near future. This work was funded by the NSF and NASA grants to NMSU. SG acknowledges support from Deutsche Akademie der Naturforscher Leopoldina with means of the Bundesministerium für Bildung und Forschung grant LPD 1996. A.V.K. was supported by NASA through Hubble Fellowship grant HF-01121.01-99A from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. We acknowledge support by NATO grant CRG 972148. Appendix: Technical details A: Halo identification: Our halo identification algorithm (see Klypin et al. 1999 for more details) starts with the search of local density maxima assuming a smoothing radius of the order of $`10h^1\mathrm{kpc}`$. This radius defines the scale of the smallest objects we are looking for. Once the centers of potential halos are found, we start the procedure of removing unbound particles and finding the size of halos. We determine the mass of the dark matter particles in concentric spherical shells around the halo center, the mean shell velocity, and the velocity dispersion relative to the mean. We assume that particles with velocities larger than the escape velocity at the position of the particle are not bound to the halo and do not take them into account when calculating halo properties. Using the density profile of the halo we estimate the maximum rotational velocity and the radius at which the maximum is reached. We will call this radius, radius of the halo. Removal of unbound particles is important in the case when a halo with a small internal velocity dispersion moves inside a larger halo. In many cases, the small-mass subhalos can be unambigously identified only after removing the unbound background particles of the larger halo. To avoid misidentifications of group or cluster halos as galaxy halos we have introduced a maximum possible halo radius of $`100h^1\mathrm{kpc}`$. We assume this radius to be the halo radius if the circular velocity profile is still increasing at this distance. The halo identification scheme is crucial for the algorithm which finds the most massive progenitor of a halo. Due to our halo definition, DM particles can belong to more than one halo (e.g., a particle may be gravitationally bound to both the subhalo and to its host halo) At a given time moment, some of the particles gravitationally bound to a halo might be outside of the formally defined halo radius. Both situations must be taken into account if one constructs the evolution history of halos using lists of particles belonging to a given halo. We use the constructed density profile and estimated radius to assign each halo a mass and circular velocity. To construct halo catalogs, we select all halos with maximum circular velocities above a certain minimum value. In addition, we have rejected all halos with the number of bound particles below a minimum threshold. There are also some special cases in halo identification. For example, after a recent merger two local density maxima could be found within a common halo. The halo finder tends to identify such a configuration as two halos at a distance of the order or smaller than the radius of the halos. Also, at a given moment small clumps in a dense environment could, by chance, appear as bound clumps which, however, disappear by the next time moment. These misidentifications could masquerade as recent merging events. To avoid these kinds of misidentifications, we have required that each halo of the sample at $`z=0`$ has a unique progenitor at the last five time steps (see § 3.3). If a progenitor has been already identified as progenitor of another halo, the smaller mass halo is discarded. This procedure, which removes about 6% of the halos, reduces the number of fake merger detections and makes the results more reliable. B: Definition of environment: We define the environment of a halo using the mass of the virialized host (if any) to which the halo belongs. We find virialized systems using a friends-of-friends algorithm with a linking length of 0.2 times the mean density. For each halo, we determine the particle with the shortest distance to the center of mass of the halo. Then we search for the virialized object to which the particle belongs. Since these objects have overdensities of the order 200, we are confident that our halo of higher internal overdensity belongs to a virialized host if the most central particle belongs to it. C: Progenitor identification: We begin by identification of all particles bound to a halo at a given redshift $`z_i`$. Then we find all objects (identified halos, small groups of particles below the detection limit of halos, isolated particles) at the previous redshift $`z_{i+1}`$ which contain any of the particles of the halo in question. The time interval between the two redshifts is typically of the order of 0.5 Gyr. We repeat this procedure for all halos at a given redshift. We thus obtain complete information on the origin of the particles found in halos at redshift $`z_i`$. However, due to the halo identification procedure described above this information is not equivalent to the mass of the progenitor objects. Indeed, we have determined the mass at the radius of the maximum rotational velocity. To get the mass growth due to merging and accretion we must compare the masses of the halo and its progenitor which are defined in the same way. It is relatively straightforward to identify the most massive progenitor for isolated halos: it is simply the halo which contains the largest fraction of particles of the descendant. In a dense environment, on the other hand, the identification is more complicated. For example, the particles of a subhalo belong both to the satellite halo and the hosting halo at an earlier moment, so that both would be identified as progenitors. To avoid this (and similar) misidentifications we identify not only the ancestors of all halos found at $`z_i`$ but also the descendants of all halos found at $`z_{i+1}`$ and check whether a given halo is really the descendant of its ancestor by searching for the maximum subsamples of particles belonging to the corresponding halos. With such somewhat extensive procedure we reduce the misidentifications of progenitors to $`\genfrac{}{}{0pt}{}{_<}{^{}}2\%`$ of all considered cases. Note, that these misidentifications lead to scatter in the mass evolution history.
warning/0004/hep-ph0004088.html
ar5iv
text
# Power-law inflation with a nonminimally coupled scalar field ## I Introduction The idea of inflation is one of the most reliable concepts in modern cosmology. It can solve the horizon and flatness problem in the standard big bang cosmology, and also provides us the seeds of the large scale structure. The inflationary period proceeds while a scalar field called inflaton slowly evolves along a sufficiently flat potential. It has been generally considered that only one scalar field determines the dynamics of inflation even if there are many scalar fields present in the inflationary epoch. However, Liddle, Mazumdar, and Schunck recently showed that we have an inflationary solution with exponential potentials in the multi-scalar field case even if individual fields do not possess flat potentials to lead to inflation. It was demonstrated that scaling solutions for exponential potentials with different slopes are the late-time attractors, and Malik and Wands confirmed this fact by choosing an appropriate rotation in field space. Copeland, Mazumdar, and Nunes examined the generalized assisted inflation where the cross-coupling terms exist between scalar fields. Exponential potentials often arise in the effective four-dimensional models induced by Kaluza-Klein theories. Kanti and Olive investigated the assisted chaotic inflation with multiple scalar fields in higher-dimensional theories. The dynamics and density perturbations in assisted chaotic inflation was studied by Kaloper and Liddle. Since most models in realistic higher-dimensional theories give rise to steep potentials by which inflation is hard to realize, the assisted mechanism by multiple scalar fields plays an important role for the realization of inflation. From a viewpoint of quantum field theory in curved spacetime, it is natural to consider that the inflaton field $`\varphi `$ couples nonminimally to the spacetime curvature $`R`$ with a coupling of $`\xi R\varphi ^2/2`$. In the new inflation model, the existence of nonminimal coupling prevents inflation in some cases, because the flatness of the potential of inflaton is destroyed around $`\varphi =0`$. In the chaotic inflation model with a nonminimally coupled inflaton field, Futamase an Maeda investigated the constraint of the coupling $`\xi `$ in two potentials of $`V(\varphi )=m^2\varphi ^2/2`$ and $`V(\varphi )=\lambda \varphi ^4/4`$. They found that $`\xi `$ is restricted as $`|\xi |\stackrel{<}{}10^3`$ to lead to sufficient inflation in the massive inflaton model. On the other hand, the constraint of $`\xi `$ is absent in the self coupling model with negative $`\xi `$, and inflation is supported for larger values of $`|\xi |`$. Fakir and Unruh examined this self-coupling model with a strong negative nonminimal coupling $`|\xi |1`$, and found that the fine tuning problem of the self-coupling $`\lambda `$ is relaxed by such large values of $`|\xi |`$. Several authors studied scalar density perturbations and tensor gravitational waves during inflation in this model. In the context of preheating after inflation, we have showed that the fluctuation of inflaton can be enhanced nonperturbatively by nonminimal coupling. In this paper, we consider power-law inflation with a nonminimally coupled inflaton field. What we are concerned with is whether power-law inflation is supported or not with the presence of nonminimal coupling. If the assisted mechanism works even in the one-field case, it is expected that this will also occur in the multi-field case. For a more complicated study of the multi-field case in the future, in this paper, we clarify in what values of $`\xi `$ assisted inflation is realized by nonminimal coupling in the one-field case. This paper is organized as follows. In the next section, we explain the model of power-law inflation with a nonminimally coupled inflaton field $`\varphi `$. In Sec. III, the dynamics of inflation is investigated in both cases of $`\xi >0`$ and $`\xi <0`$. We show in what cases the assisted inflation takes place by the effect of nonminimal coupling. We present our conclusions and discussions in the final section. ## II Basic equations We study a model where an inflaton field $`\varphi `$ is nonminimally coupled with a scalar curvature $`R`$: $`=\sqrt{g}\left[{\displaystyle \frac{1}{2\kappa ^2}}R{\displaystyle \frac{1}{2}}(\varphi )^2V(\varphi ){\displaystyle \frac{1}{2}}\xi R\varphi ^2\right],`$ (1) where $`\kappa ^2/8\pi G=m_{\mathrm{pl}}^2`$ is Newton’s gravitational constant, and $`\xi `$ is a coupling constant. In this paper, we consider an exponential potential $`V(\varphi )`$ which is described by $`V(\varphi )=V_0\mathrm{exp}\left(\sqrt{{\displaystyle \frac{16\pi }{p}}}{\displaystyle \frac{\varphi }{m_{\mathrm{pl}}}}\right),`$ (2) where $`V_0`$ and $`p`$ denote the energy scale which has the dimension of $`[\mathrm{mass}]^4`$, and the power of inflation, respectively. In the case of $`\xi =0`$, we have an inflationary solution for $`p>1`$, $`a(t)t^p,`$ (3) where $`a(t)`$ is the scale factor. However, inflation does not take place for $`p1`$ in the case of a single scalar field. Liddle et al. showed that inflation proceeds for the case of multi-scalar fields even if the individual scalar field has the power less than unity. This cooperative behavior was termed assisted inflation. In this paper, we investigate whether nonminimal coupling will assist inflation or not in the single field case. The multi-scalar field case will be discussed elsewhere. We find from the Lagrangian $`(\text{1})`$ that the effective gravitational constant $`G_{\mathrm{eff}}`$ depends on the value of the inflaton field, $`G_{\mathrm{eff}}={\displaystyle \frac{G}{1\varphi ^2/\varphi _c^2}},\mathrm{with}\varphi _c^2{\displaystyle \frac{m_{\mathrm{pl}}^2}{8\pi \xi }}.`$ (4) In order to connect to our present universe, $`G_{\mathrm{eff}}`$ needs to be positive for the case of the positive $`\xi `$, which yields $`|\varphi |<\varphi _c={\displaystyle \frac{m_{\mathrm{pl}}}{\sqrt{8\pi \xi }}}.`$ (5) When $`\xi `$ is negative, such a constraint is absent. We obtain the following field equations from the Lagrangian $`(\text{1})`$, $`{\displaystyle \frac{1\xi \kappa ^2\varphi ^2}{\kappa ^2}}G_{\mu \nu }`$ $`=`$ $`(12\xi )_\mu \varphi _\nu \varphi \left({\displaystyle \frac{1}{2}}2\xi \right)g_{\mu \nu }(\varphi )^2g_{\mu \nu }V(\varphi )+2\xi \varphi (g_{\mu \nu }\text{ }\text{ }\text{ }\text{ }\text{ }_\mu _\nu )\varphi ,`$ (6) $`\text{ }\text{ }\text{ }\text{ }\text{ }\varphi \xi R\varphi V,_\varphi =0,`$ (7) where and $`V,_\varphi `$ are defined as $`\text{ }\text{ }\text{ }\text{ }\text{ }_\mu (\sqrt{g}g^{\mu \nu }_\nu )/\sqrt{g}`$, $`V,_\varphi V/\varphi `$ respectively. Although we can analyze the evolution of the system by Eqs. $`(\text{6})`$ and $`(\text{7})`$, it is rather complicated due to the existence of nonminimal coupling. It is convenient to transform to the Einstein frame by performing a conformal transformation $`\widehat{g}_{\mu \nu }=\mathrm{\Omega }^2g_{\mu \nu },`$ (8) where $`\mathrm{\Omega }^21\xi \kappa ^2\varphi ^2`$. Then we obtain the following equivalent Lagrangian: $`=\sqrt{\widehat{g}}\left[{\displaystyle \frac{1}{2\kappa ^2}}\widehat{R}{\displaystyle \frac{1}{2}}F^2(\widehat{}\varphi )^2\widehat{V}(\varphi )\right],`$ (9) where variables with a caret denote those in the Einstein frame, and $`F^2{\displaystyle \frac{1(16\xi )\xi \kappa ^2\varphi ^2}{(1\xi \kappa ^2\varphi ^2)^2}},`$ (10) $`\widehat{V}(\varphi ){\displaystyle \frac{V(\varphi )}{(1\xi \kappa ^2\varphi ^2)^2}}.`$ (11) Introducing a new scalar field $`\mathrm{\Phi }`$ as $`\mathrm{\Phi }{\displaystyle F(\varphi )𝑑\varphi },`$ (12) the Lagrangian in the new frame is reduced to the canonical form: $`=\sqrt{\widehat{g}}\left[{\displaystyle \frac{1}{2\kappa ^2}}\widehat{R}{\displaystyle \frac{1}{2}}(\widehat{}\mathrm{\Phi })^2\widehat{V}(\mathrm{\Phi })\right].`$ (13) In this paper, we adopt the flat Friedmann-Robertson-Walker line element as the background spacetime; $`ds^2=dt^2+a^2(t)d𝐱^2=\mathrm{\Omega }^2(d\widehat{t}^2+\widehat{a}^2(\widehat{t})d𝐱^2).`$ (14) Note that $`\widehat{t}`$ and $`\widehat{a}`$ are related with those in the original frame as $`\widehat{t}={\displaystyle \mathrm{\Omega }𝑑t},\widehat{a}=\mathrm{\Omega }a.`$ (15) The evolutions of the scale factor and the $`\mathrm{\Phi }`$ field in the Einstein frame yield $`\widehat{H}^2\left({\displaystyle \frac{\widehat{a}_{,\widehat{t}}}{\widehat{a}}}\right)^2={\displaystyle \frac{\kappa ^2}{3}}\left[{\displaystyle \frac{1}{2}}\mathrm{\Phi }_{,\widehat{t}}^2+\widehat{V}(\mathrm{\Phi })\right],`$ (16) $`\mathrm{\Phi }_{,\widehat{t}\widehat{t}}+3\widehat{H}\mathrm{\Phi }_{,\widehat{t}}+\widehat{V}_{,\mathrm{\Phi }}=0,`$ (17) where $`,\widehat{t}d/d\widehat{t}`$ and $`\widehat{V}_{,\mathrm{\Phi }}{\displaystyle \frac{d\widehat{V}}{d\mathrm{\Phi }}}={\displaystyle \frac{4\xi \kappa ^2\varphi \sqrt{16\pi /pm_{\mathrm{pl}}^2}(1\xi \kappa ^2\varphi ^2)}{(1\xi \kappa ^2\varphi ^2)^2\sqrt{1(16\xi )\xi \kappa ^2\varphi ^2}}}V_0\mathrm{exp}\left(\sqrt{{\displaystyle \frac{16\pi }{p}}}{\displaystyle \frac{\varphi }{m_{\mathrm{pl}}}}\right).`$ (18) Note that $`\varphi _{,\widehat{t}}`$ and $`\mathrm{\Phi }_{,\widehat{t}}`$ are related by Eq. $`(\text{12})`$ as $`\varphi _{,\widehat{t}}={\displaystyle \frac{1\xi \kappa ^2\varphi ^2}{\sqrt{1(16\xi )\xi \kappa ^2\varphi ^2}}}\mathrm{\Phi }_{,\widehat{t}}.`$ (19) We can know the behavior of the scale factor and the inflaton field by Eqs. $`(\text{16})`$ and $`(\text{17})`$ with an effective potential $`(\text{11})`$ in the Einstein frame. Transforming back to the original frame by making use of the relation $`(\text{15})`$, we can judge in what cases nonminimal coupling assists inflation. In what follows, we will investigate these issues in details. ## III Assisted inflation with a nonminimally coupled scalar field In this section, we study the dynamics of the system in both cases of the positive and negative $`\xi `$. Let us first review the case of $`\xi =0`$ for comparison. In this case, making use of the slow-roll conditions $`\dot{\varphi }^2V,\ddot{\varphi }3H\dot{\varphi }`$, Eqs. $`(\text{16})`$ and $`(\text{17})`$ are approximately written as $`H^2`$ $``$ $`{\displaystyle \frac{\kappa ^2}{3}}V_0\mathrm{exp}\left(\sqrt{{\displaystyle \frac{16\pi }{p}}}{\displaystyle \frac{\varphi }{m_{\mathrm{pl}}}}\right),`$ (20) $`3H\dot{\varphi }`$ $``$ $`\sqrt{{\displaystyle \frac{16\pi }{p}}}{\displaystyle \frac{V_0}{m_{\mathrm{pl}}}}\mathrm{exp}\left(\sqrt{{\displaystyle \frac{16\pi }{p}}}{\displaystyle \frac{\varphi }{m_{\mathrm{pl}}}}\right),`$ (21) where we dropped a caret in the Hubble parameter. Then we find that $`\varphi `$ and $`H`$ evolve as $`\varphi `$ $``$ $`\sqrt{{\displaystyle \frac{p}{4\pi }}}\mathrm{log}\left[\sqrt{{\displaystyle \frac{\kappa ^2V_0}{3}}}{\displaystyle \frac{t}{p}}+\alpha \right]m_{\mathrm{pl}},`$ (22) $`H`$ $``$ $`\left({\displaystyle \frac{t}{p}}+\beta \right)^1,`$ (23) where $`\alpha `$ and $`\beta `$ are some constants which depend on initial values of the $`\varphi `$ field. We obtain the power-law solution $`(\text{3})`$ by integrating Eq. $`(\text{23})`$. When $`p`$ is greater than unity, the potential $`(\text{11})`$ is flat enough to lead to inflation. Since the exponential potential does not have a local minimum, the inflationary phase continues forever. In realistic models of inflation, we need to consider some exit mechanisms from power-law inflation in order to lead to the successful reheating, radiation and matter dominant universe. In the following subsections, we consider the effect of nonminimal coupling in the dynamics of power-law inflation. ### A Case of $`\xi >0`$ In this case, the inflaton field is required to lie in the region of $`(\text{5})`$. An important point with the existence of positive $`\xi `$ is that an effective potential $`(\text{11})`$ in the Einstein frame has a local minimum at $`\varphi _{}=\left(\sqrt{{\displaystyle \frac{p}{4\pi }}+{\displaystyle \frac{1}{8\pi \xi }}}\sqrt{{\displaystyle \frac{p}{4\pi }}}\right)m_{\mathrm{pl}}.`$ (24) Note that $`\varphi _{}`$ exists in the range of $`0<\varphi _{}<\varphi _c`$. The evolution of the scale factor depends on the initial condition of the inflaton field. When $`\varphi `$ is close to the critical value $`\pm \varphi _c`$, $`\widehat{V}`$ is approximately written as $`\widehat{V}(\mathrm{\Phi })A\mathrm{exp}\left(2\sqrt{{\displaystyle \frac{2}{3}}}\kappa \mathrm{\Phi }\right),`$ (25) where $`A`$ is a constant. This exponential potential has a power-law solution which is described by $`\widehat{a}\widehat{t}^{3/4}.`$ (26) Going back to the original frame, the scale factor evolves as $`at^{1/2}.`$ (27) This means that we do not have an inflationary solution when $`\varphi `$ is close to $`\pm \varphi _c`$. Let us consider the case where the initial value of inflaton ($`=\varphi _i`$) is in the range of $`\varphi _c<\varphi _i<\varphi _{}`$. Since the slope of the effective potential $`\widehat{V}(\varphi )`$ becomes steeper than in the case of $`\xi =0`$ for the values of $`\varphi <0`$ due to the existence of $`1/(1\xi \kappa ^2\varphi ^2)^2`$ term (See Fig. 1), nonminimal coupling does not support inflation when $`\varphi `$ rolls down in the region of $`\varphi <0`$. However, after inflaton passes through $`\varphi =0`$, we have a flatter effective potential compared with the case of $`\xi =0`$. We can expect that the universe evolves as inflationary even for the value of $`p1`$. For example, consider the case of $`p=1/2`$ and $`\xi =0.05`$ with the initial value $`\varphi _i`$ close to $`\varphi _c=0.892m_{\mathrm{pl}}`$. In Fig. 2, we show the evolution of the scale factor $`a`$ as a function of $`t`$. The evolution of the inflaton field $`\varphi `$ is also depicted in Fig. 3. Note that these variables are those in the original frame. We find that inflation does not take place in the region of $`\varphi <0`$. At the time of $`\overline{t}m_{\mathrm{pl}}t0.1`$, inflaton reaches $`\varphi =0`$. After that, the assisted mechanism due to nonminimal coupling begins to work, and the universe evolves as inflationary after $`\overline{t}1.5`$, which corresponds to the value of $`\varphi 0.5m_{\mathrm{pl}}`$. Even in the case of $`0<\varphi <0.5m_{\mathrm{pl}}`$, the $`\xi `$ effect makes the universe grow more rapidly than in the case of $`\xi =0`$. Assisted inflation relevantly occurs in the region of $`0.5m_{\mathrm{pl}}\stackrel{<}{}\varphi \stackrel{<}{}\varphi _{}=0.714m_{\mathrm{pl}}`$. Inflaton slowly evolves in the flat region around $`\varphi =\varphi _{}`$, and finally approaches the local minimum. In the case where inflaton is initially located in the region of $`0<\varphi _i<\varphi _{}`$, the assisted mechanism works from the beginning. After the field reaches the local minimum at $`\varphi =\varphi _{}`$, it continues to stay there, and the inflationary phase does not terminate. Hence we need some exit mechanisms from inflation as in the minimally coupled case. Let us next examine the case where $`\varphi `$ is initially in the range of $`\varphi _{}<\varphi _i<\varphi _c`$. If $`\varphi _i`$ is close to the critical value $`\varphi _c`$, we have the non-inflationary solution $`(\text{27})`$ in any value of $`p`$ and $`\xi `$. However, as $`\varphi `$ approaches the value of $`\varphi _{}`$, the assisted mechanism begins to work. Let us consider the case of $`p=1/2`$ and $`\xi =0.05`$ with the initial value of $`\varphi `$ close to $`\varphi _c=0.892m_{\mathrm{pl}}`$. We find in Fig. 2 the inflationary behavior after $`\overline{t}5`$, although the universe evolves deceleratedly as Eq. $`(\text{27})`$ at the initial stage. At $`\overline{t}=5`$, the value of inflaton is $`\varphi =0.880m_{\mathrm{pl}}`$ (See Fig. 3), which is close to the critical value $`\varphi _c`$. This means that we have the non-inflationary solution $`(\text{27})`$ only when $`\varphi `$ is very close to $`\varphi _c`$. Nonminimal coupling drives inflation while inflaton slowly evolves in the flat region of $`\varphi _{}<\varphi \stackrel{<}{}0.880m_{\mathrm{pl}}`$. Finally, inflaton reaches the local minimum of the effective potential, as is the same with the case of $`\varphi _c<\varphi _i<\varphi _{}`$. Next, we investigate the case where $`p`$ and $`\xi `$ are changed. As is found by Eq. $`(\text{24})`$, $`\varphi _{}`$ decreases with the increase of $`\xi `$. Since the effective potential $`\widehat{V}(\varphi )`$ becomes flatter in the region of $`0<\varphi <\varphi _{}`$ as $`\xi `$ increases, inflation is more easily realized in this region in spite of the decrease of $`\varphi _{}`$. In the case of $`\varphi _{}<\varphi <\varphi _c`$, since the flat region around the local minimum becomes wider for larger value of $`\xi `$, assisted inflation occurs significantly except the case where $`\varphi `$ is close to $`\varphi _c`$. This suggested that assisted mechanism works more efficiently with the increase of $`\xi `$ in the region of $`\varphi >0`$, and we have numerically confirmed this fact. If we choose the values of $`p`$ with $`p>1`$, inflation is always supported by the effect of positive $`\xi `$ when inflaton evolves in the region of $`\varphi >0`$ as long as $`\varphi `$ is not close to $`\varphi _c`$. On the other hand, since the slope of the effective potential becomes steeper with the decrease of $`p`$, we do not necessarily have inflationary solutions for the case of $`p1`$. In the case where nonminimal coupling makes the effective potential flatter than in the case of $`p=1`$ and $`\xi =0`$, assisted inflation occurs in the region of $`\varphi >0`$. For example, for $`p=2/3`$ and $`p=1/2`$ cases, the coupling $`\xi `$ is required to be $`\xi \stackrel{>}{}3\times 10^3`$ and $`\xi \stackrel{>}{}7\times 10^3`$, respectively, to lead to assisted inflation while $`\varphi `$ rolls down toward $`\varphi _{}`$. Although it is rather difficult to obtain the sufficient number of $`e`$-foldings to solve the cosmological puzzles only by the effect of nonminimal coupling in the case of $`p1`$, it may be possible to lead to sufficient inflation in the multi-field case. What we emphasize is that inflation is realized even in the one-field case with $`p1`$ in taking into account the effect of nonminimal coupling. ### B Case of $`\xi <0`$ When $`\xi `$ is negative, the shape of the effective potential $`\widehat{V}(\varphi )`$ is different depending on the relation of $`\xi `$ and $`p`$. In the case of $`|\xi |<1/2p`$, the potential does not have any extrema. However, when $`|\xi |>1/2p`$, it has both of a local minimum and a local maximum. In what follows, we examine these two different cases separately. #### 1 Case of $`|\xi |<1/2p`$ In this case, $`\widehat{V}(\varphi )`$ decreases monotonically with the increase of $`\varphi `$. Although the potential is flatter than in the case of $`\xi =0`$ for negative values of $`\varphi `$, it is steeper for $`\varphi >0`$ (See Fig. 1). Hence assisted inflation does not take place when $`\varphi `$ is initially located in the region of $`\varphi >0`$. However, for the initial values of $`\varphi <0`$, we can expect inflation to occur even in the case of $`p1`$. For example, let us consider the case of $`p=3/4`$ with the initial value of $`\varphi _i=2m_{\mathrm{pl}}`$. Numerical calculations show that we have an inflationary solution at the initial stage for the coupling of $`|\xi |\stackrel{>}{}0.3`$. We depict in Fig. 4 the evolution of the scale factor $`a`$ as a function of $`t`$ for two cases of $`\xi =0.5`$ and $`\xi =0.2`$. Although the universe evolves as non-inflationary at the whole stage in the $`\xi =0.2`$ case, inflation occurs in the $`\xi =0.5`$ case at the initial stage where $`\varphi `$ is negative. The larger values of $`|\xi |`$ ($`\stackrel{>}{}0.3`$) lead to the larger amount of inflation. If $`\varphi `$ is initially smaller than the value of $`\varphi _i=2m_{\mathrm{pl}}`$, one may consider that we will obtain the larger amount of $`e`$-foldings. However, this is not the case. With the decrease of $`\varphi `$, since the $`\xi \kappa ^2\varphi ^2`$ term in Eq. $`(\text{19})`$ increases and becomes much larger than unity, the velocity of the $`\varphi `$ field is larger than that of the $`\mathrm{\Phi }`$ field. This suggests that the $`\varphi `$ field does not evolve slowly enough to drive inflation for the smaller values of $`\varphi `$. For example, in the case of $`p=3/4`$ and $`\xi =0.5`$, we have numerically found that the universe evolves as non-inflationary for the values of $`\varphi _i\stackrel{<}{}4m_{\mathrm{pl}}`$. When $`p`$ is less than unity, inflation is not realized unless we choose rather large values of $`|\xi |`$. For the case of $`p=1/2`$ with the initial value of $`\varphi _i=2m_{\mathrm{pl}}`$, we require the values of $`|\xi |\stackrel{>}{}0.6`$ for inflation to occur. However, we should stress that assisted inflation is possible in the case of $`p1`$ for the larger values of $`|\xi |`$ which satisfy $`|\xi |<1/2p`$. #### 2 Case of $`|\xi |>1/2p`$ In this case, the potential $`\widehat{V}(\varphi )`$ is a local maximum at $`\varphi _1=\left(\sqrt{{\displaystyle \frac{p}{4\pi }}}\sqrt{{\displaystyle \frac{p}{4\pi }}+{\displaystyle \frac{1}{8\pi \xi }}}\right)m_{\mathrm{pl}},`$ (28) and a local minimum at $`\varphi _2=\left(\sqrt{{\displaystyle \frac{p}{4\pi }}}+\sqrt{{\displaystyle \frac{p}{4\pi }}+{\displaystyle \frac{1}{8\pi \xi }}}\right)m_{\mathrm{pl}}.`$ (29) Note that both of $`\varphi _1`$ and $`\varphi _2`$ are negative. With the existence of a local minimum, we can expect assisted inflation to occur in the region of $`\varphi <\varphi _1`$. We consider the case of $`p=1`$ and $`\xi =1`$ as one example. In this case, $`\varphi _1`$ and $`\varphi _2`$ correspond to $`\varphi _1=0.083m_{\mathrm{pl}}`$ and $`\varphi _2=0.482m_{\mathrm{pl}}`$, respectively. When $`\varphi `$ initially lies in $`\varphi _1<\varphi _i<0`$, assisted inflation occurs when inflaton rolls down in the region of $`\varphi <0`$ as in the case of $`|\xi |<1/2p`$. However, once inflaton passes through $`\varphi =0`$, assisted mechanism does not work since the effective potential becomes steeper compared with the $`\xi =0`$ case. When the initial value of inflaton is less than $`\varphi _1`$, the field evolves toward the local minimum at $`\varphi =\varphi _2`$ and is finally trapped in this minimum except the case where $`\varphi _i`$ is much smaller than $`\varphi _2`$. Inflation relevantly occurs as $`\varphi `$ approaches the potential minimum, since $`\widehat{V}(\varphi )`$ becomes gradually flatter. We show in Fig. 5 the evolution of the scale factor in the original frame for two cases of $`\varphi _i=0.090m_{\mathrm{pl}}`$ and $`\varphi _i=3m_{\mathrm{pl}}`$. In both cases, we find that the universe evolves as inflationary. Inflaton is finally trapped in the local minimum at $`\varphi =\varphi _2`$ after $`\overline{t}4`$ for the $`\varphi _i=0.090m_{\mathrm{pl}}`$ case, and after $`\overline{t}1.5`$ for the $`\varphi _i=3m_{\mathrm{pl}}`$ case. On the other hand, when the initial value of inflaton is $`\varphi _i\stackrel{<}{}5m_{\mathrm{pl}}`$, the $`\varphi `$ field moves rather rapidly and goes beyond the local maximum at $`\varphi _1`$. In this case, since the velocity of the $`\varphi `$ field is large due to the relation of Eq. $`(\text{19})`$, we have no inflationary solution. Namely, in the case of $`p=1`$ and $`\xi =1`$, assisted inflation is realized for the initial values of $`5m_{\mathrm{pl}}\stackrel{<}{}\varphi _i<0`$. If $`p`$ is less than unity, we require rather large values of $`|\xi |`$ greater than the order of unity due to the condition of $`|\xi |>1/2p`$. However, it is important to note that nonminimal coupling can lead to inflation even if $`p1`$ in the region of $`\varphi <0`$. We have numerically found that assisted inflation can be realized for the values of $`\xi `$ which satisfy $`|\xi |>1/2p`$. ### C Summary Finally, we present two-dimensional plots of $`\xi `$ and $`p`$ which divide the parameter regions of the inflationary and non-inflationary solutions in Fig. 6. These parameter spaces depend on initial values of inflaton. For $`\varphi _i0`$, inflation can take place even when $`p<1`$ for positive values of $`\xi `$ (see Fig. 6a). In this case, however, negative $`\xi `$ does not lead to assisted inflation because assisted mechanism is absent for $`\varphi >0`$. On the other hand, when inflaton is initially located for $`\varphi _i<0`$, inflationary behavior appears when inflaton evolves in the region of $`\varphi <0`$ even for negative $`\xi `$ (see Fig. 6b). ## IV Concluding remarks and discussions In this paper, we have investigated the dynamics of power-law inflation with a nonminimally coupled inflaton field. We studied how nonminimal coupling affects the dynamics of inflation with an exponential potential $`V(\varphi )=V_0\mathrm{exp}(\sqrt{16\pi /pm_{\mathrm{pl}}^2}\varphi )`$ in the one-field model. In the case of $`\xi >0`$, since an effective potential $`(\text{11})`$ in the Einstein frame which appears by a conformal transformation becomes flatter for $`\varphi >0`$ than in the case of $`\xi =0`$, assisted inflation can be realized by nonminimal coupling. In this case, nonminimal coupling gives rise to a potential minimum at some positive value of $`\varphi `$, and the potential becomes sufficiently flat around this area. Assisted mechanism works except the case where inflaton evolves in the region of $`\varphi <0`$ and $`\varphi `$ is close to the value of $`\varphi _c=1/\sqrt{8\pi \xi }`$. Even when the power $`p`$ is less than unity, we have an inflationary solution by choosing the appropriate values of $`\xi `$. For example, we have numerically found that inflation occurs for $`\xi \stackrel{>}{}3\times 10^3`$ when $`p=2/3`$, and for $`\xi \stackrel{>}{}7\times 10^3`$ when $`p=1/2`$. When $`\xi `$ is negative, the shape of the effective potential $`\widehat{V}(\varphi )`$ is different depending on two cases of $`|\xi |<1/2p`$ and $`|\xi |>1/2p`$. In the former case, $`\widehat{V}(\varphi )`$ is a monotonically decreasing function of $`\varphi `$. Since the potential is flat for $`\varphi <0`$ compared with the case of $`\xi =0`$, assisted mechanism works in this region as long as $`\varphi `$ is not too far from $`\varphi =0`$. However, nonminimal coupling prevents inflation for $`\varphi >0`$. In the latter case, the effective potential has a local maximum at $`\varphi _1`$ and a local minimum at $`\varphi _2`$ with $`\varphi _2<\varphi _1<0`$. When $`\varphi `$ is initially in the range of $`\varphi _1<\varphi <0`$, the assisted dynamics is the same as the former case. When $`\varphi `$ is smaller than $`\varphi _1`$ initially, inflaton evolves toward the potential minimum at $`\varphi _2`$. In this case, nonminimal coupling assists inflation to occur unless $`\varphi `$ is much smaller than $`\varphi _2`$. For summary, we conclude that assisted inflation can be realized for $`\varphi >0`$ in the case of $`\xi >0`$; and for $`\varphi <0`$ in the case of $`\xi <0`$. The higher dimensional Kaluza-Klein theories often give rise to exponential potentials which are obtained by means of a conformal transformation to the Einstein frame. Then there is a possibility that inflaton is minimally coupled to gravity in the effective four-dimensional theories. Although we did not ask the origin of the exponential potential and considered a nonminimally coupled inflaton field in the four-dimensional action as in the chaotic inflation model plus nonminimal coupling, we have to keep in mind that nonminimal coupling may not play relevant roles in realistic models of physics. What we emphasize is that assisted inflation is possible even in the one-field model by introducing nonminimal coupling. When inflaton is minimally coupled to the spacetime curvature in the effective four-dimensional theories, inflation can be assisted by considering multiple scalar fields. The problem of power-law inflation with an exponential potential is that the potential does not have a local minimum which causes a successful reheating. We find that introducing nonminimal coupling gives rise to a local minimum to which inflaton rolls down. However, since inflaton continues to possess a constant energy at this minimum, the universe inflates forever. This ever-inflating problem also occurs in the minimally coupled case. One exit mechanism is to introduce another scalar field as the hybrid inflation model, by which inflaton is to evolve toward a true minimum. Although we do not consider realistic models of exponential potentials which have graceful exit from inflation in this paper, it is of interest whether nonminimal coupling or multiple scalar fields assist inflation or not in such models. These issues are under consideration. ## ACKOWLEDGEMENTS The author would like to thank Bruce A. Bassett, Kei-ichi Maeda, Nobuyuki Sakai, Takashi Torii, and Kohta Yamamoto for useful discussions. This work was supported partially by a Grant-in-Aid for Scientific Research Fund of the Ministry of Education, Science and Culture (No. 09410217) and by the Waseda University Grant for Special Research Projects.
warning/0004/hep-th0004123.html
ar5iv
text
# 1 Introduction ## 1 Introduction In two previous papers together with Peter van Nieuwenhuizen , we showed that there exists a nonlinear embedding of 7d maximal gauged sugra into 11d sugra and proved the consistency of this truncation 11 dimensional fields to the 7 dimensional fields in the $`AdS_7\times S_4`$ background. For the $`AdS_4\times S_7`$ KK reduction of 11d sugra (to maximal d=4 gauged sugra), de Wit and Nicolai proved the consistency of the truncation indirectly (starting from another formulation of 11d sugra, with SU(8) invariance). For the $`AdS_5\times S_5`$ case presumably one can find also a consistent truncation of 10d IIB sugra to 5d maximal gauged sugra. Based on the existence of these consistent truncations we conjectured in that for the computation of correlators via the AdS-CFT correspondence , if we are interested in operators corresponding to gauged sugra fields, it is enough to take the gauged sugra action. This eliminates an ambiguity in the formulation of the correspondence. Let’s explain this further: a priori, there are two ways of dealing with the computation of correlators. The prescription says to take string theory on the $`AdS_p\times S_{Dp}`$ background, and compute the effective action as a function of the boundary fields. One way could be to take the linear KK expansion in spherical harmonics (given in ) $$\varphi _{Ai}(x,y)=\underset{I}{}\varphi _A^I(x)Y_i^I(y)$$ (1.1) and plug it into the 11d sugra action. Then the truncation to the subset of fields of interest, $`\{\varphi _A^{I_0}\}`$ is not consistent in general, because there are terms in the action linear in the fields set to zero, $`\{\varphi _A^{I_n}\}`$. That means that their equation of motion, $`\delta S/\delta \varphi _A^{I_n}=0`$, contains the fields $`\varphi _A^{I_0}(x)`$ as sources, which gives a contradiction. For the AdS-CFT correspondence, the inconsistency implies that for 4- and higher-point functions of $`\varphi _A^{I_0}(x)`$, all $`\varphi _A^{I_n}`$ will contribute through Witten diagrams involving the troublesome couplings: Another possibility appears when we can have a nonlinear ansatz relating the $`\{\varphi _{Ai}(x,y)\}`$ to $`\{\varphi _A^{I_0}(x)\}`$ such that the truncation is consistent (implying in particular that the $`\{\varphi ^{I_n}\}`$ don’t appear in Witten diagrams for $`\{\varphi ^{I_0}\}`$). A priori, we don’t know which one to take. We need a physical principle to decide. In the cases we study, 11d sugra on $`AdS_7\times S_4`$ truncated to 7d gauged sugra and 10d IIB sugra on $`AdS_5\times S_5`$ truncated to 5d gauged sugra, we will argue by examples that it is correct to take the ansatz giving a consistent truncation, and not the linear ansatz. In other words, the gauged sugra action gives the correct CFT correlators, whereas the action coming from the linear ansatz doesn’t. At this moment, it becomes clear what is the sought-for physical principle. Or rather physical principles: gauge symmetry and susy. Indeed, by taking the linearized action for the gauged sugra and imposing gauge invariance and susy (by the Noether procedure) you obtain the gauged sugra action. In fact, this is how 7d (and 5d) gauged sugra were obtained in . One might think that taking the gauged sugra action for the calculation of correlators is the natural thing to do, but this procedure is available only if there exists a consistent truncation. If there would exist an inconsistent truncation to gauged sugra, that would mean that for 4 point correlators one would have to consider the contribution of the whole tower of massive fields. So the procedure one needs in order to obtain the gauged sugra action is to modify the linearized ansatz in such a way that the action one obtains is gauge invariant and susy. This procedure can be easily generalized. The parent action was invariant under local “gauge” transformations, with parameter $`\xi _\mu =\xi ^{AB}(x)V_\mu ^{AB}(y)`$ (where $`V_\mu ^{AB}`$ is a Killing vector). After the “nonlinear redefinition” (by nonlinear redefinition we understand a nonlinear KK ansatz as opposed to a linear one) of the massless fields, this invariance is lost, and so we need a corresponding nonlinear rotation for the massive fields in order to restore it. It is not clear whether this can be done multiplet by multiplet or for the whole tower at once. We conjecture that this nonlinear ansatz, which we get after performing the rotation, is the one needed for the AdS-CFT correspondence. We have described how to obtain the nonlinear ansatz to be used for the AdS-CFT conjecture. One possible objection to this procedure is that a nonlinear redefinition of fields which doesn’t change the quadratic action, like the one from the linearized ansatz, $`\varphi _{Ai}(x,y)=_I\varphi _A^I(x)Y_i^I(y)`$ to the full nonlinear ansatz, will not change the S matrices of fields. This is so in usual field theory, but for the AdS-CFT correspondence there is one important difference: the S matrices are for the sources on the boundary. And boundary terms, which are usually neglected, become important. We will show that with a very simple example, of a $`\lambda \varphi ^3`$ theory in the bulk. Now, to show that taking the gauged sugra is the correct procedure for the AdS-CFT correspondence as opposed to taking the action coming from the linearized ansatz, we will analyze several n-point functions coming from both approaches. We will first analyze in section 2.1 some relevant 3-point functions, listing all the possible ones and discussing in particular the ones involving gauge fields. Then in section 2.2 we will analyze the CS terms and what we can say for the field theory anomalies. Finally, we will discuss the scalar 3-point functions (corresponding to CPOs) from the work of Lee et al , Corrado et al. , and Bastianelli and Zucchini and how the nonlinear rotation they found is needed to obtain a consistent truncation to gauged sugra. We also give arguments on why this is just a Taylor expansion in fluctuations of the full nonlinear rotation. ## 2 3-point functions of gauge fields ### 2.1 General considerations In this section we will make some general remarks about relevant 3-point functions, in particular about gauge fields correlators. 7d gauged sugra Bosonic fields: gauge fields $`B_\alpha ^{AB}`$ with gauge group $`SO(5)_g`$, antisymmetric tensors $`S_{\alpha \beta \gamma ,A}`$, graviton $`e_\alpha ^I`$, scalars $`\mathrm{\Pi }_{A}^{}{}_{}{}^{i}`$ in the coset $`Sl(5,𝐑)/SO(5)_c`$. Bosonic action: $`e^1={\displaystyle \frac{1}{2}}R+{\displaystyle \frac{1}{4}}m^2(T^22T_{ij}T^{ij}){\displaystyle \frac{1}{2}}P_{\alpha ij}P^{\alpha ij}{\displaystyle \frac{1}{4}}(\mathrm{\Pi }_{A}^{}{}_{}{}^{i}\mathrm{\Pi }_{B}^{}{}_{}{}^{j}F_{\alpha \beta }^{AB})^2`$ (2.1) $`+`$ $`{\displaystyle \frac{1}{2}}(\mathrm{\Pi }_{}^{1}{}_{i}{}^{}{}_{}{}^{A}S_{\alpha \beta \gamma ,A})^2+{\displaystyle \frac{1}{48}}me^1ϵ^{\alpha \beta \gamma \delta ϵ\eta \zeta }\delta ^{AB}S_{\alpha \beta \gamma ,A}F_{\delta ϵ\eta \zeta ,B}`$ $`+`$ $`{\displaystyle \frac{ie^1}{16\sqrt{3}}}ϵ^{\alpha \beta \gamma \delta ϵ\eta \zeta }ϵ_{ABCDE}\delta ^{AG}S_{\alpha \beta \gamma ,G}F_{\delta ϵ}^{BC}F_{\eta \zeta }^{DE}`$ $`+`$ $`{\displaystyle \frac{m^1}{8}}e^1\mathrm{\Omega }_5[B]{\displaystyle \frac{m^1}{16}}e^1\mathrm{\Omega }_3[B]`$ The first remark is that gravity will appear in the correct way just because of general coordinate invariance, both in the linearized ansatz and in the nonlinear one. (or rather, the 11d graviton will be nonlinearly redefined – Weyl rescaled – but only by the scalars: $`e_\alpha ^ae_\alpha ^a[detE_\mu ^m]^{1/5}`$). So we will disregard the 3-point functions involving the graviton. Also, the $`(\delta \mathrm{\Pi }_{A}^{}{}_{}{}^{i})^3`$ 3-point function will be analyzed in the last section. The remaining 3-point functions are: -Involving scalars: $`B_\alpha ^{AB}\delta \mathrm{\Pi }_{A}^{}{}_{}{}^{i}_\alpha \delta \mathrm{\Pi }_{B}^{}{}_{}{}^{i}`$, from $`[P_{\alpha ij}]^2`$ term, $`BB\delta \mathrm{\Pi }`$, from the $`dBdB\delta \mathrm{\Pi }`$ piece in $`(\mathrm{\Pi }_{A}^{}{}_{}{}^{i}\mathrm{\Pi }_{B}^{}{}_{}{}^{i}F_{\alpha \beta }^{AB})^2`$ and $`SS\delta \mathrm{\Pi }`$, from the $`[(\mathrm{\Pi }^1)_{i}^{}{}_{}{}^{A}S_{\alpha \beta \gamma ,A}]^2`$ term. -Involving no scalars: $`(B_\alpha ^{AB})^3`$, from the kinetic term, $`2dBBB`$; SSB, from the S kinetic term, i.e. $`\frac{1}{48}me^1ϵ^{\alpha \beta \gamma \delta ϵ\eta \zeta }\delta ^{AB}S_{\alpha \beta \gamma ,A}F_{\delta ϵ\eta \zeta ,B}`$, and BBS, from the term $`\frac{ie^1}{16\sqrt{3}}ϵ^{\alpha \beta \gamma \delta ϵ\eta \zeta }ϵ_{ABCDE}\delta ^{AG}S_{\alpha \beta \gamma ,G}F_{\delta ϵ}^{BC}F_{\eta \zeta }^{DE}`$. Let’s look at the $`B^3`$ term. The calculation of $`AdS`$ space correlators of gauge fields was done in . Let’s see what would happen if we took the linearized ansatz: For the $`B^3`$ term the $`AdS`$ space, the calculation of correlators was done in . In 7d, the $`dBBB`$ term comes in the nonlinear ansatz in part from the kinetic term $`F_{\alpha \beta \mu \nu }^2`$ in d=11, and so this piece will be absent if we take the linearized ansatz. But there is also a piece coming from $`\sqrt{G^{(11)}}R^{(11)}`$, which will remain. So the coefficient of the CFT correlator of 3 R-currents would get modified. Although the CFT has no lagrangean formulation, one can think of making a free field calculation, as it was done for the correlators of stress tensors in . The coefficient would not be fixed, but it can be fixed by taking susy variations on the stress tensor correlator in . One should obtain the result matching the AdS 3-point function in . So the correct result is the one coming from the nonlinear ansatz. Moreover, we clearly see that imposing gauge invariance on $`dBdB`$ we get the usual $`dFdF`$ action, so gauge invariance here is clearly the physical principle needed to modify the linearized ansatz. The same comment applies to the $`BBS`$ correlator: The $`ϵSFF`$ term in the action comes from two sources: the $`ϵ^{\mu \nu \rho \sigma }ϵ^{\alpha _1\mathrm{}\alpha _7}𝒜_{\alpha _1\alpha _2\alpha _3}F_{\mu \nu \rho \sigma }F_{\alpha _4\mathrm{}\alpha _7}`$ and $`ϵ^{\mu \nu \rho \sigma }ϵ^{\alpha _1\mathrm{}\alpha _7}𝒜_{\alpha _1\alpha _2\alpha _3}F_{\mu \nu \alpha _4\alpha _5}F_{\rho \sigma \alpha _6\alpha _7}`$. If we use the linear ansatz, the last term would give the correct piece, $`ϵSBB`$, but the former would not contribute, and so the normalization of the $`SBB`$ correlator would be wrong. Here one would have to compute the $`AdS`$ correlator first, which we leave for future work . 5d gauged sugra Bosonic fields: -ungauged model: gravitons $`e_\mu ^r`$, gauge fields $`A_\mu ^{AB}`$ scalars $`V_{AB}^{ab}`$ (27-bein), global symmetry group $`E_{6(6)}`$, composite symmetry USp(8). -gauged model: $`e_\mu ^r`$, gauge fields $`A_\mu ^{IJ},B_\mu ^{I\alpha }`$, scalars $`V^{IJab},V_{I\alpha }^{ab}`$, gauge group: $`SO(p,6p)\times SL(2,R)`$, composite symmetry USp(8). Under $`27(15,1)(6,12),ABIJI\alpha `$. Bosonic action: $`e^1_{bosonic}`$ $`=`$ $`{\displaystyle \frac{1}{4}}R+{\displaystyle \frac{1}{24}}P_{\mu abcd}P^{\mu abcd}`$ (2.2) $`{\displaystyle \frac{1}{8}}(F_{\mu \nu ab}+B_{\mu \nu ab})^2{\displaystyle \frac{1}{96}}ϵ^{\mu \nu \rho \sigma \tau }\eta _{IJ}ϵ_{\alpha \beta }B_{\mu \nu }^{I\alpha }D_\rho B_{\sigma \tau }^{J\beta }`$ $`+g^2[{\displaystyle \frac{6}{45^2}}T_{ab}^2{\displaystyle \frac{1}{96}}A_{abcd}^2]+{\displaystyle \frac{1}{12}}ϵ^{\mu \nu \rho \sigma \tau }ϵ^{IJKLMN}`$ $`(F_{IJ\mu \nu }F_{KL\rho \sigma }A_{MN\tau }+g\eta ^{PQ}F_{IJ\mu \nu }A_{KL\rho }A_{MP\sigma }A_{QN\tau }`$ $`+{\displaystyle \frac{2}{5}}g^2\eta ^{PQ}\eta ^{RS}A_{IJ\mu }A_{KP\nu }A_{MR\sigma }A_{SN\tau })`$ where in the ungauged model $`\stackrel{~}{V}_{cd}^{AB}_\mu V_{AB}^{ab}=2Q_{\mu [c}^{[a}\delta _{d]}^{b]}+P_\mu ^{ab}_{cd}`$, which becomes in the gauged model $`\stackrel{~}{V}D_\mu ^{}V=2Q_\mu +P_\mu `$, and $`\stackrel{~}{V}=V^1`$, $`D_\mu ^{}`$ is the SO(p, 6-p) covariant derivative, $`Q_\mu `$ is the USp(8) connection, $`D_\mu `$ is the full $`USp(8)\times SO(p,6p)`$\- covariant connection. Also, $`A_{abcd}`$ $``$ $`T_{a[bcd]},T_{ab}=T_{abc}^c`$ $`T_{bcd}^a`$ $``$ $`Y_{becd}^{ae}=(2V^{IKac}\stackrel{~}{V}_{beJK}V_{JK}^{ac}V_{be}^{I\alpha })\eta ^{JL}\stackrel{~}{V}_{cdIL}`$ $`F_{\mu \nu }^{ab}`$ $`=`$ $`V^{IJab}F_{\mu \nu IJ},B_{\mu \nu }^{ab}=V_{I\alpha }^{ab}B_{\mu \nu }^{I\alpha }`$ (2.3) and $`F_{\mu \nu IJ}`$ and $`B_{\mu \nu }^{I\alpha }`$ are the field strengths of $`A_\mu ^{IJ}`$ and $`B_\mu ^{I\alpha }`$, respectively. Bosonic 3-point functions\- except the ones involving the graviton, for the same reasons as in 7d, and the ones involving only scalars, which are treated in the last section. -Involving no scalars: $`AAA`$, from the $`dAAA`$ and $`dAdAA`$ terms in the action, $`BBA`$ from the $`dABB`$ and $`dBdBA`$ term in the action (namely from the $`ϵBDB`$ term), $`BBB`$ from the $`dBBB`$ term, and $`BAA`$ from the $`dBAA`$term. -Involving scalars: $`VVA`$ terms from the $`VVA`$ piece of $`P_{\mu ij}^2`$ ($`A_\mu ^{IJ}`$ coming from $`D_\mu ^{}`$), and $`BBV`$ from $`\frac{1}{96}ϵ^{\mu \nu \rho \sigma \tau }\eta _{IJ}ϵ_{\alpha \beta }B_{\mu \nu }^{I\alpha }D_\rho B_{\sigma \tau }^{J\beta }`$ (V coming from the $`Q_\mu `$ term in $`D_\lambda `$). The $`AAA`$ term was computed in and gives the correct CFT correlators (we should stress once again that the agreement between the AdS and CFT computations holds as long as one uses the gauged sugra interactions). We can easily extend this result to all the 3 point functions of gauge fields ($`BBA`$, $`BAA`$ and $`BBB`$), and all that changes are the coefficients of the terms in the action involving gauge fields, and the combinatorial factors (coming from differentiating with respect to the boundary sources of the gauge fields). On the other hand, if we take the 10d IIB sugra action, $`S_{IIB,10d}`$ $`=`$ $`{\displaystyle \frac{1}{(2\pi )^7\alpha ^{}^4}}{\displaystyle }d^{10}x\{\sqrt{g^{(10)}}e^{2\varphi }(^{(10)}+4|d\varphi |^2{\displaystyle \frac{1}{3}}|H|^2)2|dl|^2`$ (2.4) $`{\displaystyle \frac{1}{3}}|H^{}lH|^2{\displaystyle \frac{1}{60}}|M^+|^2]{\displaystyle \frac{1}{48}}C^+HH^{}\}`$ and plug in the linearized ansatz, we will again miss some terms of the type $`dAAA`$ coming from the kinetic term$`|H^{}lH|^2`$ of the antisymmetric tensors. The same comment applies to the $`dBBB`$ term for the $`BBB`$ correlator, the $`dBAA`$ term for the $`BAA`$ correlator, and $`dABB`$ for the $`ABB`$ correlator. The fact that we don’t know the nonlinear KK embedding it’s not relevant, because we know that if we have a consistent truncation, the prescription we suggest is to use the gauged sugra action. We also know the linearized KK reduction of . So we can say that the correlators obtained from the linearized ansatz will differ from the ones obtained from the nonlinear ansatz, which we know to be correct (i.e. in agreement with $`N=4`$ SYM results). Once again, the nonlinear ansatz is seen to be the correct one to take. ### 2.2 Anomalies 5 dimensions For the relation between the CS term in maximal 5d gauged sugra and the R-current anomaly in 4d N=4 SYM, Witten gave a very elegant argument in his original paper on the AdS-CFT correspondence . The argument goes as follows: If we vary the bulk gauge fields (in 5d) by $`\delta _\mathrm{\Lambda }A_\mu ^a(x)=(D_\mu \mathrm{\Lambda })^a(x)`$ the only nonzero term in the variation of the action $`\delta _\mathrm{\Lambda }S_{cl}[A_\mu ^a(x)[A_i^a\stackrel{}{x}]]`$ will be a boundary term coming from the CS term, $`\delta _\mathrm{\Lambda }S_{cl}`$ $`=`$ $`\delta _\mathrm{\Lambda }S_{CS}={\displaystyle }d^4x\mathrm{\Lambda }^a(\stackrel{}{x})({\displaystyle \frac{ik}{96\pi ^2}})d^{abc}ϵ^{ijkl}_i(A_j^a_kA_l^c`$ (2.5) $`+{\displaystyle \frac{1}{4}}f^{cde}A_j^bA_k^dA_l^e)`$ And the conjecture implies that $`S_{cl}[A_\mu ^a(x)]`$ is equal to $`W[A_i^a(\stackrel{}{x})]`$, the generating functional of connected Green’s functions on the boundary. Since also $`J_i^a(\stackrel{}{x})=\delta W[A]/\delta A_i^a(\stackrel{}{x})`$, we get $$\delta _\mathrm{\Lambda }S_{cl}=\delta _\mathrm{\Lambda }W=d^4x\mathrm{\Lambda }^a(\stackrel{}{x})D_iJ_i(\stackrel{}{x})$$ (2.6) which implies that $$<D_iJ_i(\stackrel{}{x})>=\delta _\mathrm{\Lambda }S_{CS}$$ (2.7) and gives a concrete physical interpretation to the known mathematical fact that the consistent anomaly in n dimensions is obtained by a descent equation from the 2n+1 dimensional CS action. That implies that if one takes the full 1-loop 3-point function of R-currents in SYM and one takes a divergence, it should reproduce the result for the ’Witten diagram’ of 3 gauge fields in AdS, with a divergence taken. (That is because the anomaly is only 1 loop, by the Adler-Bardeen theorem.) It is indeed so, as noted in ; but it also implies a similar result for the 4 point function. In 4d N=4 SYM, the box diagram and its anomalous part are also nonzero. A priori, there seems to be another diagram contributing an anomaly, but it actually gives only a renormalization. (The wavy lines denote gauge vector propagators while the straight lines indicate fermion propagators.) Only by summing all the one loop anomalous diagrams we get a gauge covariant anomaly $$(D^\mu J_\mu )^a=\frac{N^21}{384\pi ^2}id^{abc}ϵ^{\mu \nu \rho \sigma }F_{\mu \nu }^bF_{\rho \sigma }^c$$ (2.8) (the triangle anomaly anomaly alone yields just the $`dAdA`$ term on the right hand side). On the AdS side, there are two diagrams, the 4 point vertex for the CS term and the exchange diagrams, with 3-point vertices coming also from the CS term. The naive expectation, that the AdS 4-point vertex equals the box diagram from field theory, and the AdS exchange diagram equals the diagram with two triangles glued, is wrong, because the AdS exchange diagram gives a genuine contribution, not just a renormalization. However, Witten’s argument tells us that the sum of the AdS diagrams should be equal to the sum of the field theory diagrams, so we don’t need to worry. What would happen if we take the linearized ansatz instead? Even if we don’t know the nonlinear ansatz, we know that using the nonlinear ansatz, the action for the massless sector will be 5d maximal gauged sugra, which has a CS term, and therefore generates a R-current anomaly. On the other hand, using the linear ansatz , and substituting it in the 10d IIB sugra action, we notice that there is no surviving CS term! That is so because the linear ansatz of the 3-form field strengths $`H`$ and $`H^{}`$ does not contain the massless gauge vector fields. So, by using the linear ansatz, one would conclude that there is no R-current anomaly. And again, gauge and supersymmetry invariance tells us that we should take the nonlinear ansatz, because by imposing them both on the linearized action we get the 5d maximal gauged sugra, with a CS term. 7 dimensions Witten’s argument applies equally well in all odd dimensions, relating the anomaly in 2n dimensions to a CS term in a gauged sugra in 2n+1 dimensions. However, the only other example of maximal gauged sugra in 2n+1 dimensions is d=7. Again, by varying $$S_{CS}=\frac{m^1}{8}e^1\mathrm{\Omega }_5[B]\frac{m^1}{16}e^1\mathrm{\Omega }_3[B]$$ (2.9) where $`\mathrm{\Omega }_3[B]`$ and $`\mathrm{\Omega }_5[B]`$ are the Chern-Simons forms for $`B_\alpha ^{AB}`$ (normalized to $`d\mathrm{\Omega }_3[B]=(TrF^2)^2`$ and $`d\mathrm{\Omega }_5[B]=(TrF^4)`$), we should get the chiral anomaly in 6d. But now it is unclear how to compute this anomaly, since the dual 6d theory is a nontrivial (0,2) CFT without a lagrangean formulation. The anomaly means that the $`SO(5)`$ R-symmetry, which is part of the susy algebra, is broken by the fact that correlators of R-currents are anomalous. In 6d, the first anomalous correlator is the 4-point function (corresponding to the 4-point CS coupling $`dBdBdBB`$). And it is easy to see from the nonlinear ansatz that the 4-point CS coupling (in fact, all the CS term!) will be missing for the linearized ansatz (the 7 dimensional CS has terms with at least four fields, $`Tr(dBdBdBB)`$ and $`Tr(dBdB)Tr(dBB)`$, while the 11 dimensional CS $`dF^{(4)}dF^{(4)}A^{(3)}`$ cannot generate them after substituting a linearized ansatz). But we know that the 6d (0,2) CFT should have a chiral anomaly, because it is obtained as the IR limit of the M5-brane theory, which has an anomaly. Moreover, in , a brane calculation of the anomaly was performed, and it was found that the anomaly has the expected functional dependence, i.e. coming by descent formalism from the 7d Chern-Simons term. The only nontrivial aspect is the coefficient in front of this anomaly, which was found to be proportional to $`N^3`$. The same $`N^3`$ dependence is found also from the AdS calculation, because the sugra coupling has this dependence. One would like to understand the $`N^3`$ dependence in a field theory context, because the calculation in uses M theory, as does the AdS-CFT calculation. However, that was not done yet. The free-field calculation in for the stress-tensor correlators (trying to match with the anomaly calculation of on the AdS side), and the free-field calculation in for the R-current anomaly, both impose by hand the $`N^3`$ dependence. In , the calculation was extended to other gauge groups, and in the calculation was related to Witten’s calculation in type IIA string theory, but a real field theory explanation is still lacking. So again, since we want an anomaly in 7d, because we know it should be there by the AdS-CFT correspondence, the nonlinear ansatz is the correct one. As before, we need to impose both susy and gauge invariance on the linearized action obtained by compactification in order to recover the correct result. (The absence of a CS term respects gauge invariance alone.) ## 3 Scalar 3-point functions Let’s start by giving the $`\lambda \varphi ^3`$ example as promised in the introduction. This example emerged in a discussion we had with Fiorenzo Bastianelli. For a $`\lambda \varphi ^3`$ theory in the bulk, a redefinition of fields in the bulk doesn’t change the bulk S matrices, but does the ones computed on the boundary (via the AdS-CFT-type correspondence), because of the presence of a boundary term. Let’s start with the lagrangean $$=\frac{1}{2}(_\mu \varphi )^2+\frac{1}{2}m^2\varphi ^2+\lambda \varphi ^3$$ (3.1) The bulk 3-point function will be equal to $`\lambda `$. Let’s now redefine $`\varphi =\stackrel{~}{\varphi }+a\stackrel{~}{\varphi }^2`$. Then, $``$ $`=`$ $`{\displaystyle \frac{1}{2}}(_\mu \stackrel{~}{\varphi })^2+{\displaystyle \frac{1}{2}}m^2\stackrel{~}{\varphi }^2+\lambda \stackrel{~}{\varphi }^3+2a\stackrel{~}{\varphi }(_\mu \stackrel{~}{\varphi })^2+m^2a\stackrel{~}{\varphi }^3`$ (3.2) $`+2a^2\stackrel{~}{\varphi }^2(_\mu \stackrel{~}{\varphi })^2+{\displaystyle \frac{1}{2}}m^2\stackrel{~}{\varphi }^4+3\lambda a\stackrel{~}{\varphi }^4`$ $`+3\lambda a^2\stackrel{~}{\varphi }^5+\lambda a^3\stackrel{~}{\varphi }^6`$ But $$[2a\stackrel{~}{\varphi }(_\mu \stackrel{~}{\varphi })^2+m^2a\stackrel{~}{\varphi }^3]=a\stackrel{~}{\varphi }^2\mathrm{}\stackrel{~}{\varphi }+m^2a\stackrel{~}{\varphi }^3+2a_\mu (\stackrel{~}{\varphi }^2_\mu \stackrel{~}{\varphi })$$ (3.3) On shell, $`(\mathrm{}m^2)\stackrel{~}{\varphi }=0`$, so on-shell, the 3-point correlator is obtained from $$\lambda \stackrel{~}{\varphi }^3+a\stackrel{~}{\varphi }^2(\mathrm{}m^2)\stackrel{~}{\varphi }=\lambda \stackrel{~}{\varphi }^3$$ (3.4) and therefore the correlator is still equal to $`\lambda `$ (For the 4-point correlator, the calculation is a bit more involved, but the result is the same.). But we see that if we compute $`S_3[\stackrel{~}{\varphi }|_{bd}]`$ to get the 3-point correlator of the boundary theory, it will differ from $`S_3[\varphi |_{bd}]`$ by $`_\mu (\stackrel{~}{\varphi }^2_\mu \stackrel{~}{\varphi })`$. So a nonlinear redefinition of bulk fields, which doesn’t change the masses, does change the boundary correlators. However, the following observation was made in . The extra term $$_M_\mu (\varphi ^2_\mu \varphi )=_M\varphi (x)\varphi (x)_\mu \varphi (x)$$ (3.5) will contribute only a contact term to correlators, because we have $$\frac{\delta S}{\delta \varphi (x)\delta \varphi (y)\delta \varphi (z)}\delta (xy)$$ (3.6) where $`\varphi (x)`$ here lives on the boundary. Still, if we have the correct action from the start, no unwanted contact terms will appear. But we had only contact terms due to the simplicity of the example. The field redefinition used in involves also derivatives, and is of the type $`\varphi =\stackrel{~}{\varphi }+a\stackrel{~}{\varphi }^2+b(_\mu \stackrel{~}{\varphi })^2`$. Then, substituting into (3.1), we get an extra cubic term to be added to (3.2) (and some higher order terms too) $$[2b_\mu \stackrel{~}{\varphi }^\mu ^\nu \stackrel{~}{\varphi }_\nu \stackrel{~}{\varphi }+m^2b\stackrel{~}{\varphi }(_\mu \stackrel{~}{\varphi })^2$$ (3.7) which can be rewritten by partial integration as $$b(_\mu \stackrel{~}{\varphi })^2(m^2\mathrm{})\stackrel{~}{\varphi }b_\nu ((_\mu \stackrel{~}{\varphi })^2_\nu \stackrel{~}{\varphi })$$ (3.8) so again, on shell we get only a boundary term contribution to the 3-point correlator. But this time it is not just a contact term, as it was also noticed in . We will come back to the discussion of at the end of this section. Let’s now turn to the 3-point functions of scalars. 7 dimensions The gauged sugra scalar fields in 7 dimensions are described by a coset element $`\mathrm{\Pi }_{A}^{}{}_{}{}^{i}SL(5,𝐑)/SO(5)_c`$. In the physical gauge, it is symmetric and traceless. In terms of scalar fluctuations, $`\delta \pi _{Ai}`$, we can write: $`\mathrm{\Pi }_{A}^{}{}_{}{}^{i}`$ $`=`$ $`e^{\delta \pi _{Ai}}=[1+\delta \pi +{\displaystyle \frac{\delta \pi ^2}{2}}+{\displaystyle \frac{\delta \pi ^3}{3!}}+\mathrm{}]_{Ai}`$ $`(\mathrm{\Pi }^1)_{i}^{}{}_{}{}^{A}`$ $`=`$ $`e^{\delta \pi _{Ai}}=[1\delta \pi +{\displaystyle \frac{\delta \pi ^2}{2}}{\displaystyle \frac{\delta \pi ^3}{3!}}+\mathrm{}]_{Ai}`$ (3.9) And so $`T_{ij}`$ $`=`$ $`[12\delta \pi +2\delta \pi ^2{\displaystyle \frac{4}{3}}\delta \pi ^3+\mathrm{}]_{ij}`$ $`TrT`$ $`=`$ $`5+2Tr\delta \pi ^2{\displaystyle \frac{4}{3}}Tr\delta \pi ^3+\mathrm{}`$ (3.10) where $`T_{ij}=(\mathrm{\Pi }^1)_{i}^{}{}_{}{}^{A}(\mathrm{\Pi }^1)_{j}^{}{}_{}{}^{A}`$. From $`P_{\alpha ij}=[\mathrm{\Pi }^1_\alpha \mathrm{\Pi }]_{(ij)}`$, we get by expansion $$P_{\alpha ij}P^{\alpha ij}=Tr(_\alpha \delta \pi )^2+0\frac{5}{6}Tr[(\mathrm{}\delta \pi )\delta \pi ^3]\frac{1}{2}Tr[(_\alpha \delta \pi )^2\delta \pi ^3]$$ (3.11) We notice that the cubic terms in $`P_{\alpha ij}P^{\alpha ij}`$ cancel, but the quartic ones don’t. So, the cubic action for the scalars in the 7d gauged sugra is $$\frac{1}{2}P_{\alpha ij}P^{\alpha ij}+\frac{1}{4}(T^22T_{ij}T^{ij})=\frac{1}{2}Tr(_\alpha \delta \pi )^2+Tr\delta \pi ^2+2Tr\delta \pi ^3$$ (3.12) Let us describe the work Corrado et al. and Bastianelli et al. in 7 dimensions for computing correlators of CPOs in the boundary CFT from the scalar fields correlators in AdS space and see what one can learn from this (This procedure was introduced for the first time by Lee et al. for the study of $`AdS_5`$/N=4 SYM correspondence.). Again, if one compactifies the 11d sugra action on $`AdS_7\times S_4`$ as in , one can write an ansatz for the fields as $`G_{\mathrm{\Lambda }\mathrm{\Pi }}`$ $`=`$ $`\underset{\mathrm{\Lambda }\mathrm{\Pi }}{\overset{}{g}}+h_{\mathrm{\Lambda }\mathrm{\Pi }}`$ $`h_{\alpha \beta }`$ $`=`$ $`h_{(\alpha \beta )}^{}+({\displaystyle \frac{h^{}}{7}}{\displaystyle \frac{h_2}{5}})`$ $`h_{\mu \nu }`$ $`=`$ $`h_{(\mu \nu )}+{\displaystyle \frac{h_2}{4}}g_{\mu \nu }`$ $`F^{(4)}`$ $`=`$ $`\stackrel{}{F^{(4)}}+da^{(3)}`$ (3.13) where the notation $`\underset{\mathrm{\Lambda }\mathrm{\Pi }}{\overset{}{g}}`$ means background metric and $`h_{(\mu \nu )}`$ stands for symmetric and traceless. The indices $`\mathrm{\Lambda },\mathrm{}`$ are 11d, $`\alpha ,\mathrm{}`$ are $`AdS_7`$ and $`\mu ,\mathrm{}`$ are $`S_4`$ indices. One decomposes the sugra fields (in the gauge $`\stackrel{\mu }{\stackrel{}{D}}h_{(\mu \nu )}=\stackrel{\mu }{\stackrel{}{D}}h_{\mu \alpha }=\stackrel{\mu }{\stackrel{}{D}}a_{\mu \mathrm{\Lambda }\mathrm{\Pi }}=0)`$ in spherical harmonics: $`h_{\alpha \beta }^{}`$ $`=`$ $`{\displaystyle h_{(\alpha \beta )I}^{}Y^I}`$ $`h_{(\mu \nu )}`$ $`=`$ $`{\displaystyle \varphi ^IY_{(\mu \nu )}^I}`$ $`h^{}`$ $`=`$ $`{\displaystyle h^IY^I}`$ $`h_2`$ $`=`$ $`{\displaystyle h_{2I}Y^I}`$ $`a_{\mu \nu \rho }`$ $`=`$ $`{\displaystyle b^Iϵ_{\mu \nu \rho \sigma }D^\sigma Y_I}`$ (3.14) where $`\mathrm{}_xY^I(x)=k(k+3)Y^I(x)`$. The scalar kinetic term is diagonalized by the eigenvectors $`s^I`$ $`=`$ $`{\displaystyle \frac{k_1}{2k_1+3}}(h_2^I+32/\sqrt{2}(k+3)b^I)`$ $`t^I`$ $`=`$ $`{\displaystyle \frac{k+3}{2k+3}}(h_2^I32/\sqrt{2}kb^I)`$ $`\mathrm{}s^I`$ $`=`$ $`k(k3),k2`$ $`\mathrm{}t^I`$ $`=`$ $`(k+3)(k+6),k0`$ (3.15) Now the cubic action gives the equations of motion $`(\mathrm{}k(k+3))\varphi ^{I_1}`$ $`=`$ $`D_{I_1I_2I_3}^\varphi s^{I_2}s^{I_3}+E_{I_1I_2I_3}^\varphi D^\alpha s^{I_2}D_\alpha s^{I_3}`$ $`+`$ $`F_{I_1I_2I_3}^\varphi D^{(\alpha }D^{\beta )}s^{I_2}D_{(\alpha }D_{\beta )}s^{I_3}+\mathrm{}`$ $`(\mathrm{}k_1(k_13))s^{I_1}`$ $`=`$ $`D_{I_1I_2I_3}^{s\varphi }\varphi ^{I_2}s^{I_3}+E_{I_1I_2I_3}^{s\varphi }D^\alpha \varphi ^{I_2}D_\alpha s^{I_3}`$ (3.16) $`+`$ $`F_{I_1I_2I_3}^{s\varphi }D^{(\alpha }D^{\beta )}\varphi ^{I_2}D_{(\alpha }D_{\beta )}s^{I_3}`$ $`+`$ $`D_{I_1I_2I_3}s^{I_2}s^{I_3}+E_{I_1I_2I_3}D^\alpha s^{I_2}D_\alpha s^{I_3}`$ $`+`$ $`F_{I_1I_2I_3}D^{(\alpha }D^{\beta )}s^{I_2}D_{(\alpha }D_{\beta )}s^{I_3}+\mathrm{}`$ The condition of getting rid of nonlinear terms with derivatives in the equations of motion suggests the rotation: $`\varphi ^{I_1}`$ $`=`$ $`\stackrel{~}{\varphi }^{I_1}+J_{I_1I_2I_3}^\varphi \stackrel{~}{s}^{I_2}\stackrel{~}{s}^{I_3}+L_{I_1I_2I_3}^\varphi D^\alpha \stackrel{~}{s}^{I_2}D_\alpha \stackrel{~}{s}^{I_3}`$ $`s^{I_1}`$ $`=`$ $`\stackrel{~}{s}^{I_1}+J_{I_1I_2I_3}\stackrel{~}{s}^{I_2}\stackrel{~}{s}^{I_3}+L_{I_1I_2I_3}D^\alpha \stackrel{~}{s}^{I_2}D_\alpha \stackrel{~}{s}^{I_3}`$ (3.17) $`+`$ $`J_{I_1I_2I_3}^{s\varphi }\stackrel{~}{\varphi }^{I_2}\stackrel{~}{s}^{I_3}+L_{I_1I_2I_3}^{s\varphi }D^\alpha \stackrel{~}{\varphi }^{I_2}D_\alpha \stackrel{~}{s}^{I_3}`$ In terms of the redefined fields, the equations of motion become $`(\mathrm{}k(k+3))\stackrel{~}{\varphi }^{I_1}`$ $`=`$ $`\lambda _{I_1I_2I_3}^\varphi \stackrel{~}{s}^{I_2}\stackrel{~}{s}^{I_3}`$ $`(\mathrm{}k(k3))\stackrel{~}{s}^{I_1}`$ $`=`$ $`\lambda _{I_2,I_1I_3}^{s\varphi }\stackrel{~}{\varphi }^{I_2}\stackrel{~}{s}^{I_3}+\lambda _{I_1I_2I_3}\stackrel{~}{s}^{I_2}\stackrel{~}{s}^{I_3}`$ (3.18) where $`\lambda _{I_1I_2I_3}^\varphi `$ $`=`$ $`{\displaystyle \frac{9k_2!k_3!\mathrm{\Gamma }(k_1+5/2)2^{3k_163\mathrm{\Sigma }/2}}{\mathrm{\Gamma }(\alpha _1+1)\mathrm{\Gamma }(\alpha _2+1)\mathrm{\Gamma }(\alpha _3+1)\mathrm{\Gamma }(\frac{\mathrm{\Sigma }_5}{2})}}{\displaystyle \frac{(2\alpha _13)\mathrm{\Sigma }(\mathrm{\Sigma }+1)(\mathrm{\Sigma }+3)}{k_2(2k_2+1)k_3(2k_3+1)}}`$ (3.19) $`\alpha _1(\alpha _11)<T^{I_1}C^{I_2}C^{I_3}>`$ $`\lambda _{I_1I_2I_3}`$ $`=`$ $`{\displaystyle \frac{3k_2!k_3!\mathrm{\Gamma }(k_1+5/2)2^{3k_153\mathrm{\Sigma }/2}}{\mathrm{\Gamma }(\alpha _1+1)\mathrm{\Gamma }(\alpha _2+1)\mathrm{\Gamma }(\alpha _3+1)\mathrm{\Gamma }(\frac{\mathrm{\Sigma }+5}{2})}}`$ (3.20) $`{\displaystyle \frac{\mathrm{\Sigma }(\mathrm{\Sigma }2)(\mathrm{\Sigma }^21)(\mathrm{\Sigma }^29)}{(k_11)(2k_1+3)k_2(2k_2+1)k_3(2k_3+1)}}\alpha _1\alpha _2\alpha _3<C^{I_1}C^{I_2}C^{I_3}>`$ and where $`<C^{I_1}C^{I_2}C^{I_3}>=C_{i_1\mathrm{}i_{\alpha _2+\alpha _3}}^{I_1}C_{}^{I_2i_1\mathrm{}i_{\alpha _3}}{}_{j_1\mathrm{}j_{\alpha _1}}{}^{}C^{I_3i_{\alpha _3+1}\mathrm{}i_{\alpha _3+\alpha _2}j_1\mathrm{}j_{\alpha _1}}`$ and $`<T^{I_1}C^{I_2}C^{I_3}>=T^{I_1abi_1\mathrm{}i_{\alpha _2+\alpha _3}}C_{ai_1\mathrm{}i_{\alpha _3}j_a..j_{\alpha _11}}^{I_2}C_{bi_{\alpha _3+1}\mathrm{}i_{\alpha _3+\alpha _2}}^{I_3}{}_{}{}^{j_1\mathrm{}j_{\alpha _11}}`$, and $`\lambda ^{s\varphi }`$ is proportional to $`\lambda ^\varphi `$. We used the notation: $`\alpha _i`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{j=1}{\overset{3}{}}}k_jk_i`$ (3.21) $`\mathrm{\Sigma }`$ $`=`$ $`k_1+k_2+k_3`$ (3.22) The expression $`<T^{I_1}C^{I_2}C^{I_3}>`$ is non-zero if the ’modified triangle inequalities’ are satisfied: $`\alpha _11,\alpha _20,\alpha _30`$, together with the condition that $`\mathrm{\Sigma }`$ be even. For the s-s-s vertex, if $`k_2=k_3=2`$, we have two possibilities: $`k_1=2,4`$ . The only massive coupling ($`k_1=4`$) is extremal, and it vanishes due to the factor of $`\alpha _1=0`$, (this fact also signals the possibility of a consistent truncation to the massless sector) but the corresponding CFT correlator is finite, because it is obtained after multiplying with a factor of $$\frac{\mathrm{\Gamma }(2\alpha _1)\mathrm{\Gamma }(2\alpha _2)\mathrm{\Gamma }(2\alpha _3)}{\mathrm{\Gamma }(2k_13)\mathrm{\Gamma }(2k_23)\mathrm{\Gamma }(2k_33)}$$ (3.23) and we use that $`\mathrm{\Gamma }(2\alpha _1)/\mathrm{\Gamma }(\alpha _1)1/2`$ when $`\alpha _10`$. This analytical continuation procedure was discussed by Liu and Tseytlin who noticed that although the coupling dilaton-dilaton-massive singlet ($`M^2=32`$) vanishes, the 3-point function of associated CPOs does not. For the $`\varphi ss`$ vertex, the rescaling factor is finite, namely $$\frac{\mathrm{\Gamma }(2\alpha _3+3)\mathrm{\Gamma }(2\alpha _13)\mathrm{\Gamma }(2\alpha _2+3)}{\mathrm{\Gamma }(2k_2)\mathrm{\Gamma }(2k_3)\mathrm{\Gamma }(2k_1+6)}$$ (3.24) so the CFT correlator computed from $`\lambda _{222}^\varphi `$ remains zero. We will argue that the nonlinear field redefinition (3.17) is not just a matter of conveniently getting rid of unwanted higher-derivative terms in the scalar field action, but it is precisely (when truncated to the massless sector) a Taylor expansion of the nonlinear KK ansatz in the transverse fluctuation gauge. So, in this respect, it is not an unnatural redefinition, but it is the one which gives the gauged sugra action. For instance a nonlinear redefinition of the 3-index antisymmetric tensor $`a_{\alpha \beta \gamma }a_{\alpha \beta \gamma }+ϵ_{ABCDE}F_{[\alpha \beta }^{AB}B_{\gamma ]}^{CD}Y^E+more`$ will generate part of the 7d CS terms, which we previously argued that are absent when one uses the linearized ansatz. Start with the nonlinear KK metric ansatz $`G_{\alpha \beta }(x,y)`$ $`=`$ $`\mathrm{\Delta }^{2/5}(x,y)g_{\alpha \beta }(y)`$ (3.25) $`G_{\mu \nu }(x,y)`$ $`=`$ $`\mathrm{\Delta }^{4/5}(x,y)C_\mu ^A(x)T_{AB}^1(y)C_\nu ^B(x)`$ (3.26) $`G_{\mu \alpha }`$ $`=`$ $`2\mathrm{\Delta }^{4/5}(x,y)B_\alpha ^{AB}(y)Y^B(x)C_\mu ^C(x)T_{AC}^1(y)`$ (3.27) where $`\mathrm{\Delta }^{6/5}=YTY`$, and the spherical harmonic satisfy the following identities: $`\mathrm{}_xY^A(x)=4Y^A(x)`$, $`Y(x)Y(x)=_{A=1}^5Y^AY^A=1`$, $`_\mu Y^A(x)=C_\mu ^A(x)`$ is the conformal Killing vector and $`C_\mu ^AC_{}^{\mu }{}_{}{}^{B}=\delta ^{AB}Y^AY^B`$; indices are raised and lowered with Kronecker delta. Set the gauge fields to zero and expand in linear order in the scalar fluctuations $`\delta \pi _{AB}`$. Then $$h_{\mu \nu }=2C_\mu \delta \pi C_\nu +\frac{4}{3}Y\delta \pi Y\underset{\mu \nu }{\overset{}{g}}$$ (3.28) will not be in the transverse gauge, and we need a compensating Einstein transformation $`\stackrel{~}{g}=\frac{1}{n!}(L_\xi )^ng`$ where $`L_\xi `$ is the Lie derivative along $`\xi `$, and $`\stackrel{~}{g}`$ is the transformed metric with parameter $`\xi _\nu =C_\nu \delta \pi Y`$ to satisfy the gauge condition $`\stackrel{\mu }{\stackrel{}{D}}h_{(\mu \nu )}=0`$. Thus, in the transverse gauge, and up to quadratic order in fluctuations $`h_{(\mu \nu )}=0`$. The gauge condition $`\stackrel{\mu }{\stackrel{}{D}}h_{\mu \alpha }=0`$ also implies the need of another Einstein transformation with parameter $`\xi _\alpha =1/2Y\underset{\alpha }{\overset{}{D}}\delta \pi Y`$ and one gets $`h_{\mu \alpha }=0`$, after performing the Einstein transformations. <sup>§</sup><sup>§</sup>§ The constraint $`(h^I9/10h_2^I)\underset{(\alpha }{\overset{}{D}}\underset{\beta )}{\overset{}{D}}Y^I=0`$ is trivially satisfied on-shell if we restrict ourselves to the massless scalar sector, as it yields the scalar field eq. $`Y(\mathrm{}\delta \pi +2\delta \pi )Y=0`$. To second order in fluctuations, and in the transverse gauge we have $`h_{(\mu \nu )}`$ $`=`$ $`{\displaystyle \frac{4}{9}}C_{(\mu }\delta \pi ^2C_{\nu )}{\displaystyle \frac{4}{3}}C_{(\mu }\delta \pi C_{\nu )}Y\delta \pi Y+{\displaystyle \frac{4}{3}}C_{(\mu }\delta \pi YC_{\nu )}\delta \pi Y`$ (3.29) $`+`$ $`{\displaystyle \frac{1}{9}}C_{(\mu }\underset{\alpha }{\overset{}{D}}\delta \pi \stackrel{\alpha }{\stackrel{}{D}}\delta \pi C_{\nu )}+{\displaystyle \frac{1}{3}}C_{(\mu }\underset{\alpha }{\overset{}{D}}\delta \pi C_{\nu )}Y\stackrel{\alpha }{\stackrel{}{D}}\delta \pi Y`$ $``$ $`{\displaystyle \frac{1}{3}}C_{(\mu }\underset{\alpha }{\overset{}{D}}\delta \pi YC_{\nu )}\stackrel{\alpha }{\stackrel{}{D}}\delta \pi Y`$ where, again we needed a compensating Einstein transformation with parameter $`\stackrel{~}{\xi }_\nu `$ $`=`$ $`{\displaystyle \frac{20}{9}}C_\nu (\delta \pi )^2Y{\displaystyle \frac{1}{12}}Y\underset{\alpha }{\overset{}{D}}\delta \pi YC_\nu \stackrel{\alpha }{\stackrel{}{D}}\delta \pi Y+{\displaystyle \frac{1}{18}}Y\underset{\alpha }{\overset{}{D}}\delta \pi \stackrel{\alpha }{\stackrel{}{D}}\delta \pi C_\nu `$ Using now that the first massive mode in $`h_{\mu \nu }(y,x)=\varphi ^I(y)Y_I(x)`$ has a spherical harmonic $`Y_{\mu \nu }^{ABCD}`$ $`=`$ $`C_{(\mu }^{(A}C_{\nu )}^{B)}Y^{(C}Y^{D)}{\displaystyle \frac{1}{2}}C_{(\mu }^{(C}C_{\nu )}^{B)}Y^{(A}Y^{D)}{\displaystyle \frac{1}{2}}C_{(\mu }^{(C}C_{\nu )}^{A)}Y^{(B}Y^{D)}`$ (3.31) $``$ $`{\displaystyle \frac{1}{2}}C_{(\mu }^{(A}C_{\nu )}^{D)}Y^{(B}Y^{C)}{\displaystyle \frac{1}{2}}C_{(\mu }^{(B}C_{\nu )}^{D)}Y^{(A}Y^{C)}+C_{(\mu }^{(D}C_{\nu )}^{C)}Y^{(A}Y^{B)}`$ $``$ $`{\displaystyle \frac{2}{15}}\left(\delta ^{AB}C_{(\mu }^{(D}C_{\nu )}^{C)}+\delta ^{CD}C_{(\mu }^{(A}C_{\nu )}^{B)}\delta ^{AD}C_{(\mu }^{(C}C_{\nu )}^{B)}\delta ^{BC}C_{(\mu }^{(A}C_{\nu )}^{D)}\right)`$ where the symmetry in the $`\{ABCD\}`$ indices is given by the box Young tableau (therefore it is traceless in any pair of indices) and $`\mathrm{}_xY_{\mu \nu }^{ABCD}=8Y_{\mu \nu }^{ABCD}`$, we notice that (3.29) can be rewritten as $`h_{(\mu \nu )}`$ $`=`$ $`{\displaystyle \frac{1}{6}}(4\delta \pi ^{AB}\delta \pi ^{CD}+\underset{\alpha }{\overset{}{D}}\delta \pi ^{AB}\stackrel{\alpha }{\stackrel{}{D}}\delta \pi ^{CD})Y_{\mu \nu }^{ABCD}`$ (3.32) For the same massive mode the nonlinear redefinition reads: $$0=\varphi ^{ABCD}+const.(\underset{\alpha }{\overset{}{D}}s^{AB}\stackrel{\alpha }{\stackrel{}{D}}s^{CD}+4R^2s^{AB}s^{CD})$$ (3.33) where $`R`$ is the $`S_4`$ radius (for us, $`R=1`$, while in $`R=1/2`$). Thus we explicitly showed that the nonlinear redefinition coincides with the nonlinear KK ansatz in the transverse gauge. One cannot directly read off from here the relationship between $`\delta \pi ^{AB}`$ and $`s^{AB}`$, one of the reasons being that the spherical harmonics were normalized differently in and in . In conclusion, since after the nonlinear rotation we get a consistent truncation, and moreover, we get the correct gauged sugra terms (up to cubic order in fluctuations), we can say that the nonlinear ansatz in is the correct one to use in the AdS-CFT correspondence. Let us now discuss the 5-dimensional case. In fact, it is a characteristic of gauged sugras that the kinetic term is $`P_{\alpha ij}P^{\alpha ij}`$, $`P_{\alpha ij}=[\mathrm{\Pi }^1_\alpha \mathrm{\Pi }]_{(ij)}`$, with $`\mathrm{\Pi }`$ the scalar coset vielbein. It follows that the kinetic term has always two derivatives, and we also find no cubic term in $`P_{\alpha ij}^2`$. Now, we would like to see that the gauged sugra action is the correct one to use for the AdS-CFT correspondence. Lee et al. looked at the 10d IIB action compactified on $`AdS_5\times S_5`$. The fluctuations are written as: $`G_{mn}`$ $`=`$ $`g_{mn}+h_{mn}`$ $`h_{\alpha \beta }`$ $`=`$ $`h_{(\alpha \beta )}+{\displaystyle \frac{h_2}{5}}`$ $`h_{\mu \nu }`$ $`=`$ $`h_{(\mu \nu )}^{}{\displaystyle \frac{h_2}{3}}g_{\mu \nu }+{\displaystyle \frac{h^{}}{5}}g_{\mu \nu }`$ $`F`$ $`=`$ $`\overline{F}+5_{[i}a_{jklm]}`$ (3.34) If one decomposes linearly in spherical harmonics as: $`h_{\mu \nu }^{}`$ $`=`$ $`{\displaystyle h_{\mu \nu I}^{}Y^I}`$ $`h_2`$ $`=`$ $`{\displaystyle h_{2I}Y^I}`$ $`a_{\alpha _1\alpha _2\alpha _3\alpha _4}`$ $`=`$ $`{\displaystyle ^\alpha ϵ_{\alpha \alpha _1\alpha _2\alpha _3\alpha _4}b_IY^I}`$ $`a_{\mu _1\mu _2\mu _3\mu _4}`$ $`=`$ $`{\displaystyle a_{\mu _1\mu _2\mu _3\mu _4I}Y^I}`$ (3.35) the constraints on the fields can be solved and the $`h_2^I,B^I`$ system is diagonalized by $`s^I`$ $`=`$ $`{\displaystyle \frac{1}{20(k+2)}}[h_2^I10(k+4)b^I]`$ $`t^I`$ $`=`$ $`{\displaystyle \frac{1}{20(k+2)}}[h_2^I+10kb^I]`$ (3.36) such that $`\mathrm{}s^I`$ $`=`$ $`k(k4)s^I,k2`$ $`\mathrm{}t^I`$ $`=`$ $`(k+4)(k+8)t^I,k0`$ (3.37) If one compactifies the type IIB action in 10d one obtains the following equations of motion for the $`s`$ fields (up to quadratic order in fluctuations) $`(\mathrm{}m_{I_1}^2)s^{I_1}`$ $`=`$ $`{\displaystyle \underset{I_2,I_3}{}}(D_{I_1I_2I_3}s^{I_2}s^{I_3}+E_{I_1I_2I_3}_\mu s^{I_2}^\mu s^{I_3}`$ (3.38) $`+F_{I_1I_2I_3}^{(\mu }^{\nu )}s^{I_2}_{(\mu }_{\nu )}s^{I_3})`$ where $`D_{I_1I_2I_3},E_{I_1I_2I_3}`$ and $`F_{I_1I_2I_3}`$ are constants depending on $`k_1,k_2,k_3`$. But in order to get rid of the terms in the equations of motion nonlinear in $`s^I`$ and involving derivatives, one needs to make a nonlinear redefinition of fields, $`s^{I_1}`$ $`=`$ $`s^{,I_1}+{\displaystyle }(J_{I_1I_2I_3}s^{,I_2}S^{,I_3}+L_{I_1I_2I_3}^\mu s^{,I_2}_\mu s^{,I_3}`$ $`J_{I_1I_2I_3}`$ $`=`$ $`{\displaystyle \frac{1}{2}}E_{I_1I_2I_3}+{\displaystyle \frac{1}{4}}F_{I_1I_2I_3}(m_{I_1}^2m_{I_2}^2m_{I_3}^28)`$ $`L_{I_1I_2I_3}`$ $`=`$ $`{\displaystyle \frac{1}{2}}F_{I_1I_2I_3}`$ (3.39) which modifies the equations of motion to: $$(\mathrm{}m_{I_1}^2)s^{I_1}=\underset{I_2,I_3}{}\lambda _{I_1I_2I_3}s^{I_2}s^{I_3}$$ (3.40) where $$\lambda _{I_1I_2I_3}=\frac{k_1!k_2!k_3}{\alpha _1!\alpha _2!\alpha _3!}\frac{(k+1)2^{k+2\mathrm{\Sigma }/2}\mathrm{\Sigma }((\mathrm{\Sigma }/2)^21)((\mathrm{\Sigma }/2)^24)}{k_1(k_11)(k_2+1)(k_3+1)(\mathrm{\Sigma }/2+2)!}\alpha _1\alpha _2\alpha _3<C^{I_1}C^{I_2}C^{I_3}>$$ (3.41) We notice that the invariant tensor $`<C^{I_1}C^{I_2}C^{I_3}>`$ is nonzero (one can contract the indices correctly) only if the ’triangle inequalities’ $`\alpha _i0`$ are satisfied, together with the the condition that $`\mathrm{\Sigma }`$ is even. If the coupling $`\lambda _{I_1I_2I_3}`$ corresponds to a massive mode and two massless modes, i.e. $`k_2=k_3=2`$, then $`k_1`$ is restricted to be 2 (massless) or 4 (massive). The latter case is ’extremal’, in the sense that $`\alpha _1`$ takes the extreme value zero). But in the extremal case $`\lambda _{I_1I_2I_3}=0`$ because the $`\alpha _1`$ factor vanishes. The fact that after the nonlinear redefinition one has a consistent truncation, both in 5 and in 7 dimensions, was noticed already in a paper we wrote with Peter van Nieuwenhuizen , but we gave no details there. Afterwards, Aryutunov and Frolov wrote a series of papers where they found similar results. In , the cubic and quartic terms in the action were calculated by expanding the 10d IIB action in scalar fluctuations, and after a nonlinear redefinition of fields they also found a consistent truncation to the massless sector (the calculation also involves more fields, not just the $`s^I`$ scalars), i.e. all couplings between massless scalars and one massive scalar vanish. Moreover, in they find that the corresponding action for the massless scalars (to quartic order) coincides with the terms in the gauged sugra action. After finding the coupling $`\lambda _{I_1I_2I_3}`$, one can use it to compute correlators of CPO’s in the boundary CFT using the formulas in which are found to agree with the weak coupling result. But for that one needed a nonlinear redefinition of fields, that is effectively modifying the linear ansatz in to a nonlinear one. This nonlinear ansatz gives a consistent truncation, because one can consistently put all the massive $`s^I`$ in to zero. We notice that when going from a 3-point coupling in AdS space to a correlator on the boundary, one picks up a factor $$\frac{\mathrm{\Gamma }(\alpha _1)\mathrm{\Gamma }(\alpha _2)\mathrm{\Gamma }(\alpha _3)}{\mathrm{\Gamma }(k_12)\mathrm{\Gamma }(k_22)\mathrm{\Gamma }(k_32)}$$ (3.42) The fact that the denominator becomes infinite is absorbed in the normalization of the operators. The 2-point function of CPOs coupled to the $`k=2`$ scalar fields behaves $`(k2)^2`$, and therefore vanishes. To get a nonvanishing result (in accord with the CFT calculation) one has to rescale the supergravity fields with the infinite factor $`1/(k2)`$. This can be interpreted as another example of analytical continuation for the ‘extremal’ correlator $`k_1=k_2=2`$. We notice that the $`\mathrm{\Gamma }(\alpha _1)`$ in the numerator becomes infinite, so that the ’extremal’ correlator becomes nonzero. We will say more about that at the end of this section. In the $`AdS_5\times S_5`$ case, no fully consistent KK truncation is known, but it is generally believed that one exists. If this is the case, the procedure of taking the consistent truncation is seen to be the correct one. Another point to be stressed is that before the rotation, the action has a term cubic in scalars, but with two derivatives. We have seen that in the gauged sugra action we don’t have such a term, so the fact that the nonlinear rotation removes it is another confirmation of our procedure. Finally, we shall comment on the calculation in . This paper tried to address the following puzzle raised by the calculation in . If one takes the limit when $`k_1k_2+k_3`$ in the calculation of , the coefficient of the cubic action for the scalars tend to zero, but the integration diverges in such a way that the 3-point function becomes zero. To address this issue, D’Hoker et al. study the $`t\varphi \varphi `$ three point function, and instead of using the nonlinear redefinition of fields used in (as we argue that is the correct procedure), use equations of motion and partial integration to arrive at $`2k_5^2S_{cubic}=8{\displaystyle \frac{(\mathrm{\Sigma }+4)\alpha _1(\alpha _2+2)(\alpha _3+2)}{(k_1+3)}}{\displaystyle _{AdS_5}}a(k_1,k_2,k_3)t^{k_1}\varphi ^{k_2}\varphi ^{k_3}+`$ $`{\displaystyle _{(AdS_5)}}{\displaystyle \frac{a(k_1,k_2,k_3)}{k_1+3}}(D_n\varphi ^{k_3}D_\mu \varphi ^{k_1}D^\mu \varphi ^{k_2}D_n\varphi ^{k_2}D_\mu t^{k_1}D^\mu \varphi ^{k_3}`$ $`+D_nt^{k_1}D_{mu}\varphi ^{k_2}D^\mu \varphi ^{k_3})+\mathrm{contact}\mathrm{terms}`$ (3.43) where $`\mathrm{\Sigma }=\frac{1}{2}(k_1+k_2+k_3),\alpha _1=\frac{1}{2}(k_2+k_3k_1)`$, etc., obtaining what we described at the beginning of this section, namely that the difference between making a nonlinear redefinition of fields and using equations of motion and partial integrations is given by boundary terms. The boundary terms of the type in (3.3) are contact terms which were dropped, and the boundary terms in (3.8) are of the same type as the ones in (3.43). We indeed notice that the coefficient of the bulk integral in (3.43) becomes equal to zero for $`k_1=k_2+k_3`$. The point of view adopted in is the following. At $`k_1<k_2+k_3`$ only the bulk integral contributes, and the boundary one doesn’t. But at $`k_1=k_2+k_3`$, the situation is reversed: only the boundary integral contributes, and the boundary one doesn’t. Moreover, the result for $`k_1=k_2+k_3`$ coincides with the one from the limit $`k_1k_2+k_3`$. Our point of view is that we need to start with only the bulk integral in (3.43) (in other words make the nonlinear redefinition of fields). The analytic continuation $`k_1k_2+k_3`$ gives the correct result. With the linearized ansatz one also gets the boundary integral. If one considers it to be nonzero as does, then one can only spoil the result by a factor of 2, in this example. We note that for this case of scalar fields, this boundary terms seem to contribute only for ’extremal correlators’ ($`k_1=k_2+k_3`$), with a singular limit needed to be taken, but for general fields (gauge fields, for instance) the same will probably not happen. Indeed, as an example, for gauge fields we saw that the Chern-Simons term is completely missed by the linearized ansatz. A nonlinear redefinition in 7d which would give it would have to involve $`ϵ_{\alpha _1\mathrm{}\alpha _7}`$. And for such a redefinition the $`A_\mu `$ equation of motion ($`(\mathrm{}\delta _{\mu \nu }_\mu _\nu )A_\mu =0`$ ) is not very useful either in terms of creating the wanted Chern-Simons term by partial integration and use of the equations of motion. Neither the natural redefinition one would think of, $`A_{\alpha _1}A_{\alpha _1}+ϵ_{\alpha _1\mathrm{}\alpha _7}^{\alpha _2}A^{\alpha _3}^{\alpha _4}A^{\alpha _5}^{\alpha _6}A^{\alpha _7}`$, nor any other combination is able to reproduce the required Chern-Simons term. To be explicit, the nonlinear redefinition has to involve massive fields being redefined too: massive $``$ massive +(massless)<sup>n</sup>. Similarly, for the use of equations of motion one needs to use the massive equations of motion to obtain a Chern-Simons term. But if one takes as a starting point the linearized ansatz, one will obtain no anomalous CFT correlators. That is because for that one needs an $`ϵ`$ symbol. In 5 dimensions, the anomaly should be in the 3 point function already, but clearly the 3 point vertex is non-anomalous (because the $`ϵ`$ symbol comes only from the Chern-Simons which is now absent). In 7 dimensions, the anomaly starts at the 4 point function. So even though there is no Chern-Simons term, one might hope that the exchange diagram, where massive fields are exchanged, could contribute. But using the linearized ansatz we don’t get any couplings involving the $`ϵ`$ symbol of the type gauge field-gauge field-massive field. Therefore again, no anomaly on the CFT side. From this discussion, one concludes that, even if one does the steps in , namely partial integration and the use of the bulk equations of motion, if one obtains the Chern-Simons term, it will be together with extra boundary terms canceling the effect of the anomaly in the boundary correlators! So we have an example where, even if the methods of can be applied, the boundary terms which are generated will contribute not only to ’extremal’ correlators, but to the ’massless’ ones as well. Therefore, here (for the scalar field case) it is somewhat a matter of taste which philosophy one takes, maybe the one of D’Hoker et al. looks more attractive, however one has to consider a more general case. As we have seen, we have an argument that taking the nonlinear ansatz from the start produces the correct anomaly, the correct gauge invariant AdS correlators (so correct R-invariant CFT correlators) and gets rid of unwanted contact terms. Therefore, it is better to have a clear physical principle to deal with all of the above. For the ’massless’ (gauged sugra) AdS fields, we know that gauge invariance and susy forces us to take the gauged sugra action. For the massive fields, we don’t know the nonlinear ansatz. But then we can do what Seiberg et al. and Corrado et al. did, namely to find order by order and ansatz which removes unwanted terms in the action (with too many derivatives, for instance). Ideally, one would have to do what we sketched in the introduction, namely to use gauge invariance and susy to fix the nonlinear ansatz for the tower of massive fields. ## 4 Conclusions In this paper, we gave arguments for the previous conjecture that one has to use the gauged sugra action, as obtained from 11d, or 10d IIB sugra through a nonlinear KK ansatz, for the computation of n-point functions of CFT operators coupled to the massless (sugra) sector. A similar nonlinear rotation is needed for the massive tower in order to restore the gauge invariance of the action for the whole KK tower after performing the “nonlinear rotation” at the massless level. Part of this rotation, namely up to quadratic order in scalar fluctuations was already introduced by Lee et al. , Corrado et al. and Bastianelli and Zucchini . But their reason for doing this was to eliminate certain higher derivative couplings from the reduced action. Our arguments are based on: \- previously computed R-current correlators. In particular, we noticed that the linear KK ansatz completely misses the CS term (in both 5 and 7 dimensions) which corresponds to the R-current anomaly. -we explicitly showed that the nonlinear rotation of corresponds to a Taylor expansion of the nonlinear KK ansatz (in the transverse gauge) in massless scalar fluctuations. Acknowledgements We would like to thank Peter van Nieuwenhuizen, Fiorenzo Bastianelli, Iosif Bena, Calin Lazaroiu, Radu Roiban and Arkady Tseytlin for valuable discusions.
warning/0004/hep-ph0004143.html
ar5iv
text
# 1 Introduction ## 1 Introduction Exclusive real Compton scattering on the proton, $`\gamma p\gamma p`$, is a promising arena for studying the short-distance structure of the proton. In the limit of large energy $`\sqrt{s}`$ and fixed scattering angle $`\theta `$ in the center-of-mass frame, the real Compton amplitude should factorize as the convolution of a perturbative hard scattering matrix element with a nonperturbative light-cone distribution amplitude . The distribution amplitude is for the three valence quarks in the proton; it describes how their longitudinal momentum is partitioned when their transverse separation is very small. Contributions of Fock space states with more partons in the proton’s light-cone wave function should be suppressed by additional powers of $`s`$. However, the energy at which this asymptotic prediction of perturbative QCD (PQCD) becomes valid is not known a priori. Soft mechanisms such as the soft overlap (or handbag) model and the diquark model could be comparable to, or even dominant over, the PQCD mechanism at the presently accessible center-of-mass energies of 2–4 GeV. The PQCD prediction for $`\gamma p\gamma p`$ contains a number of uncertainties. First, only the Born level has been computed; next-to-leading-order corrections are likely to be large. Related to this, the Born level prediction is proportional to a high power of the running strong coupling constant, $`[\alpha _s(\mu )]^4`$, and its renormalization-scale ($`\mu `$) dependence leads to a large normalization uncertainty on the cross section. Second, the form of the proton distribution amplitude is not well understood. Several groups have produced model distribution amplitudes based primarily on QCD sum rule analyses . These distribution amplitudes can lead to quite different predictions for the Compton helicity amplitudes. Most of the proposed distribution amplitudes tend to peak in a region where two of the three quarks carry relatively small fractions $`x`$ of the proton longitudinal momentum. This has led to skepticism about the applicability of PQCD at accessible energies , because relatively soft sub-processes (relative to $`\sqrt{s}`$) can reorient quarks with small $`x`$ from the initial proton direction to the final proton direction. Despite all these caveats, it is still useful to know the PQCD predictions for $`\gamma p\gamma p`$, if nothing else as an asymptotic limit. There have already been four separate calculations at Born level . However, no two results agree with each other, even when the same proton distribution amplitudes are assumed. Given this discrepancy in the literature, and the need for consistent predictions from the PQCD mechanism, we undertook an independent recalculation of this process. Our results in fact differ from all previous work, although we find reasonable agreement with ref. for a subset of the helicity amplitudes, and excellent agreement with ref. for forward scattering angles. Our results are timely in view of the experimental situation. For over twenty years, the highest energy wide-angle Compton data available have been from an experiment at Cornell which investigated the energy range $`4.6`$ GeV$`{}_{}{}^{2}<s<12.1`$ GeV<sup>2</sup>. These data appear to obey an approximate $`d\sigma /dts^6`$ scaling law, as predicted by PQCD, although more precise data would be useful to confirm or refute this behavior. An experiment now underway at Jefferson Lab should soon improve the errors on the unpolarized cross section and its $`\theta `$\- and $`s`$-dependence, in the same kinematic range as the Cornell experiment. This experiment also plans to measure a polarization asymmetry, the transfer of longitudinal polarization from the incoming photon to the outgoing proton, for at least one angle. We shall discuss this asymmetry further in section 3. An upgrade of the Jefferson Lab electron beam to 12 GeV would allow for the very important extension of this experiment to higher energies. The proposed ELFE facility with a 25 GeV electron beam would also be a natural place to perform higher energy Compton measurements. The remainder of this paper is organized as follows. In section 2 we outline the calculation. In section 3 we present results for the unpolarized cross section and for some different polarization asymmetries. Section 4 contains our conclusions. ## 2 Calculation Since the general PQCD calculational framework for the Compton process has been described previously, e.g. in ref. , we will be brief here. The leading-twist PQCD factorization of the helicity amplitude $`_{hh^{}}^{\lambda \lambda ^{}}`$ for incoming (outgoing) photon helicity $`\lambda `$ ($`\lambda ^{}`$) and proton helicity $`h`$ ($`h^{}`$) is given by $`_{hh^{}}^{\lambda \lambda ^{}}`$ $`=`$ $`{\displaystyle \underset{d,i}{}}{\displaystyle _0^1}𝑑x_1𝑑x_2𝑑x_3𝑑y_1𝑑y_2𝑑y_3\delta \left(1{\displaystyle \underset{j=1}{\overset{3}{}}}x_j\right)\delta \left(1{\displaystyle \underset{k=1}{\overset{3}{}}}y_k\right)`$ $`\times \varphi _i(\stackrel{}{x})T_i^{(d)}(\stackrel{}{x},h,\lambda ;\stackrel{}{y},h^{},\lambda ^{})\varphi _i^{}(\stackrel{}{y}),`$ where the vectors $`\stackrel{}{x}(x_1,x_2,x_3)`$ and $`\stackrel{}{y}(y_1,y_2,y_3)`$ represent the quark longitudinal momentum fractions; $`i`$ labels the independent three-valence-quark Fock states of the proton, with distribution amplitudes $`\varphi _i(\stackrel{}{x})`$; and $`d`$ represents the sum over the diagrams that contribute to the hard-scattering amplitude $`T_i`$. The distribution amplitude represents the three-valence-quark component of the proton’s light-cone wave function, after the latter is integrated over transverse momenta up to a factorization scale $`\mu `$. (Moments of the distribution amplitude can also be defined via the matrix elements of appropriate local three-quark operators.) The distribution amplitude evolves logarithmically with $`\mu `$, but (as was also done in refs. ) we shall neglect this evolution here. The full distribution amplitude for a positive-helicity proton is, in the notation of ref. , $$|p_{}=\frac{f_N}{8\sqrt{6}}_0^1𝑑x_1𝑑x_2𝑑x_3\delta \left(1\underset{j=1}{\overset{3}{}}x_j\right)\underset{i=1}{\overset{3}{}}\varphi _i(\stackrel{}{x})|i;\stackrel{}{x},$$ (2) where $`|1;\stackrel{}{x}`$ $`=`$ $`|u_{}(x_1)u_{}(x_2)d_{}(x_3),`$ $`|2;\stackrel{}{x}`$ $`=`$ $`|u_{}(x_1)d_{}(x_2)u_{}(x_3),`$ (3) $`|3;\stackrel{}{x}`$ $`=`$ $`|d_{}(x_1)u_{}(x_2)u_{}(x_3).`$ The normalization constant $`f_N`$ can be determined from QCD sum rules or lattice QCD. We choose $`f_N=5.2\times 10^3`$ GeV<sup>2</sup> (as in refs. ). Fermi-Dirac statistics, isospin and spin symmetry result in only one independent distribution amplitude, $`\varphi _1`$; the other two are given by $`\varphi _2(x_1,x_2,x_3)`$ $`=`$ $`\varphi _1(x_1,x_2,x_3)\varphi _1(x_3,x_2,x_1),`$ (4) $`\varphi _3(x_1,x_2,x_3)`$ $`=`$ $`\varphi _1(x_3,x_2,x_1).`$ In addition to neglecting evolution of the distribution amplitude, we shall also take $`\alpha _s`$ to be fixed. The Born-level cross section then scales as $`\alpha _s^4\times f_N^4`$. The hard scattering amplitude is computed for three collinear incoming and outgoing quarks. The color and electric charge dependence can be factored off of each diagram as $$T_i^{(d)}(\stackrel{}{x},h,\lambda ;\stackrel{}{y},h^{},\lambda ^{})=C^{(d)}g^4Z_i^{(d)}\stackrel{~}{T}^{(d)}(\stackrel{}{x},\stackrel{}{y};h,\lambda ,h^{},\lambda ^{}),$$ (5) where $`C^{(d)}`$ is the color factor, $`g`$ is the strong coupling constant, and $`Z_i^{(d)}`$ is the appropriate product of quark electric charges, while $`\stackrel{~}{T}^{(d)}`$ is color and flavor independent. The helicities of the quarks in the hard scattering amplitude are conserved by the gauge interactions; therefore the proton helicity is conserved, and $`_{hh^{}}^{\lambda \lambda ^{}}=0`$ for $`hh^{}`$. Parity and time-reversal invariance further reduce the number of independent helicity amplitudes to three, which we take to be $$_{}^{},_{}^{},\mathrm{and}_{}^{}.$$ (6) In principle, 378 diagrams contribute to the hard scattering amplitude. However, 42 of them contain three-gluon vertices and have a vanishing color factor. Many others vanish for individual helicity configurations. We adopted the technique in ref. of using the parity symmetry (denoted $``$ therein) between certain classes of diagrams to reduce the number that had to be computed, while reserving the time-reversal symmetry as a check. All diagrams were computed by two independent computer programs, both based on the formalism outlined in ref. . These expressions were found to be identical to those used in the two most recent computations .We compared our results for each diagram to the formulae given in Tables III and IV of ref. . These tables contain three errors (found by M. Vanderhaeghen as well as us) in addition to one in diagram $`A71`$ that was published in an erratum. However, all these errors are typographical and do not affect the numerical results in that paper . The errors are: In the denominator of $`\stackrel{~}{T}^{(A44)}(x,,;y,,)`$, $`(\overline{x}_3,x_1)`$ should be $`(\overline{x}_3,y_1)`$; $`\stackrel{~}{T}^{(C75)}(x,,;y,,)`$ should be multiplied by $`1/c`$; and the diagram related to $`\overline{C77}`$ by $`𝒯`$ should be $`\overline{F11}`$, not $`\overline{F33}`$.Thus we agree completely with refs. on the hard scattering amplitude $`T_i`$. The next step is to perform the four-dimensional integration in eq. (2) over the independent quark momentum fractions. For the various model distribution amplitudes we used the coefficients of $`\varphi _1`$ listed in Table I of ref. (and eq. (6) of ref. ). Many diagrams include denominators that vanish inside the $`(\stackrel{}{x},\stackrel{}{y})`$ integration region, due to the presence of an internal quark and/or gluon that can go on shell. This is not a true long-distance singularity, and all the integrals are finite, diagram by diagram, but it is a technical obstacle to obtaining a reliable value for the integral. In the notation of ref. , the Feynman $`i\epsilon `$ prescription leads to singular denominators of the form $$\frac{1}{(x,y)+i\epsilon }=\mathrm{P}\frac{1}{(x,y)}i\pi \delta \left((x,y)\right),$$ (7) where P stands for principal part and $`(x,y)x(1ys^2)yc^2`$, with $`s=\mathrm{sin}(\theta /2)`$ and $`c=\mathrm{cos}(\theta /2)`$. Diagrams can be classified by the number of singular factors found in the denominator; for the Compton process this number can be 0, 1, 2 or 3. The presence of on-shell partons in the Born-level hard scattering amplitude (for particular values of $`(\stackrel{}{x},\stackrel{}{y})`$) leads to large phases in the PQCD amplitude . This is in contrast to the handbag model, which predicts an imaginary part that is small and beyond the accuracy of the model. At least four different numerical methods have previously been applied to handle the singular integrations. Ref. performed a Taylor expansion of the numerators of the integrand symmetrically about each singularity. Ref. kept the $`\epsilon `$ in eq. (7) explicit, and evaluated the integrals for a sequence of $`\epsilon `$ values tending to zero, looking for stable results. Ref. handled the imaginary parts of the singular integrals by solving the $`\delta `$-function constraint explicitly, and carried out the real, principal-part integrals by folding the region of integration over at the singularity, so that the integrand is manifestly finite. Finally, ref. deformed the $`(\stackrel{}{x},\stackrel{}{y})`$ integration contour into the complex plane, an elegant technique that requires relatively little bookkeeping for its implementation. We adopted a variation of the contour deformation technique . We first let $`x_1=\xi _1,x_2=(1\xi _1)(1\xi _2),x_3=(1\xi _1)\xi _2,`$ (8) $`y_1=\eta _1,y_2=(1\eta _1)(1\eta _2),y_3=(1\eta _1)\eta _2,`$ so that the four independent variables $`(\xi _1,\xi _2,\eta _1,\eta _2)`$ were integrated on the interval . We then deformed the single variable $`\eta _1`$ into the complex plane, so that it ran either over the piecewise linear contour $`0iϵ1+iϵ1`$, or over a semi-circular contour extending from 0 to 1. Note that this simultaneously deforms both $`y_1`$ and $`y_3`$, towards opposite sides of the real axis, while $`x_1`$ and $`x_3`$ remain real. Inspection of the denominator factors in Tables III and IV of ref. shows that this deformation is sufficient to correctly bypass the singularities in every Compton diagram. For example, the denominator of diagram A16 includes the factors $`[(x_1,y_1)+i\epsilon ][(\overline{x}_3,y_1)+i\epsilon ][(y_3,x_3)+i\epsilon ]`$, where $`\overline{x}_i1x_i`$, $`\overline{y}_i1y_i`$. Using the identity $`(x,y)=(\overline{y},\overline{x})`$, the singular factors can be rewritten as $`[(1y_1,\overline{x}_1)+i\epsilon ][(1y_1,x_3)+i\epsilon ][(y_3,x_3)+i\epsilon ]`$, which shows that $`y_1`$ and $`y_3`$ should indeed be deformed in opposite directions. If the diagram happens to contain denominators of the form $`(x_i,y_1)`$ or $`(y_3,x_i)`$, instead of $`(y_1,x_i)`$ or $`(x_i,y_3)`$, as does diagram A16, then the imaginary part should be multiplied by an overall minus sign (or equivalently, the contour should be deformed in the opposite direction with respect to the real axis). After making these contour deformations, the real and imaginary parts of the complex integrals were performed separately using the Monte Carlo integration routine VEGAS . Two independent versions of the contour integration were implemented numerically, with two different choices of contour (piecewise linear vs. semi-circular), and we also varied the deformation parameter $`ϵ`$, obtaining stable results. A third version of the integration program was constructed, which employed the Gauss-Legendre formalism with ten points per integration variable, instead of VEGAS. Although the Gauss-Legendre errors were larger than the VEGAS errors, the two sets of results were completely consistent with each other (and were both inconsistent with results from previous work; see section 3.1). We carried out further checks on our integration routines. For diagrams with only one singular factor in the denominator, one can integrate the imaginary part analytically. Using this procedure we checked the imaginary part of all diagrams with one singularity. One can also check the diagrams with no singularities in the same manner. A second check employed the identity $$\frac{1}{(x,y)(x,z)}=\frac{1ys^2}{c^2(yz)(x,y)}\frac{1zs^2}{c^2(yz)(x,z)}.$$ (9) Using eq. (9) one can reduce all three-singularity diagrams to two singularities. (These expressions can be reduced no further, though, because the four remaining singular variables are all different.) One can also reduce all diagrams that initially had two singularities to one-singularity diagrams, allowing their imaginary parts to be computed analytically. Our integration techniques were robust against all of these tests. Finally, Table V of ref. gives detailed results for diagram $`A51`$, which has two denominator singularities. We agree completely with these results, for both the real and imaginary parts. We note that ref. also attempted to evaluate this diagram by implementing the explicit $`\epsilon 0`$ method of ref. , but they obtained very different results for the imaginary part, compared with the results of their folding method. Ref. claims that the explicit $`\epsilon 0`$ method is not numerically stable. Since we agree with their results for diagram A51, we do not have cause to disagree with this claim. ## 3 Results ### 3.1 Comparison with previous work We computed the Compton helicity amplitudes for a variety of distribution amplitudes, which we refer to as CZ , GS , KS , COZ , HET , and ASY (the distribution amplitude for asymptotically large energy scales, $`\varphi _1(x_1,x_2,x_3)=120x_1x_2x_3`$). The CZ, KS and COZ distribution amplitudes, which satisfy the constraints imposed by QCD sum rules , are qualitatively similar. They feature a peak in $`\varphi _1`$ for $`x_11`$, $`x_{2,3}0`$; that is, the $`u`$ quark with the same helicity as the proton carries most of the momentum. The GS distribution amplitude has a peak in $`\varphi _1`$ for $`x_{1,3}1/2`$, $`x_20`$; thus it splits the momentum more equitably between the two quarks carrying the proton’s helicity. The HET distribution amplitude is intermediate in shape between GS and the $`\{`$CZ,KS,COZ$`\}`$ class. Before discussing our full results, we present a comparison of results in the literature. Here we choose the COZ distribution amplitude, since it was employed in four of the five existing calculations. (Only the earliest calculation , which was later superseded , did not use the COZ distribution amplitude.) The overall cross-section normalizations in the literature are sometimes difficult to determine, due for example to unspecified choices for $`\alpha _s`$. Therefore we choose to compare results for the following (normalization-independent) initial-state helicity correlation , $$A_{LL}\frac{\frac{d\sigma _+^+}{dt}\frac{d\sigma _+^{}}{dt}}{\frac{d\sigma _+^+}{dt}+\frac{d\sigma _+^{}}{dt}},$$ (10) where $`d\sigma _h^\lambda /dt`$ is the differential cross section for a helicity $`h`$ proton scattering off a helicity $`\lambda `$ photon. Figure 1 shows that none of the four calculations of $`A_{LL}`$ agrees completely with any other. The only two results that are very close are ours and that of ref. . These two curves are in excellent agreement for $`\theta <110^{}`$; however, we do not reproduce the prominent dip of ref. in the backward region. This statement is true for the four distribution amplitudes we have compared: KS, COZ, CZ and ASY . Figure 14 of the second reference in shows that the dip in $`A_{LL}`$ derives from $`\frac{d\sigma }{dt}(\gamma _{}p_{}\gamma p)|_{}^{}|^2+|_{}^{}|^2`$, and not from $`\frac{d\sigma }{dt}(\gamma _{}p_{}\gamma p)|_{}^{}|^2+|_{}^{}|^2`$. Indeed, we agree with their $`\gamma _{}p_{}\gamma p`$ cross section for all angles to better than 10%, up to an overall normalization factor which can be accounted for by different choices for $`\alpha _s`$. We agree with the $`\gamma _{}p_{}\gamma p`$ cross section only for $`\theta <110^{}`$, however. This suggests that the discrepancy with ref. is predominantly from the single helicity amplitude $`_{}^{}`$. The curve from ref. has the same general shape as ours, but is offset from it. The phases of the dominant helicity amplitudes given in ref. actually agree quite well with our results in figs. 57 below; the magnitudes are offset by relatively angle-independent factors. Ref. finds a very large asymmetry. We have made a detailed comparison of our COZ results with those of ref. , for the real and imaginary parts of the three independent helicity amplitudes. Each amplitude has been further split into four pieces , according to the number of singular propagators in the diagram (as determined from Tables III and IV of ref. ). The zero propagator terms (which were integrated analytically by both groups) agree to high precision (6 digits). The one propagator terms agree to within VEGAS errors, except for the imaginary part of one helicity amplitude ($`_{}^{}`$) which is within 10%. For the two propagator terms, we are in agreement on the real part of $`_{}^{}`$ and $`_{}^{}`$, but have a large discrepancy in the imaginary part. Strangely enough, for $`_{}^{}`$ we agree on the imaginary part but disagree on the real part! For the three propagator terms, both the real and imaginary parts disagree for all three helicity amplitudes. The bulk of our overall numerical disagreement comes from the two propagator terms contributing to the imaginary part of $`_{}^{}`$. The two propagator terms are often 100 times larger than ours, and they drive ref. ’s values for $`\mathrm{Im}_{}^{}`$ to be roughly a factor of 10 larger than ours. We also calculated $`A_{LL}`$ for the COZ distribution amplitude using Gauss-Legendre integration instead of VEGAS. The result agrees with our VEGAS result shown in fig. 1 (albeit with larger errors), and it disagrees with the other results, in particular that of ref. for $`\theta >110^{}`$. ### 3.2 Helicity amplitudes and unpolarized cross section In figs. 24 we display our results for the polarized differential cross sections, $$s^6\frac{d\sigma _{hh^{}}^{\lambda \lambda ^{}}}{dt}=\frac{s^4}{16\pi }|_{hh^{}}^{\lambda \lambda ^{}}|^2,$$ (11) for the three independent helicity configurations. Each figure plots the results for five different distribution amplitudes. (For HET we shall only plot the unpolarized cross section.) These plots were made for $`\alpha _{\mathrm{em}}^1=137.036`$, $`\alpha _s=0.3`$ and $`f_N=5.2\times 10^3`$ GeV<sup>2</sup>, so they can be compared directly with ref. . The phases of the helicity amplitudes are plotted in figs. 57; the GS distribution amplitude has a much different behavior and is therefore plotted separately, in fig. 8. The phases are generally large; indeed $`_{}^{}`$ is almost pure imaginary (except for the GS distribution amplitude). For reference, we also provide in Table 1 our numerical results for the real and imaginary part of $`_{}^{}`$, for the COZ distribution amplitude, including errors from the VEGAS integration. Figure 9 shows our predictions for the unpolarized differential Compton cross section, given by $$s^6\frac{d\sigma }{dt}=\frac{1}{4}\underset{\lambda ,\lambda ^{},h,h^{}}{}s^6\frac{d\sigma _{hh^{}}^{\lambda \lambda ^{}}}{dt},$$ (12) along with the experimental data from ref. . For the values used $`\alpha _s=0.3`$, $`f_N=5.2\times 10^3`$ GeV<sup>2</sup>, the predictions lie at least an order of magnitude below the data. Since the PQCD cross section scales like $`\alpha _s^4`$, accommodating a factor of 10 by changing $`\alpha _s`$ would require $`\alpha _s0.5`$. While this is not out of the question, and while some variation in $`f_N`$ could be considered as well, this may be pushing the validity of perturbation theory. On the other hand, the shape of the curves (i.e., ignoring the overall normalization) matches the data quite well for the KS, COZ, CZ, and HET distribution amplitudes. ### 3.3 Normalization by $`F_1^p(Q^2)`$ As mentioned in the introduction, the $`\alpha _s^4(\mu )`$ scaling of the proton Compton cross section at Born level introduces a large normalization uncertainty into the PQCD prediction. Uncertainty in $`f_N`$ also contributes. Both of these uncertainties can be removed at Born level by considering the dimensionless ratio $$\frac{s^6\frac{d\sigma _{\gamma p}}{dt}}{[Q^4F_1^p(Q^2)]^2},$$ (13) where $`F_1^p(Q^2)`$ is the elastic Dirac form factor for the proton at space-like momentum transfer $`Q`$. One might also imagine normalizing the Compton cross section by the time-like proton form factor. At leading order in $`\alpha _s`$, the PQCD predictions in the space-like and time-like regions are identical ; however, experimentally the time-like form factor is larger by a factor of about two . Higher order PQCD corrections can in principle account for this factor, as Sudakov effects are different in the two regions . The Compton scattering kinematics are much closer to those of the space-like proton form factor than the time-like one, at least as far as the proton is concerned. Therefore Sudakov and related higher-order effects are best cancelled by normalizing with the space-like form factor. At leading twist, $`F_1^p(Q^2)`$ is predicted to be the same as the magnetic form factor $`G_M^p(Q^2)`$. Experimentally, these are close but not identical . To normalize the experimental Compton points, we use the experimental form factor values, $$Q^4F_1^p(Q^2)Q^4G_M^p(Q^2)1.0\mathrm{GeV}^4,Q^27\text{}\mathrm{\hspace{0.17em}15}\mathrm{GeV}^2,$$ (14) which are representative of the region where both scaled form factors flatten out, and are also similar to the highest experimental values of $`s`$ available in Compton scattering.<sup>§</sup><sup>§</sup>§If one equates the four-momentum transfer to the proton in the two processes — $`Q^2`$ in the form factor and $`t`$ in Compton scattering — then the corresponding Compton $`s=2Q^2/(1\mathrm{cos}\theta )`$ should actually be considerably bigger than $`Q^2`$. At $`90^{}`$, for example, $`s=2Q^2`$. Unfortunately, there are no experimental Compton data with $`s`$ this large (all have $`t<5.3`$ GeV<sup>2</sup>), so there is not a good overlap with the region (14) where the elastic form factor is beginning to scale properly. To normalize the theoretical Compton curves, we recalculated the proton form factor at leading order in PQCD, obtaining $$Q^4F_1^p(Q^2)=\frac{(4\pi \alpha _sf_N)^2}{216}I_F,$$ (15) where $$I_F=\{\begin{array}{cc}2.500\times 10^5\hfill & \text{(CZ),}\hfill \\ 2.505\times 10^5\hfill & \text{(GS),}\hfill \\ 3.653\times 10^5\hfill & \text{(KS),}\hfill \\ 2.897\times 10^5\hfill & \text{(COZ),}\hfill \\ 3.303\times 10^5\hfill & \text{(HET),}\hfill \\ 0\hfill & \text{(ASY).}\hfill \end{array}$$ (16) These results, using the wave function (2) which is equivalent to that in ref. , are precisely a factor of two smaller than several previous calculations using the same wave functions . We do not understand the origin of this discrepancy. We do agree with the normalization of the hard scattering amplitude and the form factor in ref. (which uses, however, a different representation of the proton wave function than eq. (2)). Figure 10 shows the Compton cross section, normalized according to eq. (13), for both PQCD and the experimental data. We omit the ASY distribution amplitude, since the leading order ASY form factor vanishes. Compared with the conventionally normalized curves in fig. 9, the spread between the predictions of the three qualitatively similar distribution amplitudes, KS, COZ and CZ, has become much smaller. The theoretical curves also lie a factor of 2 to 5 closer to the data. However, they still fall about an order of magnitude below the data at the widest scattering angles. (The HET distribution amplitude does slightly better than this.) Thus it seems unlikely that the elastic proton form factor and the Compton scattering amplitude are both described by PQCD at presently accessible energies, unless there are large higher-order and process-dependent corrections. ### 3.4 Asymmetries Various polarization asymmetries can be constructed from the helicity amplitudes. These observables may provide additional diagnostic power for uncovering the Compton scattering mechanism, beyond what the unpolarized cross section provides. Figure 11 presents the perturbative QCD results for the initial state helicity correlation $`A_{LL}`$ defined in eq. (10). Also shown is the handbag model prediction for $`E_\gamma =4`$ GeV, where the form factors $`R_{V,A}`$ were evaluated using the parton distribution functions of GRV . In leading-twist PQCD, the proton helicity is conserved. The handbag model does not inherently require proton helicity conservation, but it has been assumed in ref. . Thus the PQCD and handbag curves for $`A_{LL}`$ in fig. 11 can be equated to the longitudinal photon-to-proton polarization transfer asymmetry, which is slated to be measured for at least one scattering angle in an upcoming experiment . The diquark model analyzed in ref. has nonvanishing proton helicity-flip amplitudes at finite $`s`$, making $`A_{LL}`$ and the polarization transfer into distinct asymmetries. We plot the diquark prediction for $`A_{LL}`$ from ref. . Figure 11 shows that PQCD gives quite different qualitative behavior from both the handbag and diquark models for $`A_{LL}`$, and they should be distinguishable with the help of experimental data at just a couple of backward scattering angles. A caveat is that the GS curve is somewhat oscillatory, so one might wonder whether a distribution amplitude ‘between’ GS and the $`\{`$CZ,COZ,KS$`\}`$ class of amplitudes could produce behavior similar to the handbag model. One can also define a photon spin transfer coefficient $$D_{LL}\frac{\frac{d\sigma ^{++}}{dt}\frac{d\sigma ^+}{dt}}{\frac{d\sigma ^{++}}{dt}+\frac{d\sigma ^+}{dt}},$$ (17) where now $`d\sigma ^{\lambda \lambda ^{}}/dt`$ is the differential cross section for initial and final state photon helicities $`\lambda `$ and $`\lambda ^{}`$, and unpolarized incoming and outgoing protons. Figure 12 gives the PQCD predictions for this asymmetry, as well as that of the diquark model for $`E_\gamma =4`$ GeV and a ‘standard’ distribution amplitude . The handbag model predicts $`D_{LL}=1`$, basically because the helicity-flip quark Compton amplitude $`\gamma _{}q\gamma _{}q`$ vanishes at Born level for massless quarks. The final asymmetry we plot is the photon asymmetry $$\mathrm{\Sigma }\frac{\frac{d\sigma _{}}{dt}\frac{d\sigma _{}}{dt}}{\frac{d\sigma _{}}{dt}+\frac{d\sigma _{}}{dt}},$$ (18) where $`d\sigma _{}/dt`$ and $`d\sigma _{}/dt`$ are the differential cross sections for linearly polarized photons, with the polarization plane perpendicular or parallel (respectively) to the scattering plane. Generation of this asymmetry requires a nonzero photon helicity-flip amplitude; hence the asymmetry vanishes in the handbag model. Figure 13 plots the PQCD and diquark predictions. The diquark prediction is shown for $`E_\gamma =4`$ GeV and a ‘standard’ distribution amplitude; for another distribution amplitude $`\mathrm{\Sigma }`$ can become positive in the backward region instead of negative . This asymmetry has actually been measured , however only for $`E_\gamma =3.45`$ GeV and $`\mathrm{cos}\theta >0.8`$. A high-energy wide-angle measurement would be very useful for distinguishing between handbag and PQCD mechanisms. ## 4 Conclusions Motivated by conflicting results in the literature, we have recalculated the fixed-order, Born level predictions of perturbative QCD for proton Compton scattering, for five different distribution amplitudes. While our results do not agree with those of any previous group, they do agree very well with those of ref. for $`\theta <110^{}`$, and the differences for $`\theta >110^{}`$ seem to be dominated by a single helicity amplitude, $`_{}^{}`$. From the helicity amplitudes we computed three separate polarization asymmetries. Experimental measurements of these asymmetries could be used in conjunction with the unpolarized differential cross section in order to help shed light on the mechanism involved in the Compton scattering process. We also have attempted to reduce the uncertainty in the overall normalization of the Compton cross section by normalizing it by the square of the elastic proton form factor. This exercise reduces the spread in the theoretical predictions, but it leaves them an order of magnitude below the data. Unfortunately, this result makes it difficult to simultaneously explain the current data on the elastic proton form factor and on Compton scattering in terms of perturbative QCD, without appealing to large uncalculated higher-order and process-dependent corrections. Acknowledgments We thank Stan Brodsky for suggesting this work, and for useful discussions and comments on the manuscript. We are grateful to Marc Vanderhaeghen and particularly Andreas Kronfeld for detailed discussions of their work and for providing us with unpublished data files. We thank Markus Diehl for several helpful remarks and suggestions.
warning/0004/cond-mat0004281.html
ar5iv
text
# Damaging and Cracks in Thin Mud Layers. ## I Introduction In this paper we present a careful and detailed study of a minimal fracture model that has been introduced at the aim of describing the main features of paint dessication-like phenomena. The purpose of this work is to focus on the statistical properties of these phenomena on the basis of a recent experimental work. Following the results of this work, we assumed that the main source of stress is given by the local friction between the layer of material and the bottom surface of the container. Moreover, it has been noticed that the characteristic size of crack patterns varies linearly with the layer thickness. In the limit of zero thickness crack patterns lose their polygonal structure (the characteristic size of the polygons is zero) and become branched fractals. In order to model this behavior, a minimal automaton lattice model, inspired by Invasion Percolation and by the vectorial and scalar models described in , has been recently introduced by the authors . Here, we present an extended report of the study described in , with a detailed description of the analytical calculations a new numerical and theoretical results. All the models for quasi-static fractures, describe crack evolution through a non-local Laplacian field (electric field, electric current) acting on a solid network of bonds or sites . In others the stress field evolves by keeping minimum the energy of the system. In such a case the components of the obtained vectorial equations are similar to the equations describing the action of a Laplacian field . In this model,instead, no explicit field is present. The effect of the stress is represented by an extremal breaking rule and local random breaking thresholds: at each time step, the bond with the smallest threshold is removed from the lattice. Short ranged correlations are introduced through a damaging of the non-broken nearest neighbors bonds of the just removed bond. According to the kind of fracture we deal with, one can introduce different types of damaging. In this paper two limiting cases are studied. This model is inspired by the above cited experimental observations that, in an extremely thin layer of mud or paint, the only source of stress is the local friction with the container. Moreover, since the drying mud is a mixture of a liquid and a solid (usually amorphous) phase, no long range stress relaxation is present, although the growing crack can affect the properties of the medium in its neighborhood. Some important physical properties of the model are explained by using an approach based on the Run Time Statistics (RTS) scheme . In particular, we are able to compute some relevant quantities, such as the evolution of the breaking probability, and of the probability distribution of breaking thresholds. The paper is so organized. In Sec. II the model is described. In Sect. III the results of numerical simulations are presented. In Sect. IV the model is studied analytically and theoretical and numerical results are compared. ## II The Model A square lattice is considered and a quenched random variable $`x_i`$ is assigned to each bond $`i`$. The $`x_i`$’s are independently extracted from an uniform probability density between $`0`$ and $`1`$. At each time-step $`t`$, the unbroken bond in the lattice with the lowest value of the variable is broken (removed). Then damage (weakening) is applied, in a way explained below, to the unbroken nearest neighbors (n.n.) of the just removed bond. After having introduced the damaging, the breaking and damaging steps are repeated until a connected, percolating, subset (infinite cluster) of removed bonds appears, dividing the system into two disconnected parts. Before explaining the definition of damaging, it is necessary to introduce some notations. The set of broken bonds up to time $`t`$ is indicated with $`C_t`$, and the set of non-broken bonds with $`C_t`$. The number of bonds belonging to $`C_t`$ is $`C_t=t`$, while $`C_t=Nt`$ (where $`N`$ is the total number of bonds in the lattice), in fact $`C_t`$ is composed by the whole lattice minus the bonds in $`C_t`$. The definition of $`C_t`$ is independent of the model. That of $`C_t`$, instead, can differ a lot from a model to another; for instance in Invasion Percolation (IP) it is simply given by the set of nearest neighbors of the bonds in $`C_t`$. Two different kinds of weakening of the unbroken neighbors are studied: Either by direct weakening or by re-distribution of the “stress”. In the first case (rule 1), the unbroken n.n. are weakened, by extracting a new threshold $`x_i^{}`$ between $`0`$ and the former value $`x_i`$. In this case an average weakening of one half of the former value at time is obtained. In the second case (rule 2), instead, each neighbor has a threshold weakened by a fraction of the threshold of the bond just removed. Both cases mimic the damaging produced by the enhancement of the stress nearby crack tips: the first case refers to a situation where stochasticity (thermal fluctuations) is important in the determination of the new thresholds , the second case refers to a deterministic effect around the crack tip. From the point of view of mud cracking, the two-dimensional lattice represents a very thin layer of mud (or paint), and quenched disorder accounts for local stress induced by inhomogeneous desiccation of the sample. Since the evolution of cracks in mud desiccation is assumed to be a slow process, the dynamics is assumed to be quasi-static, i.e. one microscopical breaking with relative damaging for each time-step. Some authors correctly point out that otherwise time-dependent effects and a non-equilibrium dynamics are relevant in crack propagation . In this model the explicit presence of an external field (applied stress) and of the response of the material (strain of bonds) have been eliminated. The only quantity present is the breaking threshold, the dynamics of which is chosen to reproduce the evolution of cracks. This simulates the presence of a local stress field, acting not on the boundaries but directly on each bond. Our assumption is based on the experimental results in Ref. , where, as the mud layer becomes thinner, only the inhomogeneities drive the nucleation of cracks. Furthermore, the hypothesis of crack developing under the same state of strain not only is usually applied in the presence of thermal gradients , but is also commonly reported in experiments of loading of softened material . Hence, such a model is particularly suitable to describe, for example, paint drying, where the stress applied to the painted surface depends on the local action of external conditions (density gradient in the paint). Moreover, its simplicity allows us to study analytically its properties, which is a non common feature for fracture models. ## III Numerical Results Numerical simulations, with cylindrical symmetry (periodic boundary conditions in the horizontal direction) for various system sizes $`L`$ have been performed. The dynamics stops as soon as a crack spanning the system in the vertical direction appears. Both damaging rules are implemented, and they are discussed in parallel. Despite the simplicity of the dynamical rules, the results are rather interesting. We have computed the fractal dimension of the percolating cluster, the distribution of the size of clusters of broken bonds, the avalanche size-distribution (in order to check if long range temporal correlations are present), and the probability distribution of the breaking thresholds at the percolation time. An avalanche can be defined as an ensemble of causally and geometrically connected breakdowns (see below for a rigorous definition). Under this respect the size-distribution of such avalanches represents the probability of a large or small response of the system to an external solicitation. For example a power law distribution represents a critical state of the system where the response has not a characteristic size. The fractal dimension $`D_f`$ of the percolating cluster is computed using the box-counting method. The analysis is restricted to the spanning cluster to reduce the finite size effects present for the smaller clusters. The results of the box-counting analysis are reported in Table I for the different sizes and for the two damaging rules. The values of $`D_f`$ for the two damaging rules coincide within the error bars. The connected clusters of broken bonds are identified with a standard cluster counting procedure, based on the Hoshen-Kopelman algorithm . Also the distribution of finite clusters is nontrivial, showing a clear power law with exponent $`\tau _c=1.54(2)`$ (see Fig. 1(a)) for rule 1 and $`\tau _c=1.57(3)`$ for rule 2. The plots labelled with (b) in Fig. 1 refer to the avalanche size distribution. This quantity is interesting with respect to recent experiments and models where a power law behavior of the acoustic emission has been related to Self-Organized Criticality (SOC). The presence of a SOC-like behavior would mean that the dynamics of fractures itself leads the sample to a steady state where small variation of the external field can trigger reaction at any length-scale. In particular the external field in this case is the applied stress, and the response of the sample can be considered as the energy released (acoustic emission) by one avalanche of cracks, where avalanche means a causally and geometrically connected series of breakdowns. In this oversimplified model the external stress can be considered constant, since the only change after any single breakdown is due the damaging of the n.n.. Consequently, in this work the size of an avalanche is monitored as a measure for the acoustic emission. An avalanche can be defined as follows. Let us suppose that a bond $`i`$ grows (i.e. it is broken) at time $`t`$; this is the initiator of an avalanche, which is defined as the set of events geometrically and causally connected to the initial one (bond $`i`$). “Causal” connection refers to the weakening following any bond breaking. In particular, when bond $`i`$ grows at time $`t`$, the avalanche goes on at time $`t+1`$ if a unbroken first neighbor bond $`j`$ of $`i`$ is removed. At time $`t+2`$ the avalanche goes on if a bond $`k`$ grows where $`k`$ is a unbroken first neighbor of $`i`$ or $`j`$ and so on. A linear-log plot of the probability distribution of avalanche size, versus sample size $`L`$ is shown in Fig. 1b. After a power law transient, an exponential distribution is reached, indicating that a characteristic size exists for the avalanches. One can note that both for weakening rule 1 and weakening rule 2 simulations give qualitatively similar results, although for rule 2 the characteristic time of avalanches is smaller. This is easily explained, since the damaging rule 2 is less strong than rule 1, and, consequently, the causal connection between subsequent breaking events is weaker. This result for avalanches is similar to those obtained for a scalar model of Dielectric Breakdown, but differs from the avalanche behavior in models of fracture . The explanation of this behavior is motivated by two arguments. Firstly, in the present definition of an avalanche the threshold is changed only for the n.n.. This introduces a typical length scale, while other definitions consider as the threshold the ratio between local field and resistivity, thus giving the possibility of large scale correlations. Secondly, in this model broken bonds are removed from the system. This represents a substantial difference with many SOC models with quenched disorder presented in the literature. For example, in a simple toy model of SOC due to Bak and Sneppen (where a similar refresh of thresholds is present) the dynamics produces clear power laws in the avalanche distribution. There, each site (species) deleted is replaced by a new one and is not definitively removed. In our model, instead, the number of candidates $`C_t`$ to be broken at each time-step decreases in time. This is a crucial point, since indeed power law behavior in the presence of a scalar field seems to be related to a “reconstructing rule” that allows one to deal with a system where removed bonds are replaced by new ones. Therefore, only in the case of plastic deformation, one is in the presence of a steady state, as correctly pointed out by Ref.. Therefore, the fractal dimension, the cluster size distribution and, to some extent, the avalanche size distribution seem to be universal with respect to the two different local damaging rules. In the next section the study will focus on some quantities which, instead, are not universal and reproduce the evolution of the mechanical properties of the material during the fracturing dynamics. These quantities are the average probability density of breaking thresholds, or histogram, $`\varphi _t(x)`$, and as a by-product the mean breaking threshold $`x(t)`$, which expresses the average resistance to breaking, or rigidity, of the system at time $`t`$. These quantities will be studied both numerically, and analytically, by using a probabilistic tool called Run Time Statistics (RTS) . ## IV RTS Derivation of the average weakening of the material As seen above, the evolution of the crack is described by a quasi-static extremal dynamics in a medium with quenched disorder. The most important question for a theoretical comprehension of the model is: which is the source of the spatio-temporal correlations developed by the dynamics? As pointed in in relation to Invasion Percolation (IP), the source can be found in the memory effects developed by the evolution of the dynamics itself via an interplay between dynamical rules and quenched disorder. This can be simply realized observing that the knowledge of the growth history up to a time $`t`$, provides information about the probability distribution and the correlations of the random bond-thresholds. This information has to be added to the original information that the thresholds are independently extracted from the uniform probability density in the interval $`[0,1]`$. Moreover this information influences the probabilities of the different possible continuations of the dynamics for larger time. This memory effects can be studied using carefully the notion of conditional probability. This kind of approach to growth dynamics with quenched disorder has been developed in , with particular reference to IP. This peculiar probabilistic algorithm is called Run Time Statistics (RTS). A particular modification of this tool is presented here taking into account the damaging mechanism, which is not present in IP-like models. Finally, RTS is used in order to predict analytically some relevant quantities as the evolution of both the average probability density of breaking thresholds of unbroken bonds and of the mean resistance to breakdown $`x(t)`$ of the material. Here we provide directly the final RTS formulas together with a brief sketch about their meaning. A detailed derivation of the analytical results of this section is given in Appendix $`A`$. The RTS approach permits mainly to answer the following two questions, once given a certain time-ordered geometrical path followed by the dynamics up to time $`t`$: 1. which is the effective probability density function of the variables $`x_i`$ of the lattice conditioned to the knowledge of this fixed past dynamical history; 2. which is the conditional probability of any further growth event at the next time-step. In order to introduce operative formulas, let us think to know the “one-bond” effective probability density functions $`p_{i,t}(x)`$ (conditioned to the past dynamical history) for each non-broken bond $`i`$. As it is clarified in Appendix A, this “one-bond” formulation of RTS is an approximate of the rigorous one. However, as shown in , it is a good approximation when the number of random numbers is large (as in this case). First of all one can write the breaking probability $`\mu _{i,t}`$ for each bond $`i`$ at that time-step: $$\mu _{i,t}=_0^1𝑑xp_{i,t}(x)\left[\underset{k(i)}{\overset{C_t}{}}_x^1𝑑yp_{k,t}(y)\right],$$ (1) where $`C_t`$ is the whole set of unbroken bonds. Note that at time $`t`$ the number of bonds in $`C_t`$ is $`(2L^2t)`$, i.e., the total number $`2L^2`$ of bonds in a square lattice of side $`L`$ minus the number of broken bonds before time $`t`$. Eq. (1) expresses nothing else than the effective probability that $`x_i`$ is the minimum in the set $`C_t`$ conditioned to the past history. The next important step is to update each $`p_{j,t}(x)`$ by conditioning them to this latest growth event. In this way one obtains the $`p_{j,t+1}(x)`$’s conditioned to the history up to the time-step $`t+1`$. The effective probability density at time $`t+1`$ of the latest grown bond $`i`$ is usually called $`m_{i,t+1}(x)`$, in order to distinguish it from the densities of still unbroken bonds. It is given by $$m_{i,t+1}(x)=\frac{1}{\mu _{i,t}}p_{i,t}(x)\left[\underset{k(i)}{\overset{C_t}{}}_x^1𝑑yp_{k,t}(y)\right].$$ (2) Eq. (2) (multiplied by $`dx`$) gives the “effective” probability that $`xx_ix+dx`$, conditioned to the past fixed dynamical history (time-ordered path) up to time $`t+1`$: the “memory” of the history up to time $`t`$ is “recorded” in the set of functions $`p_{k,t}(x)`$, where $`k`$ runs over all the bond belonging to $`C_t`$, while the last step is recorded in the particular functional relationship between $`m_{i,t+1}(x)`$ and the set $`\left\{p_{k,t}(x)\right\}`$ itself. This relationship is imposed by the order relation among the interface variable $`x_k`$, i.e. by the fact that $`x_i`$ is the minimum in $`C_t`$. Note that, once a bond is broken, it does not participate anymore to the dynamics. For this reason, the “effective” probability density function of its threshold does not change anymore in time and is given definitely by Eq. (2). For the remaining bonds one has to distinguish among the unbroken bonds far away from the bond $`i`$ and the unbroken nearest neighbors bonds, which will be weakened by the growth of bond $`i`$. The updating rules, for the two different mechanisms of damaging, differ only for this last set of bonds. For the non-weakened bonds, one has in both cases the following updating equation: $$p_{j,t+1}(x)=\frac{1}{\mu _{i,t}}p_{j,t}(x)_0^x𝑑yp_{i,t}(y)\left[\underset{k(i,j)}{\overset{C_t}{}}_y^1𝑑zp_{k,t}(z)\right]$$ (3) The updating equations for the weakened bonds are instead the following: (1) For the damaging mechanism 1: $`p_{j,t+1}(x)={\displaystyle \frac{1}{\mu _{i,t}}}{\displaystyle _0^1}dy{\displaystyle \frac{1}{y}}\theta (yx)p_{j,t}(y)\times `$ (4) $`\times {\displaystyle _0^y}dzp_{i,t}(z)\left[{\displaystyle \underset{k(i,j)}{\overset{C_t}{}}}{\displaystyle _z^1}dup_{k,t}(u)\right].`$ (5) (2) For the damaging mechanism 2 (see Appendix A): $`p_{j,t+1}(x)=`$ (6) $`={\displaystyle \frac{1}{\mu _{i,t}}}{\displaystyle _0^1}dy\left[{\displaystyle \underset{k(i,j)}{\overset{C_t}{}}}{\displaystyle _y^1}dzp_{k,t}(z)\right]p_{i,t}(y)p_{j,t}(x+{\displaystyle \frac{y}{n_{i,t}}})\times `$ (7) $`\times \theta \left({\displaystyle \frac{n_{i,t}}{n_{i,t}1}}xy\right)\theta \left(n_{i,t}(1x)y\right)`$ (8) Note that the main difference between Eq. (5) and Eq. (8) is due to the fact that the number of n.n. $`n_{i,t}`$ of the bond $`i`$ at time $`t`$ appears explicitly only in the latter, i.e. only in the second model (rule 2) the damaging is an explicit function of the geometry, while in the former (rule 1) the damaging is a “one-bond” process. Eqs. (1-3) coincide with the ones introduced for the RTS approach to IP (apart from the different definition of the growth interface $`C_t`$). Eqs. (5), (8), instead, are new and account for the n.n. weakening. Eqs. (1-8) allow one to study the extremal deterministic dynamics as a kind of stochastic process with memory. In particular, $`\mu _{i,t}`$ can be used to evaluate systematically the statistical weight of a fixed time-ordered growth path, while the $`p_{j,t}(x)`$’s store information about the growth history. A very important quantity to characterize the properties of the dynamics is the empirical distribution (or histogram) of unbroken thresholds. This quantity is defined as: $$h_t(x)=\underset{jC_t}{}p_{j,t}(x)$$ (9) where, $`h_t(x)dx`$ is the number of non-broken bonds between $`x`$ and $`x+dx`$ at time $`t`$, conditioned to the past dynamical history. Considering the effect of the growth of bond $`i`$ at time $`t`$ on this quantity, one gets $$h_{t+1}(x)=h_t(x)m_{i,t+1}(x)\underset{j(i)}{}p_{j,t}(x)+\underset{j(i)}{}p_{j,t+1}(x)$$ (10) where $`j(i)`$ indicates the sum over the $`n_{i,t}`$ unbroken n.n. of $`i`$. Moreover, $`m_{i,t+1}(x)`$ and $`p_{j,t+1}(x)`$ are given respectively by Eq. (2) and Eqs. (3), (5) ( (8) for rule 2) . Being the histogram an almost self-averaging quantity of the model in the large time limit, one can evaluate its shape in the “typical” realization of the dynamics taking the average over all the possible histories up to time $`t+1`$. The notation $`\mathrm{}`$ is introduced to indicate for this average. The l.h.s. of Eq. (10) can be computed as $$h_{t+1}(x)=C_{t+1}\varphi _{t+1}(x)=[N(t+1)]\varphi _{t+1}(x),$$ (11) where $`N=2L^2`$ is the total number of bonds in the lattice and $`\varphi _t(x)`$ represents the average thresholds density function over the unbroken bonds at time $`t`$ (normalized to $`1`$), i.e. $`\varphi _t(x)=p_{k,t}(x)`$ where $`k`$ is a generic interface bond. For the r.h.s. of Eq. (10) the main difficulty arises in the evaluation of $`m_{i,t+1}`$ and $`_{j(i)}p_{j,t+1}(x)`$. Following , one can write $$m_{i,t+1}(Nt)\varphi _t(x)\left[1_0^x𝑑y\varphi _t(y)\right]^{Nt1}$$ (12) In obtaining Eq. (12), we used the definition of $`\varphi _t(x)`$ and the following approximation: $$\underset{kC_t}{}p_{k,t}(x_k)=\underset{kC_t}{}p_{k,t}(x_k)=\underset{kC_t}{}\varphi _t(x_k).$$ (13) Using again the definition of $`\varphi _t(x)`$, one gets $$\underset{j(i)}{}p_{j,t}(x)=n_t\varphi _t(x),$$ (14) where $`n_t=n_{i,t}`$. Using Eq. (5), corresponding to the weakening rule 1, and the approximations given by Eq. (13), one has $`{\displaystyle \underset{j(i)}{}}p_{j,t+1}(x)={\displaystyle \frac{n_t(Nt)}{Nt1}}{\displaystyle _x^1}dy{\displaystyle \frac{\varphi _t(y)}{y}}\times `$ (15) $`\times \left\{1\left[1{\displaystyle _0^y}𝑑z\varphi _t(z)\right]^{Nt1}\right\}.`$ (16) The equation for the $`\varphi _{t+1}(x)`$ for rule 1 will finally read: $`\varphi _{t+1}(x)={\displaystyle \frac{Ntn_t}{Nt1}}\varphi _t(x)+`$ (17) $`{\displaystyle \frac{Nt}{Nt1}}\varphi _t(x)\left[1{\displaystyle _0^x}𝑑y\varphi _t(y)\right]^{Nt1}+`$ (18) $`+n_t{\displaystyle \frac{Nt}{(Nt1)^2}}{\displaystyle _x^1}dy{\displaystyle \frac{\varphi _t(y)}{y}}\times `$ (19) $`\times \left\{1\left[1{\displaystyle _0^y}𝑑z\varphi _t(z)\right]^{Nt1}\right\}`$ (20) Note that even at percolation time $`Nt`$ is a large number. For this reason terms in Eq. (20) containing the term $`\left[1_0^x𝑑y\varphi _t(y)\right]^{Nt1}`$ are negligible for $`x`$ such that $`_0^x𝑑y\varphi _t(y)`$ is finite (i.e. larger than $`1/(Nt)`$). It is easy to show that the continuum limit of Eq. (20), for such values of $`x`$, is invariant under the rescaling $`LaL`$ (i.e. $`Na^2N)`$ and $`ta^2t`$. This result is based on the assumption that: $$n_t(L)=n_{a^2t}(aL)$$ (21) The numerical simulations suggest the following scaling form for $`n_t(L)`$ (see Fig. 2): $$n_t(L)=n_{max}\left[\frac{1}{1+t/AL^2}\right]^\beta ,$$ (22) where $`\beta =0.23(2)`$, $`A=0.030(2)`$ and $`n_{max}=6`$ is the lattice coordination number. This form for $`n_t`$ satisfies Eq. (21). The study of the weakening rule 2 is quite similar. Eqs. (1-3) keep the same, while the conditioned probability density for the weakened bonds is given by Eq. (8). By following the same steps as above, the following equation for the $`\varphi _{t+1}(x)`$ is obtained: $`\varphi _{t+1}(x)={\displaystyle \frac{Ntn_t}{Nt1}}\varphi _t(x)+`$ (23) $`{\displaystyle \frac{Nt}{Nt1}}\varphi _t(x)\left[1{\displaystyle _0^x}𝑑y\varphi _t(y)\right]^{Nt1}+`$ (24) $`+n_t{\displaystyle \frac{Nt}{Nt1}}{\displaystyle _0^1}dy[1{\displaystyle _0^y}dz\varphi _t(z)]^{Nt2}\varphi _t(y)\times `$ (25) $`\times \varphi _t\left(x+{\displaystyle \frac{y}{n_t}}\right)\theta ({\displaystyle \frac{n_t}{n_t1}}xy)\theta (n_t(1x)y)`$ (26) All the assumptions we made for the case 1, including the scaling ansatz given in Eq. (21), are valid for case 2. In particular, from numerical simulations, one can find the following behavior for $`n_t(L)`$ (see Fig. 3) : $$n_t(L)n_{max}\mathrm{exp}\left(\frac{t}{AL^2}\right).$$ (27) The analytical study of both kinds of weakening allows to make three important predictions: (1) Firstly, we find both theoretically, from the numerical solution of Eqs.(20, 26), and from the numerical simulations of the model, a discontinuity in the histogram (see Fig. 4 and 5), indicating that the system evolves in such a way as to remove all bonds with threshold smaller than some critical value $`x_c`$. (2) Second, from the symmetry properties of Eqs. (2026), we deduce that the number $`t_{sp}(L)`$ of broken bonds at the percolation time is proportional to $`L^2`$, even though the percolating cluster is fractal. This result, confirmed by numerical simulations (see Fig. 6a, Fig. 7a), and compatible with the scaling function (22) for $`n_t(L)`$, is deduced supposing that at the percolation time the shape of the histogram is independent of $`L`$, an assumption which fits well with the numerical histogram (see Fig. 4a and Fig. 5a). (3) Finally, we present an approximated result for the dynamical behavior of the average value (over the unbroken bonds) of the thresholds $`x(t)`$. This quantity can be seen as a characterization of the average resistance of the material in time. In order to find the evolution equation of $`x(t)`$ for the damaging rule 1, it is enough to multiply both sides of, respectively, Eq. (20) and Eq. (26) for $`x`$ and integrate them in the whole interval $`[0,1]`$. Then one finds: $`x(t+1)=\left(1{\displaystyle \frac{n_t2}{2(Nt1)}}\right)x(t)+`$ (28) $`{\displaystyle \frac{1+n_t/[2(Nt1)]}{Nt1}}{\displaystyle _0^1}𝑑x\left[1{\displaystyle _0^x}𝑑y\varphi _t(y)\right]^{Nt}.`$ (29) For the damaging rule 2, the way to find the equation for $`x`$ is even simpler. In fact, it is enough to consider that at each time step, the global effect on $`x`$ is equivalent to remove two bonds with the resistance equal to the minimal one at that time. Therefore, one can write: $`x(t+1)=\left(1+{\displaystyle \frac{1}{Nt1}}\right)x(t)+`$ (30) $`{\displaystyle \frac{2}{Nt1}}{\displaystyle _0^1}𝑑x\left[1{\displaystyle _0^x}𝑑y\varphi _t(y)\right]^{Nt}.`$ (31) For rule 1, it is simple to see, from Eq. (29), that $`x(t+1)<x(t)`$ until $`n_t>2`$ (which is verified for all the times). This means that on average the medium weakens during the evolution even if the weakest bond is removed at any time step. This is due to the fact that, in this case the weakening of the neighbors of the weakest interface bond has a stronger effect on the material than the removal of the weakest bond itself. For the rule 2, instead, one finds that $`x(t+1)>x(t)`$, if $`x(t)`$ is larger than the double of the average minimal threshold, and, due to the extremal nature of the dynamics, this is verified always in the large $`N`$ limit, i.e. in the limit of a large number of bonds in the interface at any time-step. This means that in this second case, damaging is not strong enough to allow a global weakening of the system, which becomes more and more rigid. This is reasonable since in rule 2 the stress on the weakest bond is redistributed to the nearest neighbors, and the total initial stress is conserved, while in rule 1 there is not total stress conservation. In other words, in the model with rule 1 the damaging is a multiplicative effect, i.e. the damaging is proportional to the old threshold (which can be big), in the model with rule 2 the damaging is quite reduced by the fact that at each time-step it is proportional to the minimal threshold in the whole system. In Figs. 6b, 7b the time evolution of $`x(t)`$ obtained from computer simulations is compared with the theoretical prediction. Our analytical results are in good agreement with numerical simulations. For rule 2, numerical simulations of the histogram evidentiate a low $`x`$ tail below the critical threshold, which tends to disappear as the system size grows, and a non zero slope of the part just above the critical threshold. The first one is a clear finite size effect, which is less important in the simulations of rule 1, because for rule 1 the critical threshold is very small. Of course, such a finite size effect is absent in the theoretical results, as in all mean field (MF) approaches. The second effect could be due to spatial correlation induced by the damaging rule 2, which in the analytical approach are neglected. This second effect does not disappear as the system size grows. Consequently the agreement between the numerical simulations of $`x(t)`$ and Eq. (31) is less good than for rule 1. The numerical $`x(t)`$, mainly because of the nonzero negative slope of $`\varphi (x)`$above $`x_c`$, is a bit smaller than the theoretical prediction. With respect to real fracturing processes the behavior of the average resistance $`x(t)`$ obtained with rule 2 is more realistic, since in real materials one usually observes that the material during micro-cracks formation becomes more rigid, although more fragile, since the number of bond one has to break to have global breakdown becomes smaller and smaller. Moreover it is worth to note that, apart the shape of $`\varphi (x)`$ and the behavior of $`x(t)`$, all the other statistical properties of the system do not depend on the used weakening rule. Finally, it is worth to point out that, to our knowledge, apart the qualitative results of , no quantitative experimental results are available. For example, a measurement of the fractal dimension of cracks or their size distribution would be extremely useful to further test the predictions of this model. At the moment, this model seems able to capture, with its extremely simplified dynamics, some basic properties of fracturing processes. In conclusion, we have presented a new model for fractures, which is useful in describing in a semi-quantitative way some basic mechanism in drying paint-like and mud-like cracking processes, for extremely thin samples. Due to its extreme simplicity, the model is particularly suitable for large scale simulations and takes into account the damaging effects involved in fracture propagation. Even in this simple model we are able to analyze which conditions trigger SOC behavior in such systems. Furthermore, the change in the threshold distribution, induced by the damaging mechanism, allows us to write down explicitly the form of the breakdown probability for the bonds of the sample. Possible further research could consist, for damaging rule 2, in a more refined calculational scheme, in which two variable probability densities are also considered. This would be the first order correction to our MF approach considering only one-variable distributions, and could allow us to take into account correlations induced by the damaging rule. Such a generalization of the RTS theory, formally discussed in , is however technically very difficult. Another research direction we are following is the application of real space techniques, combined with the RTS approach, to calculate the critical exponents of the model. The authors acknowledge the support of the EU grant Contract No. FMRXCT980183. ## A RTS for the damaged system In this appendix a simple explication of the RTS probabilistic equations is provided. First of all, one has to note that in this kind of models (as well as in Invasion Percolation) the initial condition of the system is characterized by independent variables (the breaking thresholds of the bonds) identically and uniformly distributed. However, once the minimal value in the set is found and the relative bond broken, the knowledge of this event makes the variables of the remaining non-broken bonds no longer simply uniformly distributed in the interval $`[0,1]`$, and correlated (no more independent one each other). In fact, after the breaking of the bond with the minimal threshold, one has to condition the probability of any further event to the last known event. This information influences the probability distribution of the remaining bonds of the system and creates correlations among them . The systematic study of this “memory” effect is what is called Run Time Statistics . In order to clarify the “step by step” mechanism of storage of conditional information, let us think to have fixed a time-order path $`A_t`$, i.e. an history of the dynamics up to time $`t`$. $`A_t`$ is given by the time ordered sequence $`\{i_0,i_2,\mathrm{},i_{t1}\}`$ of the broken bonds up to time $`t`$. Let us suppose to know the joint threshold probability density function $`P_t(\{x\}_{C_t}|A_t)`$ of the whole set of non-broken bonds conditioned to the knowledge of the past history $`A_t`$. $`P_t(\{x\}_{C_t}|A_t)`$ represents the “effective” distribution of the disorder at the $`t^{th}`$ time-step of a fixed history $`A_t`$. Note that at time $`t=0`$, one has: $$P_0(\{x\}_{C_0})=\underset{kS}{}p_0(x_k)=1,$$ (A1) where $`S`$ is the whole lattice, as no information from the dynamics is still present. Since any kind of “order” relation, superimposed to a set of independent stochastic variables, introduces correlations, in general $`P_t(\{x\}_{C_t}|A_t)`$ does not factorize in the product of single-bond “effective” density functions for $`t>0`$ . That is, it is not possible to write: $$P_t(\{x\}_{C_t}|A_t)=\underset{kC_t}{}p_{k,t}(x_k).$$ (A2) However, as shown in , in the limit of large number of variables the “geometrical” correlations in $`P_t(\{x\}_{C_t}|A_t)`$ become negligible, and one can make, at any time step, the approximation given by Eq. (A2). Therefore, we consider the approximated case where the “effective” probability density function of the disorder of the system, with all the information about the past history stored, is given by the set of “effective” one-bond functions $`p_{k,t}(x)`$ for each non-broken bond $`k`$. The rigorous exposition of RTS, by using the non-factorizable function $`P_t(\{x\}_{C_t}|A_t)`$ at each $`t`$ is given in . Knowing the set of functions $`p_{k,t}(x)`$, one can write the “effective” probability $`\mu _{i,t}`$ that a given bond $`i`$ of the set is broken at that time. It is simply the probability, conditioned to the whole past history, that $`x_i`$ is the minimum in the set of non-broken bond variables. Consequently, it is given by Eq. (1), i.e. $$\mu _{i,t}=_0^1𝑑xp_{i,t}(x)\left[\underset{k(i)}{\overset{C_t}{}}_x^1𝑑yp_{k,t}(y)\right].$$ (A3) The set of $`\mu _{i,t}`$ for each non-broken bond and for each time-step, defines a branching process of the dynamics; i.e. each history $`A_t`$ at time $`t`$ continues with a certain probability $`\mu _{i,t}`$ in a different history $`A_{t+1}`$ at time $`t+1`$ for each breaking bond $`i`$ at time $`t`$. In order to continue the probabilistic description of the branching at further time-steps, one should obtain the new set of functions $`p_{k,t+1}(x)`$ for these different cases of breaking at time $`t`$, using only the “old” set of $`p_{k,t}(x)`$ and the set of probabilities $`\mu _{i,t}`$ defining the branching. This is possible by using the notions of conditional probability. Here the simple rule relating the conditional to joint probability of a first event $`A`$ to a second event $`B`$ is reminded : $$\text{Prob}(A|B)=\frac{\text{Prob}(AB)}{\text{Prob}(B)},$$ (A4) where, as usual, $`A|B`$ means the event $`A`$ conditioned to the event $`B`$, while $`AB`$ the event $`A`$ joint to the event $`B`$. Note that “memory” up to time $`t`$, for a fixed history $`A_t`$ in the branching of all the possible histories, is already stored in the functions $`p_{k,t}(x)`$. Consequently, in order to obtain the set of probability functions $`p_{k,t+1}(x)`$ for the history $`A_{t+1}`$ obtained from $`A_t`$ adding the breaking of bond $`i`$ at time $`t`$, one has to store only information about the last step. At this point one has to distinguish the three cases: (1) the just broken bond $`i`$, (2) a non-broken bond $`j`$ far from $`i`$, and (3) a non-broken neighbor $`l`$ of $`i`$. (1) In this case let us call the conditioned probability density of bond $`i`$ after its braking with $`m_{i,t+1}(x)`$ instead of $`p_{i,t+1}(x)`$, remarking with this that after its breakdown, bond $`i`$ is removed definitely from the interface. Note that, since after $`t`$ the bond $`i`$ does not participate to the dynamics, its “effective” probability density will not change anymore. $`m_{i,t}(x)dx`$ is the probability that $`x<x_ix+dx`$, conditioned to the the past history up to its breaking. However, since the memory up to the time-step just before its breaking is stored in the known functions $`p_{k,t}(x)`$, $`m_{i,t}(x)dx`$ is the probability, calculated using the set of functions$`\{p_{k,t}(x)\}`$, that $`x<x_ix+dx`$ (event $`A`$ of Eq. (A4)) conditioned to the fact that the bond $`x_i`$ is the minimum in the set of interface bonds at that time (event $`B`$ of Eq. (A4)). Therefore, from Eq. (A4), one has Eq. (2): $$m_{i,t+1}(x)=\frac{1}{\mu _{i,t}}p_{i,t}(x)\left[\underset{k(i)}{\overset{C_t}{}}_x^1𝑑yp_{k,t}(y)\right].$$ (A5) In a quite similar way we can update the effective probability densities for the cases (2) and (3). In the case (2), using the set of functions $`\{p_{k,t}(x)\}`$, $`p_{j,t+1}(x)dx`$ is the probability that $`x<x_jx+dx`$ (event $`A`$) conditioned to the fact that $`x_i`$ was the minimal in the interface at time $`t`$. Again from Eq. (A4) one has Eq. (3): $$p_{j,t+1}(x)=\frac{1}{\mu _{i,t}}p_{j,t}(x)_0^x𝑑yp_{i,t}(y)\left[\underset{k(i,j)}{\overset{C_t}{}}_y^1𝑑zp_{k,t}(z)\right].$$ (A6) In the case (3) one has to distinguish the two different damaging rules, and the conditioning events are more complex. For rule 1, using the set of function $`\{p_{k,t}(x)\}`$, $`p_{l,t+1}(x)dx`$ is the probability that $`x<x_lx+dx`$ (event $`A`$) conditioned to the fact that $`x_i`$ was the minimum and that the value of $`x_l`$ at this time-step differs from the value at the previous time-step for a random fraction of itself (event $`B`$). One, then, gets Eq. (5): $`p_{j,t+1}(x)={\displaystyle \frac{1}{\mu _{i,t}}}{\displaystyle _0^1}dy{\displaystyle \frac{1}{y}}\theta (yx)p_{j,t}(y)\times `$ (A7) $`\times {\displaystyle _0^y}dzp_{i,t}(z)\left[{\displaystyle \underset{k(i,j)}{\overset{C_t}{}}}{\displaystyle _z^1}dup_{k,t}(u)\right].`$ (A8) Finally, for rule 2, always using the set of functions $`\{p_{k,t}(x)\}`$, $`p_{l,t+1}(x)dx`$ is the probability that $`x<x_lx+dx`$ (event $`A`$) conditioned to the fact that $`x_i`$ was the minimum and that the value of $`x_l`$ at this time-step differs from the value at the previous time-step for a fraction $`1/n_{i,t}`$ of $`x_i`$ (event $`B`$). From this one has eq. (8): $`p_{j,t+1}(x)=`$ (A9) $`={\displaystyle \frac{1}{\mu _{i,t}}}{\displaystyle _0^1}dy\left[{\displaystyle \underset{k(i,j)}{\overset{C_t}{}}}{\displaystyle _y^1}dzp_{k,t}(z)\right]p_{i,t}(y)p_{j,t}(x+{\displaystyle \frac{y}{n_{i,t}}})\times `$ (A10) $`\times \theta \left({\displaystyle \frac{n_{i,t}}{n_{i,t}1}}xy\right)\theta \left(n_{i,t}(1x)y\right).`$ (A11)
warning/0004/hep-th0004173.html
ar5iv
text
# 1 Introduction ## 1 Introduction In recent years quantum mechanics on sphere $`S^n`$ have been studied in various aspects. Interestingly, the gauge fields are seen to emerge at the quantum level, which in turn specify the possible inequivalent quantizations and they are coupled with a particle constrained to move on the sphere. In ref. we have presented a picture where the wave functions are constrained to the sphere by the “quantum constraint”. Namely we utilize, à la Dirac, a square root of the usual on sphere constraint. In this approach, the induced gauge fields appear in the Hopf map from the wave function space to the real space. In this note we show the following : 1. By considering the path integral quantization on $`S^2`$ through the use of the wave functions mentioned above, we show that the Berry’s connection appears as a topological term in the effective action. 2. The topological term is nothing but the Chern-Simons term of gauge variables, which appear when we enlarge our space $`S^2`$ to $`SO(3)`$, by the use of the relation $`S^2=SO(3)/SO(2)`$. The gauge variables absorb the extra degrees of freedom of the enlarged space. 3. We generalize the arguments in 1. and 2. to arbitrary sphere $`S^n`$. We recapitulate in section 2 what we have done in the previous paper. Then using the wave function introduced in section 2 we perform the path integral quantization and obtain the effective action on $`S^2`$ in section 3. We shall be concerned with the cases $`n=3,4`$ in section 4. In section 5 we consider the general case. Section 6 is devoted to discussions. ## 2 Hopf Map and Quantization on a Sphere In the previous paper we have quantized a system constrained to move on a sphere by considering a square root of the “on sphere condition” and have arrived at the fibre bundle structure of the Hopf map in the cases of $`S^2`$and $`S^4`$. This leads to more geometrical understanding of monopole and instanton gauge structures that emerge in the course of quantization. We have seen that square root of the “on sphere condition” $`x_M^2r^2=0`$ can be written as $$\left(x_M\mathrm{\Gamma }_Mr\right)\mathrm{\Phi }\left(\stackrel{}{x}\right)=0,$$ (1) where $`\mathrm{\Gamma }_M`$ are the Pauli matrices for $`S^2`$ and are the Dirac matrices for $`S^4`$ . A solution of eq.(1) is expressed as $$\mathrm{\Phi }=v\varphi ,$$ (2) with $`\varphi `$ an arbitrary complex function on the sphere. The solution to the constraint is the space projected by $`P`$ $$P\mathrm{\Phi }=\mathrm{\Phi },$$ (3) where the projection operator $`P`$ to the space spanned by $`v`$ is defined as $`P`$ $``$ $`vv^{},`$ (4) $`P^2`$ $`=`$ $`vv^{}vv^{}=P,`$ $`Pv`$ $`=`$ $`v.`$ The explicit form of $`v`$ can be written as $$v=\frac{1}{\sqrt{2r\left(r+x_3\right)}}\left(\begin{array}{c}r+x_3\\ x_1+ix_2\end{array}\right),$$ (5) for $`S^2`$ and $$v=\frac{1}{\sqrt{2r\left(r+x_5\right)}}\left(\begin{array}{c}r+x_5\\ x_4i\stackrel{}{x}\stackrel{}{\sigma }\end{array}\right),$$ (6) for $`S^4`$ . Then evaluation of the projected derivative $`P\mathrm{\Phi }`$, through $`P\mathrm{\Phi }`$ $`=`$ $`vv^{}\left(v\varphi \right)`$ $`=`$ $`vv^{}\left(v\varphi +v\varphi \right)`$ $`=`$ $`vD\varphi ,`$ where $`D`$ $``$ $`+A,`$ (8) $`A`$ $``$ $`v^{}v,`$ leads to the induced magnetic monopole gauge potential for $`S^2`$ $$Av^{}dv=\frac{i}{2r\left(r+x_3\right)}\left(x_1dx_2x_2dx_1\right),$$ (9) which was obtained in refs.,, and the instanton gauge potential for $`S^4`$ $$A=i\frac{1}{2r\left(r+x_5\right)}\sigma _{\mu \nu }x_\mu dx_\nu ,$$ (10) discussed in refs.,. ## 3 Path Integral Quantization on $`S^2`$ According to the arguments of the previous section, the wave function on $`S^2`$ can be written as $$\mathrm{\Phi }=v\varphi .$$ (11) Using this wave function we calculate the transition amplitude $`Z`$ $``$ $`\stackrel{}{x},v;t_f|\stackrel{}{x},v;t_i`$ (12) $`=`$ $`{}_{f}{}^{}\stackrel{}{x},v|e^{iH(t_ft_i)}|\stackrel{}{x},v_{i}^{}`$ $`=`$ $`{\displaystyle [d\stackrel{}{x}]v_f|\stackrel{}{x}_f|e^{iH\mathrm{\Delta }t}|\stackrel{}{x}_{N1}|v_{N1}v_{N1}|\stackrel{}{x}_{N1}|e^{iH\mathrm{\Delta }t}|\stackrel{}{x}_{N2}|v_{N2}}`$ $`\mathrm{}v_1|\stackrel{}{x}_1|e^{iH\mathrm{\Delta }t}|\stackrel{}{x}_i|v_i,`$ which by use of $$v_k|\stackrel{}{x}_k|e^{iH\mathrm{\Delta }t}|\stackrel{}{x}_{k1}|v_{k1}=[d\stackrel{}{p}]v_k|e^{i(p_k\frac{\mathrm{\Delta }x_k}{\mathrm{\Delta }t}H(x_k))\mathrm{\Delta }t}|v_{k1},$$ (13) can be written as $$\stackrel{}{x},v;t_f|\stackrel{}{x},v;t_i=[d\stackrel{}{x}][d\stackrel{}{p}]e^{i{\scriptscriptstyle (\stackrel{}{p}\dot{\stackrel{}{x}}H(\stackrel{}{x}))𝑑t}}v_f|v_{N1}v_{N1}|v_{N2}\mathrm{}v_1|v_i,$$ (14) where, for example, $`[d\stackrel{}{x}]{\displaystyle \underset{i=1}{\overset{3}{}}}dx_i\delta ({\displaystyle \underset{j=1}{\overset{3}{}}}x_j^2r^2)`$. Furthermore, if we express $$v_k|v_{k1}1v|_\stackrel{}{x}|v\mathrm{\Delta }\stackrel{}{x}e^{i\omega },$$ we have $$\stackrel{}{x},v;t_f|\stackrel{}{x},v;t_i=[d\stackrel{}{x}][d\stackrel{}{p}]e^{i{\scriptscriptstyle (\stackrel{}{p}\dot{\stackrel{}{x}}H(\stackrel{}{x}))𝑑t}}e^{i{\scriptscriptstyle \omega }},$$ (15) where $$\omega =iv|_\stackrel{}{x}|v\mathrm{\Delta }\stackrel{}{x}=\frac{ϵ_{3ij}}{2r(r+x_3)}x_i\dot{x}_j\mathrm{\Delta }tA_0\mathrm{\Delta }t.$$ Consequently, to the original action $$S_0=(\stackrel{}{p}\dot{\stackrel{}{x}}H(\stackrel{}{x}))𝑑t,$$ (16) we have to add $$S^{}=A_0𝑑t.$$ (17) Which implies emergence of the geometrical term added to the original Lagrangian and the induced Lagrangian should be $$L_{S^2}=L_{S^2}^0+A_0.$$ (18) In order to see the meaning of the induced term, let us consider $`H`$-covariant formulation of Lagrangian on the coset space $`G/H`$ , which in our case is $`S^2=SO(3)/SO(2)`$. We can express the original Lagrangian in terms of $`SO(3)`$ variables $`U=u_0+i\stackrel{}{\sigma }\stackrel{}{u}`$ , $$L_{SO(3)/SO(2)}^0=r^2\mathrm{Tr}[DU(DU)^{}],$$ (19) where $`D{\displaystyle \frac{d}{dt}}+i𝒜{\displaystyle \frac{\sigma _3}{2}}`$ and $`𝒜`$ is a “gauge variable” that compensates the redundant freedom of $`SO(2)`$ contained in $`U`$ . We can see the equivalence of this Lagrangian with $`L_{S^2}^0`$ by integrating out $`𝒜`$ in eq.(19), i.e. $$L_{SO(3)/SO(2)}^0\frac{r^2}{4}\mathrm{Tr}\left[\frac{d}{dt}(U^{}\sigma _3U)\right]^2=\frac{1}{2}\left(\dot{\stackrel{}{x}}\right)^2=L_{S^2}^0,$$ (20) since $`U^{}\sigma _3U={\displaystyle \frac{\stackrel{}{x}}{r}}\stackrel{}{\sigma }`$ and $`\stackrel{}{x}^2=r^2`$. The induced Lagrangian (18) can be written as $$L_{SO(3)/SO(2)}=L_{SO(3)/SO(2)}^0+k𝒜,$$ (21) in $`H`$-covariant form, where $`k=1/2`$ . Indeed, integrating out $`𝒜`$ once again we arrive at $$L_{SO(3)/SO(2)}L_{S^2}^0i\frac{1}{2}\mathrm{Tr}(\dot{U}U^{}\sigma _3)\frac{1}{8r^2},$$ (22) which is equivalent to $`L_{S^2}`$ (18) up to total divergence and constant term. That is, when the system is described in terms of $`SO(3)`$ variables, a term proportional to the “gauge variable” $`𝒜`$, which was introduced in order to absorb the extra degrees of freedom, is induced. This term, which has topological origin, can be considered as a $`(0+1)`$-dimensional Chern-Simons term. ## 4 $`S^3`$ and $`S^4`$ The $`S^3`$ and $`S^4`$ cases go along the same line. In the case of $`S^4`$, the quantum constraint can be written as $$(x_M\gamma ^Mr)|v=0,$$ (23) where $$\begin{array}{cc}\gamma ^M\hfill & =(\gamma ^1,\gamma ^2,\gamma ^3,\gamma ^4,\gamma ^5),\hfill \\ \stackrel{}{\gamma }\hfill & =\left(\begin{array}{cc}0\hfill & i\stackrel{}{\sigma }\hfill \\ i\stackrel{}{\sigma }\hfill & 0\hfill \end{array}\right),\gamma ^4=\left(\begin{array}{cc}0\hfill & 1\hfill \\ 1\hfill & 0\hfill \end{array}\right),\gamma ^5=\left(\begin{array}{cc}1\hfill & 0\hfill \\ 0\hfill & 1\hfill \end{array}\right),\hfill \\ \stackrel{}{\gamma }\hfill & =(\gamma ^1,\gamma ^2,\gamma ^3).\hfill \end{array}$$ (24) Solution of the equation (23) is $$|v=\frac{1}{\sqrt{2r(r+x_5)}}\left(\begin{array}{c}r+x_5\\ x_4i\stackrel{}{x}\stackrel{}{\sigma }\end{array}\right),$$ (25) and the transition amplitude can be written as $$Z_{S^4}=[dx_M][dp_M]\mathrm{exp}\left[i\left(p_N\dot{x}_NH_{S^4}^0\right)𝑑t\right]\mathrm{Tr}\mathrm{exp}\left[A_{S^4}^0𝑑t\right],$$ (26) where $$A_{S^4}^0=\frac{1}{2r(r+x_5)}\sigma _{\mu \nu }x_\mu \dot{x}_\nu .$$ (27) That is, a coupling with the instanton solution is induced. As for the case of $`S^3`$, the transition amplitude that follows from the wave function is nothing but eq.(26) with $`x_5=0`$, that is, $$Z_{S^3}=[dx_\mu ][dp_\mu ]\mathrm{Tr}\mathrm{exp}\left[i\left(p_\nu \dot{x}_\nu H_{S^3}^0\right)𝑑t\right]\mathrm{Tr}\mathrm{exp}\left[A_{S^3}^0𝑑t\right],$$ (28) the additional term being a coupling with the meron solution $$A_{S^3}^0=\frac{1}{2r^2}\sigma _{\mu \nu }x_\mu \dot{x}_\nu .$$ (29) As in the case for $`S^2`$ , let us consider $`H`$-covariant formulation of the transition amplitude on the coset space $`G/H`$ , which is $`SO(4)/SO(3)`$ in the case for $`S^3`$ and is $`SO(5)/SO(4)`$ in the case for $`S^4`$. Let us start with the case of $`S^3=SO(4)/SO(3)`$. We write the $`SO(4)`$ element in the block diagonal $`4\times 4`$ matrix form $$SO(4)G_4=\left(\begin{array}{cc}e^{i\mathrm{\Theta }_{\mu \nu }\sigma _{\mu \nu }}& 0\\ 0& e^{i\overline{\mathrm{\Theta }}_{\mu \nu }\overline{\sigma }_{\mu \nu }}\end{array}\right)\left(\begin{array}{cc}g& 0\\ 0& \overline{g}\end{array}\right),$$ (30) where $`\sigma _{ij}=\overline{\sigma }_{ji}=\frac{1}{2}ϵ_{ijk}\sigma _k,\sigma _{i4}=\overline{\sigma }_{i4}=\frac{1}{2}\sigma _i`$. The Lagrangian on $`S^3`$ can be written in terms of $`G_4`$ as $$L_{SO(4)/SO(3)}^0=\frac{r^2}{2}\mathrm{Tr}(G_4^1DG_4(G_4^1DG_4)^{}),$$ (31) where $`D{\displaystyle \frac{d}{dt}}+i𝒜^{SO(3)}`$ and $`𝒜^{SO(3)}`$ is the gauge variable that absorbs the extra $`SO(3)`$ degrees of freedom, which we write as $$𝒜^{SO(3)}\left(\begin{array}{cc}𝒜_i\frac{\sigma _i}{2}& 0\\ 0& 𝒜_i\frac{\sigma _i}{2}\end{array}\right).$$ (32) We find that the Lagrangian (31) is equivalent to the naive Lagrangian for a particle on $`S^3`$. Indeed, integrating out $`𝒜^{SO(3)}`$ we arrive at the following Lagrangian on $`S^3`$ $`L_{SO(4)/SO(3)}^0`$ $``$ $`{\displaystyle \frac{r^2}{4}}\mathrm{Tr}(g\dot{g}^1\overline{g}\dot{\overline{g}}^1)^2`$ $`=`$ $`{\displaystyle \frac{r^2}{4}}\mathrm{Tr}\left[{\displaystyle \frac{d}{dt}}(\overline{g}^1g){\displaystyle \frac{d}{dt}}(g^1\overline{g})\right]`$ $`=`$ $`{\displaystyle \frac{r^2}{4}}\mathrm{Tr}\dot{Q}_3\dot{Q}_3^1={\displaystyle \frac{1}{2}}\dot{x}_\mu ^2=L_{S^3}^0,`$ where $`Q_3\overline{g}^1g={\displaystyle \frac{x_4}{r}}+i{\displaystyle \frac{x_i}{r}}\sigma _i`$ and $`{\displaystyle \underset{\mu =1}{\overset{4}{}}}x_\mu ^2=r^2.`$ Next we add a term $`\mathrm{Tr}(K𝒜^{SO(3)})`$ to the Lagrangian, where the constant $`K`$ is given by $$K=\left(\begin{array}{cc}k_i\sigma _i& 0\\ 0& k_i\sigma _i\end{array}\right).$$ (34) Then, integrating out $`𝒜^{SO(3)}`$, we find that the Lagrangian, $$L_{SO(4)/SO(3)}=L_{SO(4)/SO(3)}^0+\mathrm{Tr}(K𝒜^{SO(3)}),$$ (35) goes to $$L_{SO(4)/SO(3)}L_{S^3}^0i\mathrm{Tr}\left[G_4\dot{G}_4^1K\right]\frac{1}{2r^2}\mathrm{Tr}(K^2).$$ (36) Furthermore, we can show that the transition amplitude $`Z_{SO(4)/SO(3)}`$ corresponding to this Lagrangian is identical to $`Z_{S^3}`$. As we can write $$G_4=h^1\left(\begin{array}{cc}\overline{g}^1g& 0\\ 0& 1\end{array}\right),$$ (37) we have $$G_4\dot{G}_4^1=h^1(4iA_{S^3}^0)h+h^1\dot{h},$$ (38) where $`h`$ is an $`SO(3)`$element written in $`4\times 4`$ block diagonal matrix form. Thus the Lagrangian is expressed as $$L_{SO(4)/SO(3)}L_{S^3}^0i\mathrm{Tr}(Kh^1\dot{h})4\mathrm{T}\mathrm{r}(hKh^1A_{S^3}^0)+\mathrm{const}..$$ (39) We note that this Lagrangian has the same form as that obtained in ref.. Here if we define $$\left(\begin{array}{cc}S^i\sigma _i& 0\\ 0& S^i\sigma _i\end{array}\right)\frac{1}{2}h^1\left(\begin{array}{cc}\sigma _3& 0\\ 0& \sigma _3\end{array}\right)h,$$ (40) the corresponding Hamiltonian can be written as $$H=H_{S^3}^02S^i(A_{S^3}^0)^i,$$ (41) where $`S^i`$ is a spin variable that satisfies $`[S^i,S^j]=iϵ^{ijk}S^k.`$Based on this Hamiltonian with $`k_i=\frac{1}{4}\delta _{i3},`$ we derive the transition amplitude by integrating the spin degrees of freedom, $$Z_{SO(4)/SO(3)}=[dx_\mu ][dp_\mu ]\mathrm{exp}\left[i\left(p_\nu \dot{x}_\nu H_{S^3}^0\right)𝑑t\right]\mathrm{Tr}\mathrm{exp}\left[iA_{S^3}^0𝑑t\right].$$ (42) We have confirmed this equation under the gauge condition $`(A_{S^3}^0)^1=(A_{S^3}^0)^2=0.`$ This expression coincides completely with the previous $`Z_{S^3}.`$ Next we turn to the discussion on $`S^4`$. The naive Lagrangian for the particle on $`S^4`$ in terms of $`SO(5)`$ variable $`G_5`$ can be written as $$L_{SO(5)/SO(4)}^0=\frac{r^2}{2}\mathrm{Tr}\left(G_5^1DG_5\left(G_5^1DG_5\right)^{}\right),$$ (43) where $`D{\displaystyle \frac{d}{dt}}+i𝒜^{SO(4)}`$ is the $`SO(4)`$-covariant derivative. By integrating out $`𝒜^{SO(4)}`$, this Lagrangian can be seen to be equivalent to $`L_{S^4}^0`$. This may be explicitly shown in the representation such as $$𝒜^{SO(4)}=\left(\begin{array}{cc}𝒜_{\mu \nu }\sigma _{\mu \nu }& 0\\ 0& 𝒜_{\mu \nu }\overline{\sigma }_{\mu \nu }\end{array}\right)=\left(\begin{array}{cc}𝒜_+& 0\\ 0& 𝒜_{}\end{array}\right)𝒜_a^{SO(4)}T_a,$$ (44) where $`T_a`$ is the generator corresponding to the $`SO(4)`$ subgroup of $`SO(5)`$. We separate as $`G_5\dot{G}_5^1ig_{}^aT_a+ig_{}^\alpha T_\alpha G_{}+G_{}`$ ( $`T_\alpha `$ is the remaining generator of $`SO(5)`$ ). Integrating out $`𝒜^{SO(4)}`$, we see $$L_{SO(5)/SO(4)}^0\frac{r^2}{2}\mathrm{Tr}G_{}^2=\frac{1}{2}\dot{x}_M^2=L_{S^4}^0,$$ (45) where $`G_5^1\gamma ^5G_5={\displaystyle \underset{M=1}{\overset{5}{}}}{\displaystyle \frac{x_M}{r}}\gamma _M`$and$`{\displaystyle \underset{M=1}{\overset{5}{}}}x_M^2=r^2`$. Next we assume that the term $`\mathrm{Tr}(K𝒜^{SO(4)})`$ has been induced to the system on $`S^4`$ , where the constant $`K`$ is the algebra of $`SO(4)`$, the general form of which is given by $$K=\left(\begin{array}{cc}K_+& 0\\ 0& K_{}\end{array}\right)=K_aT_a.$$ (46) Then integrating out $`𝒜^{SO(4)}`$ we obtain the Lagrangian on $`S^4`$ $$L_{SO(5)/SO(4)}L_{S^4}^0i\mathrm{Tr}(G_5\dot{G}_5^1K)\frac{1}{2r^2}\mathrm{Tr}(K^2).$$ (47) As in the case of $`S^3`$, we calculate transition amplitude by means of the Hamiltonian $$H=H_{S^4}^0+2S^i(A_{S^4}^0)^i,$$ (48) which corresponds to the effective Lagrangian (47), and integrate out the spin variables $`S^i`$ defined by $$\left(\begin{array}{cc}0& 0\\ 0& S^i\sigma _i\end{array}\right)\frac{1}{2}\stackrel{~}{h}^1\left(\begin{array}{cc}0& 0\\ 0& \sigma _3\end{array}\right)\stackrel{~}{h},$$ (49) where $`\stackrel{~}{h}`$ is an element of the $`SO(4)`$ subgroup. Then, we arrive at $$Z_{SO(5)/SO(4)}=[dx_M][dp_M]\mathrm{exp}\left[i\left(p_N\dot{x}_NH_{S^4}^0\right)𝑑t\right]\mathrm{Tr}\mathrm{exp}\left[iA_{S^4}^0𝑑t\right],$$ (50) where the choice has been used of $`K_+=0,K_{}=\frac{1}{2}\sigma _3.`$ Thus, we see that for both cases $`S^3`$and $`S^4`$, the $`(0+1)`$-dimensional Chern-Simons terms, that are proportional to the gauge variable, are induced in the course of quantization. ## 5 $`S^n`$ What has been argued so far can be generalized to all $`n`$ . First, we obtain the transition amplitude for a particle on $`S^n`$ that follows from the wave function with the “on sphere condition” taken into account. In order to do this, we consider the cases $`n=2m`$ and $`n=2m1`$ separately. We start with $`n=2m`$ case and prepare $`2^m\times 2^m`$ matrices $`\mathrm{\Gamma }_N^{(2m)}\left(N=1,2,\mathrm{},2m+1\right)`$ $$\{\mathrm{\Gamma }_M^{(2m)},\mathrm{\Gamma }_N^{(2m)}\}=2\delta _{MN},[\mathrm{\Gamma }_M^{(2m)},\mathrm{\Gamma }_N^{(2m)}]=2i\mathrm{\Sigma }_{MN}^{(2m)}.$$ (51) The explicit form of $`\mathrm{\Gamma }_N^{(2m)}`$ can be obtained as $`\mathrm{\Gamma }_i^{(2m)}`$ $`=`$ $`\sigma _2\mathrm{\Gamma }_i^{(2m2)}(i=1,\mathrm{},2m1),`$ $`\mathrm{\Gamma }_{2m}^{(2m)}`$ $`=`$ $`\sigma _11,`$ (52) $`\mathrm{\Gamma }_{2m+1}^{(2m)}`$ $`=`$ $`\sigma _31,`$ and $`SO(2m)`$ generators $`\mathrm{\Sigma }_{\mu \nu }^{(2m)}(\mu ,\nu =1,2,\mathrm{},2m)`$ are expressed in block diagonal form as $$\left(\begin{array}{cc}\mathrm{\Sigma }_{\mu \nu }^{(2m)+}& 0\\ 0& \mathrm{\Sigma }_{\mu \nu }^{(2m)}\end{array}\right).$$ (53) The spinor wave function $`|v`$ in $`2^m\times 2^{m1}`$ matrix form is constrained to satisfy the “on sphere condition” $$(x_N\mathrm{\Gamma }_N^{(2m)}r)|v=0.$$ (54) The non-trivial solution to the equation is given as $$|v=\frac{1}{\sqrt{2r(r+x_{2m+1})}}\left(\begin{array}{c}r+x_{2m+1}\\ x_{2m}ix_i\mathrm{\Gamma }_i^{(2m2)}\end{array}\right).$$ (55) With the help of $`|v`$ we perform the path integral to obtain the transition amplitude, $$Z_{S^{2m}}=[dx_M][dp_M]\mathrm{exp}\left[i(p_N\dot{x}_NH_{S^{2m}}^0)𝑑t\right]\mathrm{Tr}\mathrm{exp}\left[iA_{S^{2m}}^0𝑑t\right],$$ (56) where $$A_{S^{2m}}^0=\frac{1}{2r(r+x_{2m+1})}\mathrm{\Sigma }_{\mu \nu }^{(2m)+}x_\mu \dot{x}_\nu .$$ (57) Thus, we can claim that a coupling to the generalized instanton configuration is induced . The transition amplitude for a particle on $`S^{2m1}`$ is nothing but eq.(56) with $`x_{2m+1}=0`$. That is, $$Z_{S^{2m1}}=[dx_\mu ][dp_\mu ]\mathrm{exp}\left[i(p_\nu \dot{x}_\nu H_{S^{2m1}}^0)𝑑t\right]\mathrm{Tr}\mathrm{exp}\left[iA_{S^{2m1}}^0𝑑t\right],$$ (58) The additional term is a coupling to the generalized meron solution in arbitrary odd dimensions, $$A_{S^{2m1}}^0=\frac{1}{2r^2}\mathrm{\Sigma }_{\mu \nu }^{(2m)+}x_\mu \dot{x}_\nu .$$ (59) Next, exactly as in the previous discussions, having in mind that $`S^n=SO(n+1)/SO(n)`$ we can describe the system using the elements of $`SO(n+1)`$. This suggest that the induced term, which was obtained through the above mentioned path integration, appears as a term proportional to the “gauge variable” $`𝒜^{SO(n)}`$ that was introduced in order to absorb the extra degrees of freedom. Thus, for the description of quantum mechanics on $`S^n`$ using the $`SO(n+1)`$ variables, the gauge variable $`𝒜^{SO(n)}`$ is expected to be induced. Namely, we claim that the induced term is a term proportional to the “gauge variable” also for the general dimension $`n`$. ## 6 Discussions We have seen that, when describing quantum mechanics on the sphere in terms of the wave functions that satisfy the “square root” of “on sphere condition”, a particular gauge configuration appears in the transition amplitude due to the geometrical reasons. These results are consistent with those obtained in ref. from a different point of view. These configurations are monopoles and (generalized) instantons for even dimensional spheres and (generalized) merons for odd dimensional spheres. Furthermore, we have shown in this note that it is possible to interpret this situation as an induction of a term proportional to the “gauge variable” that was introduced in order to absorb the extra degrees of freedom, when we describe the $`S^n`$ system in terms of the $`SO(n+1)`$ variables according to the relation $`S^n=SO(n+1)/SO(n)`$. The induced terms have topological meaning and can be considered as a Chern-Simons terms in $`0+1`$ dimensions. If we extend this result, obtained in quantum mechanics, to the field theories where the fields are constrained to the sphere, we could expect an induction of Chern-Simons gauge fields. For example, for $`O(3)`$ sigma model in $`2+1`$ dimensions, fields are on $`S^2=SU(2)/U(1)`$ and we expect $`U(1)`$ Chern-Simons term to be induced in this case. This possibility was also suggested in refs.,.
warning/0004/cond-mat0004217.html
ar5iv
text
# Cooper pair dispersion relation for weak to strong coupling \[ ## Abstract Cooper pairing in two dimensions is analyzed with a set of renormalized equations to determine its binding energy for any fermion number density and all coupling assuming a generic pairwise residual interfermion interaction. Also considered are Cooper pairs (CPs) with nonzero center-of-mass momentum (CMM)—usually neglected in BCS theory—and their binding energy is expanded analytically in powers of the CMM up to quadratic terms. A Fermi-sea-dependent linear term in the CMM dominates the pair excitation energy in weak coupling (also called the BCS regime) while the more familiar quadratic term prevails in strong coupling (the Bose regime). The crossover, though strictly unrelated to BCS theory per se, is studied numerically as it is expected to play a central role in a model of superconductivity as a Bose-Einstein condensation of CPs where the transition temperature vanishes for all dimensionality $`d2`$ for quadratic dispersion, but is nonzero for all $`d1`$ for linear dispersion. PACS #: 74.20.Fg; 64.90+b; 05.30.Fk; 05.30.Jp \] The original Cooper pair (CP) problem in two (2D) and three (3D) dimensions possesses ultraviolet divergences in momentum space that are usually removed via interactions regularized with large-momentum cutoffs . One such regularized potential is the BCS model interaction which is of great practical use in studying Cooper pairing and superconductivity . Although there are controversies over the precise pairing mechanism, and thus over the microscopic Hamiltonian appropriate for high-$`T_c`$ superconductors, some of the properties of these materials have been explained satisfactorily within a BCS-Bose crossover picture via a renormalized BCS theory for a short-range interaction. In the weak-coupling limit of the BCS-Bose crossover description one recovers the pure mean-field BCS theory of weakly-bound, severely-overlapping CPs. For strong coupling (and/or low density) well-separated, nonoverlapping (so-called “local”) pairs appear in what is known as the Bose regime. It is of interest to detail how renormalized Cooper pairing itself evolves independently of the BCS-Bose crossover picture in order to then discuss the possible Bose-Einstein (BE) condensation (BEC) of such pairs. We address this here in a single-CP picture, while considering also the important case (generally neglected in BCS theory) of non-zero center-of-mass-momentum (CMM) CPs that are expected to play a significant role in BE condensates at higher temperatures. In this Report we derive a renormalized Cooper equation for a pair of fermions interacting via either a zero- or a finite-range interaction. We find an analytic expression for the CP excitation energy up to terms quadratic in the CMM which is valid for any coupling. For weak coupling only the linear term dominates, as it also does for the BCS model interaction . The linear term was mentioned for 3D as far back as 1964 (Ref. p. 33). For strong coupling we now find that the quadratic term dominates and is just the kinetic energy of the strongly-bound composite pair moving in vacuum. The CP dispersion relation enters into each summand in the BE distribution function of the boson number equation from which the critical temperature $`T_c`$ of BEC of CPs is extracted. The linear dependence on the CMM of CP binding for weak coupling leads to novel transition properties even in a heuristic BEC picture of superconductivity as BE-condensing CPs. It is well-known that BEC is possible only for dimensionalities $`d>2`$ for usual nonrelativistic bosons with quadratic dispersion; this limitation reappears in virtually all BEC schemes thus far applied to describe superconductivity (Refs. among others). But for bosons with a linear dispersion relation as found here in weak coupling, BEC can now occur for all $`d>1`$. Here we discuss the CP dispersion relation only in 2D. We have also performed a similar analysis in 3D and obtained the linear to quadratic crossover in the dispersion relation. Consider a two-fermion system in the CM frame, with each fermion of mass $`m`$ , interacting via the purely attractive short-range separable potential $$V_{pq}=v_0g_pg_q,$$ (1) where $`v_00`$ is the interaction strength and the $`g_p`$’s are the dimensionless form factors $`g__p(1+p^2/p_0^2)^{1/2}`$, where the parameter $`p_0`$ is the inverse range of the potential so that, e.g.,$`p_0\mathrm{}`$ implies $`g__p=1`$ and corresponds to the contact or delta potential $`V(r)=v_0\delta (𝐫)`$. The interaction model (1) mimics a wide variety of possible dynamical mechanisms in superconductors: mediated by phonons, or plasmons, or excitons, or magnons, etc., or even a purely electronic interaction. In the first instance mentioned, two terms of the form (1) can simulate a coulombic interfermion repulsion surrounded by a longer-ranged electron-phonon attraction. The momentum-space Schrödinger eigenvalue equation for a two-particle bound state in vacuum with binding energy $`B_20`$ for interaction (1) is $$\frac{1}{v_0}=\underset{k}{}\frac{g_k^2}{B_2+\mathrm{}^2k^2/m},\text{ }$$ (2) where $`k`$ is the wavenumber in the CM frame and $`\mathrm{}^2k^2/2m`$ the single-particle energy. On the other hand, the CP equation for two fermions above the Fermi surface with momenta wavevectors $`𝐤_1`$ and $`𝐤_2`$ (and arbitrary CMM wavevector $`𝐊𝐤_1+𝐤_2`$) is given by $$[\frac{\mathrm{}^2k^2}{m}E_K+\frac{\mathrm{}^2K^2}{4m}]C_k=_q^{^{}}V_{kq}C_q,$$ (3) where $`𝐤\frac{1}{2}(𝐤_1𝐤_2)`$ is the relative momentum, $`E_K2E_F\mathrm{\Delta }_K`$ the total pair energy, $`\mathrm{\Delta }_K0`$ the CP binding energy, $`C_qq\mathrm{\Psi }`$ its wave function in momentum space, and the prime on the summation implies restriction to states above the Fermi surface: viz., $`|𝐤\pm 𝐊/2|>k_F`$. For the separable interaction (1) Eq. (3) becomes $$\underset{k}{}{}_{}{}^{^{}}\frac{g_k^2}{\mathrm{}^2k^2/m+\mathrm{\Delta }_K2E_F+\mathrm{}^2K^2/4m}=\frac{1}{v_0}.$$ (4) Although the summand in Eq. (4) is angle-independent, the restriction on the sum arising from the filled Fermi sea is a function of the relative wave vector $`𝐤`$, and therefore angle-dependent. The potential strength $`v_0`$ can be eliminated between Eqs. (2) and (4) leading to the renormalized CP equation $`{\displaystyle \underset{k}{}}`$ $`{\displaystyle \frac{g_k^2}{B_2+\mathrm{}^2k^2/m}}`$ (5) $`=`$ $`{\displaystyle \underset{k}{}}{}_{}{}^{^{}}{\displaystyle \frac{g_k^2}{\mathrm{}^2k^2/m+\mathrm{\Delta }_K2E_F+\mathrm{}^2K^2/4m}},`$ (6) Instead of the arbitrary cutoff usually employed in dealing with delta interactions, in Eq. (6) we rely on physical “observables” for the sake of renormalization, viz., the ground-state binding energy $`B_2`$ in vacuum. The sums in Eq. (6) can be transformed to integrals; the restriction in the second term arising from the filled Fermi sea leads to two different expressions depending on whether $`\stackrel{~}{K}K/k_F`$ is $`<2`$ or $`>2`$, as discussed in the Appendix. Letting all variables be dimensionless by expressing them either in units of the Fermi wavenumber $`k_F`$ or of the Fermi energy $`E_F\mathrm{}^2k_F^2/2m`$, viz., $`\xi k/k_F`$, $`\stackrel{~}{B}_2B_2/E_F`$, $`\stackrel{~}{\mathrm{\Delta }}_K\mathrm{\Delta }_K/E_F`$, etc., we define $`\alpha _K^21\stackrel{~}{\mathrm{\Delta }}_K/2\stackrel{~}{K}^2/4\beta _K^2`$, and $`\theta `$ the angle between wavevectors $`𝐤`$ and $`𝐊`$ so that $`\xi _0(\theta )\sqrt{1\stackrel{~}{K}^2\mathrm{sin}^2\theta /4}+\stackrel{~}{K}\mathrm{cos}\theta /2`$ and $`\xi _0^{}(\theta )\sqrt{1\stackrel{~}{K}^2\mathrm{sin}^2\theta /4}+\stackrel{~}{K}\mathrm{cos}\theta /2`$. For a zero-range interaction, $`g_k=1`$, after some algebra one gets $$_0^{\pi /2}𝑑\theta \mathrm{ln}[\xi _0^2(\theta )\alpha _K^2]=\frac{\pi }{2}\mathrm{ln}\frac{\stackrel{~}{B}_2}{2},\stackrel{~}{K}<2,$$ (7) $$\frac{\pi }{2}\mathrm{ln}[\beta _K^2]_0^{\theta _0}𝑑\theta \mathrm{ln}\frac{\xi _0^2(\theta )+\beta _K^2}{\xi _0^2(\theta )+\beta _K^2}=\frac{\pi }{2}\mathrm{ln}\frac{\stackrel{~}{B}_2}{2},\stackrel{~}{K}>2,$$ (8) where $`\theta _0=\mathrm{arcsin}(2/\stackrel{~}{K})<\pi /2`$. For $`\stackrel{~}{K}=0`$ only Eq. (7) applies, in which case $`\xi _0(\theta )=1,`$ $`\alpha _K^2\alpha _0^2=1\stackrel{~}{\mathrm{\Delta }}_0/2`$ and we obtain the surprising result $`\stackrel{~}{\mathrm{\Delta }}_0=\stackrel{~}{B}_2`$, i.e., for an attractive delta interaction the vacuum and CP binding energies for zero CMM coincide for all coupling, a result apparently first obtained in Ref. . For a nonzero CMM the CP binding energies $`\mathrm{\Delta }_K`$ can be calculated from Eqs. (7) and ( 8). For $`\stackrel{~}{K}0`$ a minimum threshold value of $`B_2/E_F`$ is found to be required to bind a CP. Equations (7) and (8) can be solved numerically for the CP binding $`\mathrm{\Delta }_K`$ for any CMM. For small CMM only Eq. (7) is relevant; and this equation for small but non-zero $`\stackrel{~}{K}`$ and for $`\stackrel{~}{K}=0`$ can be subtracted one from the other. This gives the small-CMM expansion valid for any coupling $`B_2/E_F0`$, $`\epsilon _K`$ $``$ $`(\mathrm{\Delta }_0\mathrm{\Delta }_K)={\displaystyle \frac{2}{\pi }}\mathrm{}v_FK`$ (9) $`+`$ $`\left[1\left\{2\left({\displaystyle \frac{4}{\pi }}\right)^2\right\}{\displaystyle \frac{E_F}{B_2}}\right]{\displaystyle \frac{\mathrm{}^2K^2}{2(2m)}}+O(K^3),`$ (10) where a nonnegative CP excitation energy $`\epsilon _K`$ has been defined, and the Fermi velocity $`v_F`$ comes from $`E_F/k_F=\mathrm{}v_F/2`$. The leading term in (10) is linear in the CMM, followed by a quadratic term. The linear CP dispersion term should not be confused with that of the many-body (collective) excitation spectrum in weak coupling. Only CPs can undergo BEC while bosonic “excitations” (or modes or phonons) cannot since the former are fixed in number while the latter are not. Indeed, the particle-hole \[sometimes called the Anderson-Bogoliubov-Higgs (ABH)\] modes of excitation energy $`\mathrm{}v_FK/\sqrt{d}`$ in $`d`$ dimensions in the zero coupling limit are larger than the weak-coupling CP dispersion energies $`(2/\pi )\mathrm{}v_FK`$ and $`\frac{1}{2}\mathrm{}v_FK`$ (Ref. p. 33) in 2D and 3D , respectively, while in 1D they happen to coincide—in spite of the fact that CPs and ABH-like modes are physically distinct entities. The coefficient of the quadratic term in (10) changes sign at $`B_2/E_F=\mathrm{\Delta }_0/E_F0.379,`$ as one goes from weak ($`B_2=\mathrm{\Delta }_0E_F`$) to strong ($`B_2=\mathrm{\Delta }_0E_F`$) coupling. If $`v_F`$ (or $`E_F`$) $`0`$ explicitly (dilute limit) the first two terms of Eq. (10) reduce simply to $$\epsilon _K\frac{\mathrm{}^2K^2}{2(2m)}\text{}$$ (11) for any coupling. This is clearly just the familiar nonrelativistic kinetic energy in vacuum of the composite (so-called “local”) pair of mass $`2m`$ and CMM $`K`$. The same result (11) is also found to hold in 3D. Figure 1 shows exact numerical results for the zero-range potential ($`g_k=1`$) for different couplings of a CP excitation energy $`\epsilon {}_{K}{}^{}/\mathrm{\Delta }_0`$ as function of CMM $`K/k_F`$, both dimensionless. We note that the CPs break up whenever $`\mathrm{\Delta }_{K\text{ }}`$ turns from positive to negative, i.e., vanishes, or by Eq. (10) when $`\epsilon {}_{K}{}^{}/\mathrm{\Delta }_0=1`$. These points are marked in the figure by dots. In addition to the exact results obtained by solving Eqs. (7) and (8), we also exhibit the results for the linear approximation \[first term on the right-hand side of Eq. (10), dot-dashed lines, virtually coinciding with the exact curve for all $`B_2/E_F0.1`$\] as well as for the quadratic approximation (dashed parabolas) as given by Eq. (11) for stronger couplings. For weak enough coupling or large enough fermion density at any nonzero coupling the exact dispersion relation is virtually linear—in spite of the divergence of the isolated quadratic term in Eq. (10) as $`B_2/E_F0`$. As coupling is increased the quadratic dispersion relation (11) slowly begins to dominate. The crossover from a linear to a quadratic dispersion relation manifests itself by a change in curvature from concave down to concave up—these two regions being separated by an inflection point that moves down towards the origin as coupling is increased to infinity. Figure 2 shows the CP excitation energy $`\epsilon {}_{K}{}^{}/\mathrm{\Delta }_0`$ as a function of CMM $`K/k_F`$ calculated for the finite-range interaction form factor $`g_p=(1+p^2/p_0^2)^{1/2}`$ with $`p_0=k_F`$ for weak to moderate coupling. To compare, we also plot the zero-range result as well as the linear relation given by the first term on the right-hand side of Eq. (10). The finite-range curves are closer to the corresponding zero-range ones if labeled by $`\mathrm{\Delta }_0/E_F`$ instead of by $`B_2/E_F`$ , as was done with all four sets of curves. Figure 3 exhibits $`\epsilon {}_{K}{}^{}/\mathrm{\Delta }_0`$ as a function of $`K/k_F`$ for the finite-range interaction with $`p_0=k_F`$ for stronger couplings. In this case there is no special advantage in labeling the dispersion curves by $`\mathrm{\Delta }_0`$ so $`B_2`$ was used with results for $`B_2/E_F=`$ 3, 10 and 20 shown. In the zero-range case the curves gradually tend to the quadratic form as $`B_2`$ increases. For finite-range, $`p_0=k_F`$, the curves develop a maximum followed by a minimum with a point of inflection in between. The slope at the point of inflection is now negative. Although each curve tends to a quadratic form for large enough $`K/k_F`$, they are quite different from it for small $`K/k_F`$. These “looped” dispersion curves are reminiscent of the “roton” excitation spectrum in liquid $`^\text{4}`$He. To summarize, the single CP problem with non-zero CMM is tracked as it evolves in varying the interfermion short-range pair interaction from weak to strong or varying fermion density from high to low, respectively, for any fixed nonzero coupling. The CP excitation energy is exhibited as a function of its CMM. For weak coupling, the excitation energy is a linear dispersion relation in the CMM, and changes very gradually to a quadratic relation as coupling increases. For a zero-range pair interaction in the dispersion curve one typically has a point of inflection with a positive slope separating a region of concave-down curvature for small CMM from a region of concave-up curvature for large CMM. For finite-range pair interactions of sufficiently long range the slope at the point of inflection changes from positive to zero and eventually becomes negative. This leads to maxima and “roton-like” minima in the CP dispersion curves. These results will play a critical role in a model of superconductivity based on BE condensation of CPs as they will yield, even in 2D as in the cuprates , BEC transition temperatures $`T_c`$ that interpolate between nonzero values in weak coupling with a linear CP dispersion relation down to the expected $`T_c0`$ value in strong coupling with a quadratic relation. Appendix: The restriction that both particles lie above the Fermi sea in Eq. (6) can be written as $$\left(𝐤/k_F\pm 𝐊/2k_F\right)^21=\xi ^2\pm \xi \stackrel{~}{K}\mathrm{cos}\theta +\stackrel{~}{K}^2/410,$$ (12) where $`\xi k/k_F`$ and $`\stackrel{~}{K}K/k_F`$. The equality leads to two pairs of roots in $`\xi `$, say $`\xi _{1,2}=a\pm b`$ and $`\xi _{3,4}=a\pm b`$, where $`a(\stackrel{~}{K}/2)\mathrm{cos}\theta `$, $`b\sqrt{1(\stackrel{~}{K}^2/4)\mathrm{sin}^2\theta }`$, and $`\theta `$ the angle between $`𝐤`$ and $`𝐊`$. For $`\stackrel{~}{K}<2`$, $`b>a`$, one root of the two pairs is positive and the other negative. Thus, Eq. (12) can be satisfied provided that $`\xi `$ $`>\xi _1,\xi _2,\xi _3,\xi _4`$, or specifically, if $`\xi >\xi _0(\theta )a+b.`$ For $`\stackrel{~}{K}>2`$ and $`\theta >\theta _0\mathrm{arcsin}(2/\stackrel{~}{K})`$, $`b`$ becomes imaginary and Eq. (12) is satisfied for all $`\xi `$. Therefore, there is no restriction in the integration over $`\xi `$. However, for $`\stackrel{~}{K}>2`$ and $`\theta <\theta _0`$, $`b<a`$ the pair of roots $`\xi _{1,2}`$ are both negative while the pair $`\xi _{3,4}`$ are both positive (with $`\xi _3>\xi _4`$). Consequently, in both cases Eq. (12) is satisfied only if $`\xi `$ is in the interval $`[0,\xi _0^{}(\theta )ab],`$ and in the interval $`[\xi _0(\theta ),\mathrm{}]`$ , respectively. Using these restrictions on the $`\xi `$ integration in Eq. ( 6) one eventually arrives at Eqs. (7) and (8). Acknowledgments: M.deLl. thanks S. Fujita for discussions, D.M. Eagles for reading the manuscript, and V.V. Tolmachev for extensive correspondence. Partial support from UNAM-DGAPA-PAPIIT (México) # IN102198, CONACyT (México) # 27828 E, DGES (Spain) # PB95-0492 and FAPESP (Brazil) is gratefully acknowledged. Fig. 1. Dimensionless CP excitation energy $`\epsilon _K/\mathrm{\Delta }_0`$ vs $`K/k_{F\text{ }}`$, calculated from Eqs. (7) and (8) for different couplings $`B_2/E_F`$, full curves. The dot-dashed line is the linear approximation (virtually coincident with the exact curve for $`B_2/E_F`$ $``$ $`0.1`$) while the dashed curve is the quadratic term of Eq. (11). Dots denote values of CMM wavenumber where the CP breaks up, i.e., where $`\mathrm{\Delta }_K`$ $`0`$. Fig. 2. Same as Fig. 1 but for couplings expressed as $`\mathrm{\Delta }_0/E_F`$. The dot-dashed line is the linear approximation; the dashed curve is the result for the finite-range interaction $`p_0=k_F`$, and the full curve is the zero-range result. For the finite-range potential $`\mathrm{\Delta }_0/E_F=0.1,`$ $`0.5,`$ $`1.0`$ and $`2.0`$ correspond to $`B_2/E_F=0.469,`$ $`1.4,`$ $`2.45`$ and $`4`$, respectively. Dots and squares mark values of CMM wavenumber where the CP breaks up. Fig. 3. Same as Fig. 1 but including also finite range at stronger couplings $`B_2/E_F`$. The full curve is the exact zero-range result; the short-dashed one the quadratic approximation; the long-dashed one the exact finite-range result with $`p_0=k_F`$. Each set of three curves is labeled by different values of $`B_2/E_F`$.
warning/0004/gr-qc0004063.html
ar5iv
text
# Quantum lightcone fluctuations in theories with extra dimensions ## I Introduction One of the most challenging problems in modern physics is the unification of the gravitational interaction with other known interactions in nature. Many attempts, from early Kaluza-Klein theory to present supergravity and superstring theories, all involve going to higher dimensions and postulating the existence of extra spatial dimensions. The modern view regarding these postulated extra dimensions is that they are a physical reality, rather than merely a technical intermediate device for obtaining rather complicated four-dimensional theories from simpler Lagrangians in higher dimensions. If these extra dimensions really exist, one must explain why they are not seen. The usual answer is that they curl into an extremely small compactified manifold, possibly as small as the Planck length scale, $`l_{pl}=1.6\times 10^{33}`$ cm. Therefore low-energy physics should be insensitive to them until distances of the compactification scale are being probed. In general, one has the possibility of observing the presence of the extra dimensions in a scattering experiment in which energies greater than that associated with the compactification scale are achieved. Various upper bounds have been put on the size of possible extra dimensions . For example, an upper bound of $``$ 1 Tev was given in orbifold compactifications of superstrings . However, if only gravity propagates in the extra dimensions, the upper bound can be much larger. A recent proposal is that the fundamental scale of quantum gravity can be as low as few Tev and the observed weakness of gravity is the result of large extra dimensions in which only gravity can propagate . This scenario could be realized in the context of several string models in which one has a set of three-branes (3+1 dimensional spacetime) in the entire spacetime with extra dimensions. The Standard Model particles are confined to one of the branes, while gravitons propagate freely in the entire bulk. The size of extra dimensions could then be as large as 1 mm in this type of model. Extra dimensions of sufficiently large size may manifest themselves in particle colliders and in the possible deviation from Newton’s law at short distances, and they may also have implications in gauge unification and cosmology. However, a question arises naturally as to whether there are any lower bounds on the sizes of extra dimensions. It is the common belief that the existence of extra dimensions has no effect on low-energy physics as long as they are extremely small. Recently, we have argued that this is not the case, because of lightcone fluctuations arising from the quantum gravitational vacuum fluctuations due to compactification of spatial dimensions . An explicit calculation was carried out in the five-dimensional prototypical Kaluza-Klein model which showed that the periodic compactification of the extra spatial dimensions gives rise to stochastic fluctuations in the apparent speed of light which grow as the compactification scale decreases and are in principle observable. Basically, the smaller the size of the compactified dimensions , the larger are the fluctuations that result. This is closely related to the Casimir effect, the vacuum energy occurring whenever boundary conditions are imposed on a quantum field. The gravitational Casimir energy in the five-dimensional case with one compactified spatial dimension was studied in , where a nonzero energy density was found, which tends to make the extra dimension contract. This raises the question of stability of the extra dimensions. It is possible, however, that the Casimir energy arising from the quantum gravitational field and other matter fields may be made to cancel each other , thus stabilizing the extra dimensions. Quantum lightcone fluctuations due to the compactification of spatial dimensions , although similar in nature to the Casimir effect, come solely from gravitons. Hence, no similar cancellation is to be expected. In an earlier work , we examined lightcone fluctuation in a five-dimensional model with periodic compactification. We found that there seem to be observable effects which essentially rule out this model. That is, we derived a lower bound on the compactification scale which is larger than upper bounds derived from other considerations. The purpose of the present paper is threefold: to further develop the basic formalism, to provide more details of the five-dimensional calculation, and to extend the analysis to some higher dimensional models. In Section II, we give a brief review of the formalism, discuss the observability of lightcone fluctuations, and derive the graviton two-point functions in the transverse tracefree gauge for spacetimes with an arbitrary number of dimensions (detailed calculations will given in the Appendix). In Section III, we examine light cone fluctuations in spacetimes with extra dimensions periodically compactified into a torus. The five-dimensional prototypical Kaluza-Klein model, of which some results have already been reported, will be studied in great detail. Higher dimensions up through 11 will also be discussed. Section IV deals with light cone fluctuations in the brane-world scenario, as motivated by a recent proposal of extra dimensions of macroscopic size. We will summarize and conclude in Section V. ## II Observability of light cone fluctuations and the graviton two-point function in the TT gauge To begin, let us examine a $`d=4+n`$ dimensional spacetime with $`n`$ extra dimensions. Consider a flat background spacetime with a linearized perturbation $`h_{\mu \nu }`$ propagating upon it , so the spacetime metric may be written as $`ds^2=(\eta _{\mu \nu }+h_{\mu \nu })dx^\mu dx^\nu =dt^2d𝐱^2+h_{\mu \nu }dx^\mu dx^\nu ,`$ where the indices $`\mu ,\nu `$ run through $`0,1,2,3,\mathrm{},3+n`$. Let $`\sigma (x,x^{})`$ be one half of the squared geodesic distance between a pair of spacetime points $`x`$ and $`x^{}`$, and $`\sigma _0(x,x^{})`$ be the corresponding quantity in the flat background . In the presence of a linearized metric perturbation, $`h_{\mu \nu }`$, we may expand $`\sigma =\sigma _0+\sigma _1+O(h_{\mu \nu }^2).`$ Here $`\sigma _1`$ is first order in $`h_{\mu \nu }`$. If we quantize $`h_{\mu \nu }`$, then quantum gravitational vacuum fluctuations will lead to fluctuations in the geodesic separation, and therefore induce lightcone fluctuations. In particular, we have $`\sigma _1^20`$, since $`\sigma _1`$ becomes a quantum operator when the metric perturbations are quantized. The quantum lightcone fluctuations give rise to stochastic fluctuations in the speed of light, which may produce an observable time delay or advance $`\mathrm{\Delta }t`$ in the arrival times of pulses. ### A Observability of light cone fluctuation Here, we shall discuss how light cone fluctuations characterized by $`\sigma _1^2`$ are related to physical observable quantities. For this purpose, let us consider the propagation of light pulses between a source and a detector separated by a distance $`r`$ on a flat background with quantized linear perturbations. For a pulse which is delayed or advanced by time $`\mathrm{\Delta }t`$, which is much less than $`r`$, one finds $$\sigma =\sigma _0+\sigma _1+\mathrm{}.=\frac{1}{2}[(r+\mathrm{\Delta }t)^2r^2]r\mathrm{\Delta }t.$$ (1) Square the above equation and take the average over a given quantum state of gravitons $`|\varphi `$ (e.g. the vacuum states associated with compactification of spatial dimensions ), $$\mathrm{\Delta }t_\varphi ^2=\frac{\varphi |\sigma _1^2|\varphi }{r^2}.$$ (2) This result is, however, divergent due to the formal divergence of $`\varphi |\sigma _1^2|\varphi `$. One can define an observable $`\mathrm{\Delta }t_{obs}`$ by subtracting from Eq. (2) the corresponding quantity, $`\mathrm{\Delta }t_0^2`$, for the vacuum state as follows $$\mathrm{\Delta }t_{obs}^2=|\mathrm{\Delta }t_\varphi ^2\mathrm{\Delta }t_0^2|=\frac{|\varphi |\sigma _1^2|\varphi 0|\sigma _1^2|0|}{r^2}\frac{|\sigma _1^2_R|}{r}.$$ (3) Here we take the absolute value of the difference between $`\mathrm{\Delta }t_\varphi ^2`$ and $`\mathrm{\Delta }t_0^2`$, because the observable quantity $`\mathrm{\Delta }t_{obs}^2`$ has to be a positive real number. Note that we can also get this result from the gravitational quantum average of the retarded Green’s function $`G_{ret}(x,x^{})`$ when $`\sigma _1^2_R>0`$. Therefore, the root-mean-squared deviation from the classical propagation time is given by $$\mathrm{\Delta }t_{obs}=\frac{\sqrt{|\sigma _1^2_R|}}{r}.$$ (4) At this point, a question may arise as to whether the formal procedure of taking the absolute value in deriving the relation between $`\mathrm{\Delta }t`$ and $`\sigma _1^2_R`$, Eq. (4), is a reasonable one, or whether a meaningful relation between $`\mathrm{\Delta }t`$ and $`\sigma _1^2_R`$ can be established only when $`\sigma _1^2_R>0`$. We shall argue that a result essentially the same as Eq. (4) can be obtained by the following analysis which avoids the sign problem. Instead of squaring the Eq. (1), we take the fourth power of both sides and average over a quantum vacuum state to yield $$\mathrm{\Delta }t_\varphi ^4=\frac{\varphi |\sigma _1^4|\varphi }{r^4}.$$ (5) We can regularize $`\varphi |\sigma _1^4|\varphi `$ by normal ordering and define $$\varphi |\sigma _1^4|\varphi _R=\varphi |:\sigma _1^4:|\varphi .$$ (6) For a free field $`\psi `$ and a quantum vacuum state one can show by use of Wick’s theorem that $$:\psi ^4:=3:\psi ^2:^2=3\psi ^2_R^2.$$ (7) Therefore one finds that $$\mathrm{\Delta }t_{obs}=\frac{3^{1/4}\sqrt{|\sigma _1^2_R|}}{r}\frac{\sqrt{|\sigma _1^2_R|}}{r}.$$ (8) This is essentially the same as Eq. (4), apart from a dimensionless factor of order unity. We should note that there may be other ways to define the quartic operator, $`\sigma _1^4`$. One possibility is to let $`\sigma _1^4=(:\sigma _1^2:)^2`$, provided that the integrals involved can be defined. Both of these definitions were discussed in Ref. . There it was found in some model cases that the two definitions yield the same result, apart from numerical factors of order unity, which are not important for the present purposes. Note that $`\mathrm{\Delta }t`$ is the ensemble averaged deviation, not necessarily the expected variation in flight time, $`\delta t`$, of two pulses emitted close together in time. The latter is given by $`\mathrm{\Delta }t`$ only when the correlation time between successive pulses is less than the time separation of the pulses. This can be understood physically as due to the fact that the gravitational field may not fluctuate significantly in the interval between the two pulses. This point is discussed in detail in Ref. . These stochastic fluctuations in the apparent velocity of light arising from quantum gravitational fluctuations are in principle observable, since they may lead to a spread in the arrival times of pulses from distant astrophysical sources, or the broadening of the spectral lines. Lightcone fluctuations and their possible astrophysical observability have been recently discussed in a somewhat different framework in Refs. . ### B An Alternative Derivation of $`\mathrm{\Delta }t`$ In order to find $`\mathrm{\Delta }t`$ in a particular situation, we need to calculate the quantum expectation value $`\sigma _1^2_R`$ in any chosen quantum state $`|\psi `$, which can be shown to be given by Although the derivations there were given in 3+1 dimensions, the generalization to arbitrary dimensions is straightforward. $$\sigma _1^2_R=\frac{1}{8}(\mathrm{\Delta }r)^2_{r_0}^{r_1}𝑑r_{r_0}^{r_1}𝑑r^{}n^\mu n^\nu n^\rho n^\sigma G_{\mu \nu \rho \sigma }^R(x,x^{}).$$ (9) Here $`dr=|d𝐱|`$, $`\mathrm{\Delta }r=r_1r_0`$ and $`n^\mu =dx^\mu /dr`$. The integration is taken along the null geodesic connecting two points $`x`$ and $`x^{}`$, and $$G_{\mu \nu \rho \sigma }^R(x,x^{})=\psi |h_{\mu \nu }(x)h_{\rho \sigma }(x^{})+h_{\mu \nu }(x^{})h_{\rho \sigma }(x)|\psi $$ (10) is the graviton Hadamard function, understood to be suitably renormalized. The gauge invariance of $`\mathrm{\Delta }t`$, as given by Eq. (4), has been analyzed recently . In this subsection, we wish to rederive $`\mathrm{\Delta }t`$ using the geodesic deviation equation. This derivation allows us to see the gauge invariance more clearly, and to discuss the issue of Lorentz invariance of lightcone fluctuations. Let us consider a pair of timelike geodesics with tangent vector $`u^\mu `$, and $`n^\mu `$ as a unit spacelike vector pointing from one geodesic to the other (See Fig. 1). The geodesic deviation equation is given by $$\frac{D^2n^\mu }{d\tau ^2}=R_{\alpha \nu \beta }^\mu u^\alpha n^\nu u^\beta ,$$ (11) where $`R_{\alpha \nu \beta }^\mu `$ is the Riemann tensor. The relative acceleration per unit proper length of particles on the neighboring geodesics is $$\alpha n_\mu \frac{D^2n^\mu }{d\tau ^2}=R_{\mu \alpha \nu \beta }n^\mu u^\alpha n^\nu u^\beta .$$ (12) Thus if $`ds`$ is the spatial distance between the two particles, then $`\alpha ds`$ is their relative acceleration. It follows that the relative change in displacement of the two particles after a proper time $`T`$ is $$ds_0^T𝑑\tau _0^\tau 𝑑\tau ^{}\alpha (\tau ^{},0),$$ (13) Now consider the case of two observers (particles) separated by a finite initial distance $`s_0`$ as illustrated in Fig. (2). We can find the relative change in displacement of these two observers by integrating on $`s`$: $$\mathrm{\Delta }s=_0^{s_0}𝑑s_0^T𝑑\tau _0^\tau 𝑑\tau ^{}\alpha (\tau ^{},0).$$ (14) This is the relative displacement measured at the same moment of proper time for both observers. Let us now consider a light signal sent from one observer to the other. If $`\alpha =0`$, the distance traveled by the light ray is $`s_0`$. When $`\alpha 0`$, this distance becomes $`s_0+\mathrm{\Delta }s`$, where now $$\mathrm{\Delta }s=_0^{s_0}𝑑s\underset{}{_0^s𝑑\tau _0^\tau 𝑑\tau ^{}\alpha (\tau ^{},s)}.$$ (15) Here the under-braced integral is the displacement per unit $`s`$ of a pair of observers at a distance $`s`$ from the source. The domain of the final two integrations is illustrated in Fig. (3). If gravity is quantized, the Riemann tensor will fluctuate around an average value of zero due to quantum gravitational vacuum fluctuations. This leads to $`\alpha =0`$, and hence $`\mathrm{\Delta }s=0`$. Notice here that $`\alpha `$ becomes a quantum operator when metric perturbations are quantized. However, in general, $`(\mathrm{\Delta }s)^20`$, and we have $$(\mathrm{\Delta }s)^2=_0^{s_0}𝑑s_1_0^{s_0}𝑑s_2_0^{s_1}𝑑\tau _1_0^{\tau _1}𝑑\tau _1^{}_0^{s_2}𝑑\tau _2_0^{\tau _2}𝑑\tau _2^{}\alpha (\tau _1^{},s_1)\alpha (\tau _2^{},s_2).$$ (16) Thus the root-mean-squared fluctuation in the flight path is $`\sqrt{(\mathrm{\Delta }s)^2}`$, which can also be understood as a fluctuation in the speed of light. It entails an intrinsic quantum uncertainty in the measurement of distance. Therefore, spacetime becomes fuzzy at a scale characterized by $`\sqrt{(\mathrm{\Delta }s)^2}`$. The integrand in Eq. (16) is obviously invariant under any coordinate transformation while the integral is gauge invariant within the linear approximation. We now wish to show that this gauge-invariant quantity is the same as Eq. (4) when calculated in the transverse-tracefree (TT) gauge. Choose a coordinate system where the source and the detector are both at rest, and suppose that the light ray propagates in the $`x`$-direction, then we have $$u^\mu =(1,0,0,0),$$ (17) $$n^\mu =(0,1,0,0),$$ (18) and $$\alpha =R_{xtxt}=\frac{1}{2}h_{xx,tt}.$$ (19) Substitution of the above results into Eq. (16) leads to $`(\mathrm{\Delta }s)^2`$ $`=`$ $`{\displaystyle ^r}𝑑x_1{\displaystyle _0^r}𝑑x_2{\displaystyle _0^{x_1}}𝑑t_1{\displaystyle _0^{t_1}}𝑑t_1^{}{\displaystyle _0^{x_2}}𝑑t_2{\displaystyle _0^{t_2}}𝑑t_2^{}\alpha (t_1^{},s_1)\alpha (t_2^{},s_2)`$ (20) $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle _0^r}𝑑x_1{\displaystyle _0^r}𝑑x_2h_{xx}(x_1,x_1)h_{xx}(x_2,x_2)={\displaystyle \frac{1}{r}}\sigma _1^2,`$ (21) where we have set $`s_0=r`$ and used the fact that along the light ray $`x=t`$. Thus, one has $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2}}{r}=\sqrt{(\mathrm{\Delta }s)^2}$$ (22) which also demonstrates the gauge-invariance of $`\mathrm{\Delta }t`$. Now we wish to discuss the rather subtle issue of the relation of lightcone fluctuations to Lorentz symmetry. It is sometimes argued that lightcone fluctuations are incompatible with Lorentz invariance. The most dramatic illustration of this arises when a time advance occurs, that is, when a pulse propagates outside of the classical lightcone. In a Lorentz invariant theory, there will exist a frame of reference in which the causal order of emission and detection is inverted, so the pulse is seen to be detected before it was emitted. Thus the lightcone fluctuation phenomenon, if it is to exist at all, seems to be incompatible with strict Lorentz invariance. Our view of the situation is the following: lightcone fluctuations respect Lorentz symmetry on the average, but not in individual measurements. The symmetry on the average insures that the mean lightcone be that of classical Minkowski spacetime. The average metric is that of Minkowski spacetime provided that $`h_{\mu \nu }=0.`$ However, a particular pulse effectively measures a spacetime geometry which is not Minkowskian and not Lorentz invariant. A simple model may help to illustrate this point. Consider a quantum geometry consisting of an ensemble of classical Schwarzschild spacetimes, but with both positive and negative values for the mass parameter $`M`$. (The fact that the $`M<0`$ Schwarzschild spacetime has a naked singularity at $`r=0`$ need not concern us. For the purpose of this model, we can confine our discussion to a region where $`r|M|.`$) Suppose that this ensemble has $`M=0`$, but $`M^20`$. It is well known that light propagation in a $`M>0`$ Schwarzschild spacetime can exhibit a time delay relative to what would be expected in flat spacetime. This is the basis for the time delay tests of general relativity using radar signals sent near the limb of the sun. In the present model, however, the time difference is equally likely to be a time advance rather than a time delay. A measurement of the time difference amounts to a measurement of $`M`$. This model is Lorentz invariant on the average because $`M=0`$ and the average spacetime is Minkowskian. However, a specific measurement selects a particular member of the ensemble, which is generally not Lorentz invariant. In addition to the fact that the mean metric is Minkowskian, there is another sense in which lightcone fluctuations due to compactification exhibit average Lorentz invariance. Note that $`\mathrm{\Delta }s`$, and hence $`\mathrm{\Delta }t`$, depends on the Riemann tensor correlation function $`R_{xtxt}(x_1)R_{xtxt}(x_2)`$, which is invariant under Lorentz boosts along the $`x`$-axis. Thus if we were to repeat the above calculations of $`\mathrm{\Delta }s`$ in a second frame moving with respect to the first, the result will be the same. In both cases one is assuming that the detector is at rest relative to the source. This is a reflection of the Lorentz invariance of the spectrum of fluctuations, which is exhibited by the compactified flat spacetimes studied in this paper, but not by the Schwarzschild spacetime with a fluctuating mass. ### C Graviton two-point functions We shall use a quantization of the linearized gravitational perturbations $`h_{\mu \nu }`$ which retains only physical degrees of freedom. That is, we are going to work in the TT gauge in which the gravitational perturbations have only spatial components $`h_{ij}`$, satisfying the transverse, $`^ih_{ij}=0`$, and tracefree, $`h_i^i=0`$ conditions. Here $`i,j`$ run from 1 to $`3+n=d1`$. These $`2d`$ conditions remove all of the gauge degrees of freedom and leave $`\frac{1}{2}(d^23d)`$ physical degrees of freedom. We have $$h_{ij}=\underset{𝐤,\lambda }{}[a_{𝐤,\lambda }e_{ij}(𝐤,\lambda )f_𝐤+H.c.].$$ (23) Here H.c. denotes the Hermitian conjugate, $`\lambda `$ labels the $`\frac{1}{2}(d^23d)`$ independent polarization states, $`f_𝐤`$ is the mode function, and the $`e_{\mu \nu }(𝐤,\lambda )`$ are polarization tensors. (Units in which $`32\pi G_d=1`$, where $`G_d`$ is Newton’s constant in d dimensions and in which $`\mathrm{}=c=1`$ will be used in this paper.) Let us now calculate the Hadamard function, $`G_{\mu \nu \rho \sigma }(x,x^{})`$, for gravitons in the Minkowski vacuum state in the transverse tracefree gauge. It follows that $$G_{ijkl}(x,x^{})=\frac{2Re}{(2\pi )^{d1}}\frac{d^{d1}𝐤}{2\omega }\underset{\lambda }{}e_{ij}(𝐤,\lambda )e_{kl}(𝐤,\lambda )e^{i𝐤(𝐱𝐱^{})}e^{i\omega (tt^{})}.$$ (24) The summation of polarization tensors in the transverse tracefree gauge is (See the Appendix in Ref. .<sup>§</sup><sup>§</sup>§ The tensorial argument given there applies in any number of dimensions ) $`{\displaystyle \underset{\lambda }{}}`$ $`e_{ij}(𝐤,\lambda )e_{kl}(𝐤,\lambda )=\delta _{ik}\delta _{jl}+\delta _{il}\delta _{jk}\delta _{ij}\delta _{kl}+\widehat{k}_i\widehat{k}_j\widehat{k}_k\widehat{k}_l`$ (26) $`+\widehat{k}_i\widehat{k}_j\delta _{kl}+\widehat{k}_k\widehat{k}_l\delta _{ij}\widehat{k}_i\widehat{k}_l\delta _{jk}\widehat{k}_i\widehat{k}_k\delta _{jl}\widehat{k}_j\widehat{k}_l\delta _{ik}\widehat{k}_j\widehat{k}_k\delta _{il},`$ where $`\widehat{k}_i=\frac{k_i}{k}.`$ We find that $`G_{ijkl}`$ $`=`$ $`2F_{ij}\delta _{kl}+2F_{kl}\delta _{ij}2F_{ik}\delta _{jl}2F_{il}\delta _{jk}2F_{jl}\delta _{ik}2F_{jk}\delta _{il}`$ (28) $`+2H_{ijkl}+2D(x,x^{})(\delta _{ik}\delta _{jl}+\delta _{il}\delta _{jk}\delta _{ij}\delta _{kl}).`$ Here $`D(x,x^{})`$, $`F_{ij}(x,x^{})`$ and $`H_{ijkl}(x,x^{})`$ are functions which are defined as follows: $$D^n(x,x^{})=\frac{Re}{(2\pi )^{3+n}}\frac{d^{3+n}𝐤}{2\omega }e^{i𝐤(𝐱𝐱^{})}e^{i\omega (tt^{})},$$ (29) $$F_{ij}^n(x,x^{})=\frac{Re}{(2\pi )^{3+n}}_i_j^{}\frac{d^{3+n}𝐤}{2\omega ^3}e^{i𝐤(𝐱𝐱^{})}e^{i\omega (tt^{})},$$ (30) and $$H_{ijkl}^n(x,x^{})=\frac{Re}{(2\pi )^{3+n}}_i_j^{}_k_l^{}\frac{d^{3+n}𝐤}{2\omega ^5}e^{i𝐤(𝐱𝐱^{})}e^{i\omega (tt^{})}.$$ (31) These functions are calculated in the Appendix. Let $$R=|𝐱𝐱^{}|,\mathrm{\Delta }t=tt^{}.$$ (32) For $`n=2m+1`$, an odd number of extra dimensions, the results are $$D^{2m+1}=\{\begin{array}{cc}\frac{(2m+1)!!}{2(2\pi )^{m+2}}\frac{1}{(R^2\mathrm{\Delta }t^2)^{m+\frac{3}{2}}},\hfill & \text{ for}R^2>\mathrm{\Delta }t^2,\hfill \\ 0\hfill & \text{ for}R^2<\mathrm{\Delta }t^2,\hfill \end{array}$$ (33) $$F_{ij}^{2m+1}=\frac{1}{2(2\pi )^{m+2}}_i_j^{}\left[\frac{(2m)!!}{R^{2m}}S(0)\underset{k=0}{\overset{m}{}}\frac{(2k+1)!!}{(2k)!!(2k+1)(2k1)}\frac{R^{2k}}{(R^2\mathrm{\Delta }t^2)^k}\right],$$ (34) and $`H_{ijkl}^{2m+1}`$ $`=`$ $`{\displaystyle \frac{1}{2(2\pi )^{m+2}}}_i_j^{}_k_l^{}\{{\displaystyle \frac{(2m)!!}{R^{2m2}}}[{\displaystyle \frac{1}{2}}Q(1)`$ (36) $`+{\displaystyle \underset{j=0}{\overset{m2}{}}}{\displaystyle \frac{(mj1)(2j+1)!!}{4m(j+1)(2j)!!(2j+1)(2j1)}}{\displaystyle \frac{R^{2j}}{(R^2\mathrm{\Delta }t^2)^j}}S(0)].`$ In particular, for $`n=1`$ ($`d=5`$), we have $$H_{ijkl}^1=0$$ (38) Here $$S(0)=\{\begin{array}{cc}\frac{\sqrt{R^2\mathrm{\Delta }t^2}}{R^2}\hfill & \text{ for}R^2>\mathrm{\Delta }t^2,\hfill \\ 0\hfill & \text{ for}R^2<\mathrm{\Delta }t^2,\hfill \end{array}$$ (39) and $$Q(1)=\left(\frac{1}{3}\frac{\mathrm{\Delta }t^2}{3R^2}\right)S(0).$$ (40) For $`n=2m`$, an even number of extra dimensions, we have $$D^{2m}=\frac{2^mm!}{(2\pi )^{m+2}}\frac{1}{(R^2\mathrm{\Delta }t^2)^{m+1}},$$ (41) $`F_{ij}^{2m}`$ $`=`$ $`{\displaystyle \frac{1}{(2\pi )^{m+2}}}_i_j^{}\{{\displaystyle \frac{(2m1)!!}{R^{2m}}}[1{\displaystyle \frac{\mathrm{\Delta }t}{4R}}\mathrm{ln}\left({\displaystyle \frac{R+\mathrm{\Delta }t}{R\mathrm{\Delta }t}}\right)^2`$ (43) $`{\displaystyle \frac{1}{R^2}}{\displaystyle \underset{k=2}{\overset{m}{}}}{\displaystyle \frac{2^{k2}\mathrm{\Gamma }(k1)}{(2k1)!!}}{\displaystyle \frac{R^{2k}}{(R^2\mathrm{\Delta }t^2)^{k1}}}]\}m1,`$ and $`H_{ijkl}^{2m}`$ $`=`$ $`{\displaystyle \frac{1}{(2\pi )^{m+2}}}_i_j^{}_k_l^{}\{{\displaystyle \frac{(2m1)!!}{R^{2m4}}}[{\displaystyle \frac{\mathrm{\Delta }t}{24R^3}}({\displaystyle \frac{\mathrm{\Delta }t^2}{R^2}}1)\mathrm{ln}\left({\displaystyle \frac{R+\mathrm{\Delta }t}{R\mathrm{\Delta }t}}\right)^2{\displaystyle \frac{1}{18R^2}}`$ (47) $`{\displaystyle \frac{\mathrm{\Delta }t^2}{6R^4}}{\displaystyle \underset{k=3}{\overset{m}{}}}{\displaystyle \frac{1}{(2k1)(2k3)}}({\displaystyle \frac{1}{R^2}}{\displaystyle \frac{\mathrm{\Delta }t}{4R^3}}\mathrm{ln}\left({\displaystyle \frac{R+\mathrm{\Delta }t}{R\mathrm{\Delta }t}}\right)^2)`$ $`+{\displaystyle \frac{1}{R^4}}{\displaystyle \underset{j=2}{\overset{m2}{}}}{\displaystyle \frac{(mj1)2^{j2}\mathrm{\Gamma }(j1)}{(2m1)(2j+1)!!}}{\displaystyle \frac{R^{2j}}{(R^2\mathrm{\Delta }t^2)^{j1}}}]\}m2.`$ For the case $`n=2`$ ($`d=6`$), we have for $`H`$ $$H_{ijkl}^2=\frac{1}{(2\pi )^3}_i_j^{}_k_l^{}[\frac{1}{6}\mathrm{ln}(R^2\mathrm{\Delta }t^2)\frac{\mathrm{\Delta }t^2}{6R^2}+\frac{\mathrm{\Delta }t}{8R}(\frac{\mathrm{\Delta }t^2}{3R^2}1\left)\mathrm{ln}\left(\frac{R+\mathrm{\Delta }t}{R\mathrm{\Delta }t}\right)^2\right].$$ (49) The case of $`n=0`$ ($`d=4`$) was given in Ref. . ## III Periodic compactification scenario Let us now suppose that the extra $`n`$ dimensions $`z_1,\mathrm{},z_n`$ are compactified with periodicity lengths $`L_1,\mathrm{},L_n`$, namely spatial points $`z_i`$ and $`z_i+L_i`$ are identified. For simplicity, we shall assume in this paper that $`L_1=\mathrm{}=L_n=L`$. The effect of imposition of the periodic boundary conditions on the extra dimensions is to restrict the field modes to a discrete set $$f_𝐤=(2\omega (2\pi )^3L^n)^{\frac{1}{2}}e^{i(𝐤𝐱\omega t)},$$ (50) with $$k_i=\frac{2\pi m_i}{L},i=1,\mathrm{},n,m_i=0,\pm 1,\pm 2,\pm 3,\mathrm{}.$$ (51) Let us denote the associated vacuum state by $`|0_L`$. In order to calculate the gravitational vacuum fluctuations due to compactification of extra dimensions, we need the renormalized graviton Hadamard function with respect to the vacuum state $`|0_L`$, $`G_{\mu \nu \rho \sigma }^R(x,x^{})`$, which is given by a multiple image sum of the corresponding Hadamard function for the Minkowski By Minkowski we mean flat spacetime with all dimensions uncompactified vacuum, $`G_{\mu \nu \rho \sigma }`$ : $$G_{\mu \nu \rho \sigma }^R(t,z_i,t^{},z_i^{})=\underset{i=1}{\overset{n}{}}\underset{m_i=\mathrm{}}{\overset{+\mathrm{}}{}}^{}G_{\mu \nu \rho \sigma }(t,z_i,t^{},z_i^{}+m_iL).$$ (52) Here the prime on the summation indicates that the $`m_i=0`$ term is excluded and the notation $$(t,\stackrel{}{x},z_1,..,z_n)(t,z_i)$$ (53) has been adopted. We are mainly concerned about how lightcone fluctuations arise in the usual uncompactified space as a result of compactification of extra dimensions. So we shall examine the case of a light ray propagating in one of the uncompactified dimensions. Take the direction to be along the $`x`$-axis in our four dimensional world, then the relevant graviton two-point function is $`G_{xxxx}`$, which can be expressed as $`G_{xxxx}(t,\stackrel{}{x},z_i,t^{},\stackrel{}{x}^{},z_i^{})`$ $`=`$ $`2[D(t,\stackrel{}{x},z_i,t^{},\stackrel{}{x}^{},z_i^{})2F_{xx}(t,\stackrel{}{x},z_i,t^{},\stackrel{}{x}^{},z_i^{})`$ (55) $`+H_{xxxx}(t,\stackrel{}{x},z_i,t^{},\stackrel{}{x}^{},z_i^{})].`$ Assuming that the propagation goes from point $`(a,0,\mathrm{},0)`$ to point $`(b,0,\mathrm{},0)`$, we have $`\sigma _1^2`$ $`=`$ $`{\displaystyle \frac{1}{8}}(ba)^2{\displaystyle _a^b}dx{\displaystyle _a^b}dx^{}G_{xxxx}^R(t,x,\mathrm{𝟎},t^{},x^{},\mathrm{𝟎},),`$ (56) $`=`$ $`{\displaystyle \frac{1}{8}}(ba)^2{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \underset{m_i=\mathrm{}}{\overset{+\mathrm{}}{}}^{}}G_{xxxx}(t,x,\mathrm{𝟎},t^{},x^{},0,0,m_1L,\mathrm{},m_iL).`$ (57) With these results, we can in principle calculate lightcone fluctuations in spacetimes with arbitrary number of flat extra dimensions. In what follows, we first examine some particular cases, then make some observations for the general case. ### A The five dimensional Kaluza-Klein model One of the most intriguing and elegant ways of unifying gauge field theories with gravitation is the higher-dimensional generalizations of Kaluza-Klein theory. The original suggestion of Kaluza and Klein was that electromagnetism and general relativity could be unified by starting with a five-dimensional version of the latter and then somehow arranging for the fifth dimension to become unobservable. This idea was further generalized to higher dimensions in attempts to unify non-Abelian gauge fields with gravitation and has been extensively studied in recent years in the context of supergravity and superstring theories. In the course of investigation of new features arising from the introduction of extra dimensions, the five-dimensional Kaluza-Klein theory has always been taken as a prototypical model to carry out explicit calculations to obtain a basic understanding. This is also our strategy here. In this subsection, we will derive the results which were summarized in Ref. . #### 1 Calculation of $`\mathrm{\Delta }t`$ To begin, let us examine the influence of the compactification of the fifth (extra) dimension on the light propagation in our four dimensional world, by considering a light ray traveling along the $`x`$-direction from point $`a`$ to point $`b`$, which is perpendicular to the direction of compactification. Define $$\rho =xx^{},ba=r$$ (59) and note the fact that the integration in Eq. (LABEL:eq:sigma1) is to be carried out along the classical null geodesic on which $`tt^{}=\rho `$. Then we obtain, after performing the differentiation in both $`D`$ and $`F_{xx}`$, $`G_{xxxx}(t,x,0,0,0,t^{},x^{},0,0,mL^{})`$ (60) $`={\displaystyle \frac{1}{4\pi ^2}}[{\displaystyle \frac{\rho ^6}{(\rho ^2+m^2L^2)^3|mL|^3}}{\displaystyle \frac{7\rho ^4}{(\rho ^2+m^2L^2)^3|mL|}}`$ (61) $`+{\displaystyle \frac{9\rho ^2|mL|}{(\rho ^2+m^2L^2)^3}}{\displaystyle \frac{|mL|^3}{(\rho ^2+mL^2)^3}}].`$ (62) Thus, we have $`\sigma _1^2`$ $`=`$ $`{\displaystyle \frac{1}{8}}r^2{\displaystyle _a^b}dx{\displaystyle _a^b}dx^{}G_{xxxx}^R(t,x,\mathrm{𝟎},t^{},x^{},\mathrm{𝟎},),`$ (63) $`=`$ $`{\displaystyle \frac{1}{8}}r^2{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}{\displaystyle \underset{m=\mathrm{}}{\overset{+\mathrm{}}{}}^{}}G_{xxxx}(t,x,\mathrm{𝟎},t^{},x^{},0,0,mL)`$ (64) $`=`$ $`{\displaystyle \frac{r^2}{32\pi ^2L}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\left[{\displaystyle \frac{8}{m}}\mathrm{ln}\left(1+{\displaystyle \frac{\gamma ^2}{m^2}}\right){\displaystyle \frac{2\gamma ^2}{m^3}}{\displaystyle \frac{8\gamma ^2}{(\gamma ^2+m^2)m}}\right],`$ (65) where we have introduced a dimensionless parameter $`\gamma =r/L`$. We are interested here in the case in which $`\gamma 1`$. It then follows that the summation is dominated, to the leading order, by the second term, $$\sigma _1^2_R\frac{r^2}{16\pi ^2L}\underset{n=1}{\overset{\mathrm{}}{}}\frac{\gamma ^2}{n^3}=\frac{\zeta (3)r^2\gamma ^2}{16\pi ^2L},$$ (66) where $`\zeta (3)`$ is the Riemann-zeta function. So, the mean deviation from the classical propagation time due to the lightcone fluctuations is $$\mathrm{\Delta }t\sqrt{\frac{\zeta (3)}{16\pi ^2L}}\gamma =\sqrt{\frac{\zeta (3)}{16\pi ^2L}}\sqrt{32\pi G_5}\gamma =\sqrt{\frac{2\zeta (3)G_4}{\pi }}\gamma \left(\frac{r}{L}\right)t_{pl}.$$ (67) Here we have used the fact that $`G_5=G_4L`$, and $`t_{pl}5.39\times 10^{44}s`$ is the Planck time. This result reveals that here the mean deviation in the arrival time increases linearly In the usual four dimensional case with one compactified spatial dimension $`\mathrm{\Delta }t`$ grows linearly with the square root of $`r`$ (see Ref. ). with $`r`$ and inversely with the size of the extra dimension. This effect is counter-intuitive in the sense that it grows as the size of the compactified dimension decreases. When $`r`$ is of cosmological dimensions and $`L`$ is sufficiently small, the effect is potentially observable. #### 2 Choice of Renormalization and Changes in $`L`$. Note that here $`\sigma _1^2_R`$ has been renormalized to zero as $`L\mathrm{}`$. This is the most natural choice of renormalization, corresponding to the effect of the graviton fluctuations vanishing in the limit of noncompactified spacetime. This is analogous to setting a Casimir energy density to zero in the limit of infinite plate separation, However, if instead of renormalizing $`\sigma _1^2`$ against the vacuum with respect to $`L\mathrm{}`$, we take the manifold with compactified extra dimensions at some fixed sizes $`L`$ to have $`\sigma _1^2_R=0`$, then the lightcone fluctuations could seem to be renormalized away. The latter renormalization scheme is a logical possibility that we can consider, although it is unnatural as there seems to be nothing in the theory which picks out a particular finite value of $`L`$. In any case, if $`L`$ is somehow allowed to vary, for example, as the universe evolves, then lightcone fluctuations would produce noticeable effects, as one could at most set $`\sigma _1^2_R=0`$ at one point along the path of a light ray. It is particularly so, when we try to detect the spread in the arrival times of pulses from distant astrophysical sources, where we are looking back in time. Hence significant lightcone fluctuations may arise no matter what renormalization scheme one chooses if the size $`L`$ is allowed to vary. To get an understanding for the case of a changing $`L`$, let us assume that $`L`$ changes with time at an extremely small rate, which is in fact required by experimental data on the time evolution of fundamental constants (see, for example, Refs. ). Then the evolution of $`L`$ can be reasonably well approximated by a linear function of time: <sup>\**</sup><sup>\**</sup>\**Strictly speaking, the functional dependence of $`L`$ on time should be given by a yet-unknown underlying dynamical theory which governs how the extra dimensions evolve. However, the assumption of a linear dependence is good enough for our purpose of getting a basic idea about how the variation of $`L`$ over time would affect our results. $$L=L_i+\alpha t$$ (68) Using this expression for $`L`$ and redoing the calculations <sup>††</sup><sup>††</sup>††It is worth pointing out here that the graviton two-point function obtained simply by replacing the constant $`L`$ with a changing one is not the exact two-point function that satisfies the appropriate equations, but it is a very good leading order approximation provided that $`\alpha 1`$, one finds, $`\sigma _1^2_R`$ $`=`$ $`{\displaystyle \frac{r^2}{32\pi ^2L_f}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\left({\displaystyle \frac{L_f}{L_i}}\right)^2\left[{\displaystyle \frac{8}{m}}\mathrm{ln}\left(1+{\displaystyle \frac{\gamma ^2}{m^2}}\right){\displaystyle \frac{2\gamma ^2}{m^3}}{\displaystyle \frac{8\gamma ^2}{(\gamma ^2+m^2)m}}\right]`$ (69) $``$ $`{\displaystyle \frac{\zeta (3)r^2\gamma ^2}{16\pi ^2L_f}}\left({\displaystyle \frac{L_f}{L_i}}\right)^2,`$ (70) where $`L_i`$ is the initial compactification size when the light ray is emitted, $`L_f=L_i+\alpha r`$ is the final size at the time of reception and $`\gamma =r/L_f`$. Here $`\sigma _1^2_R`$ is renormalized with respect to $`L\mathrm{}`$. Therefore, one has for the mean time deviation from the classical propagation time $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\left(\frac{L_f}{L_i}\right)\left(\frac{r}{L_f}\right)t_{pl}.$$ (71) Another possibility, as we have mentioned earlier, is to renormalize $`\sigma _1^2`$ against that corresponding to the current size $`L_f`$, which implements the idea of setting the renormalized $`\sigma _1^2_R`$ to be zero if $`L`$ is fixed always or at least during the propagation of the light. This is accomplished by taking the difference between Eq. (70) and Eq. (66) with $`L`$ being replaced by $`L_f`$ to obtain $$\sigma _1^2_R\left[1\left(\frac{L_f}{L_i}\right)^2\right]\frac{\zeta (3)r^2\gamma ^2}{16\pi ^2L_f},$$ (72) which leads to $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\sqrt{\left|1\left(\frac{L_f}{L_i}\right)^2\right|}\left(\frac{r}{L_f}\right)t_{pl}.$$ (73) Equations (71) and (73) demonstrate clearly that no matter what renormalization scheme is employed, one gets a nonzero lightcone fluctuation effect as long as $`L`$ is changing. We want to point out here again that renormalizing $`\sigma _1^2`$ with respect to $`L\mathrm{}`$ is far more natural than to a particular finite size $`L_f`$, since the latter seems to pick out a preferred size $`L_f`$ without any convincing reason to do so. #### 3 Correlation of Pulses The fluctuation in the flight time of pulses, $`\mathrm{\Delta }t`$, can apply to the successive wave crests of a plane wave. This leads to a broadening of spectral lines from a distant source. Note, however, that $`\mathrm{\Delta }t`$ is the expected variation in the arrival times of two successive crests only when the successive pulses are uncorrelated . To determine the correlation, we need to compare $`|\sigma _1^2|`$ and $`|\sigma _1\sigma _1^{}|`$. The latter quantity is defined by $$\sigma _1\sigma _1^{}=\frac{1}{8}(\mathrm{\Delta }r)^2_{r_0}^{r_1}𝑑r_{r_0}^{r_1}𝑑r^{}n^\mu n^\nu n^\rho n^\sigma G_{\mu \nu \rho \sigma }^R(x,x^{}),$$ (74) where the $`r`$-integration is taken along the mean path of the first pulse, and the $`r^{}`$-integration is taken along that of the second pulse. Here we will assume that $`\mathrm{\Delta }t\mathrm{\Delta }r`$, so the slopes, $`v`$ and $`v^{}`$, of the two mean paths are approximately unity. Thus the two-point function in Eq. (74) will be assumed to be evaluated at $`\rho =|𝐱𝐱^{}|=|rr^{}|`$ and $`\tau =|tt^{}|=|rr^{}t_0|`$. If $`|\sigma _1\sigma _1^{}||\sigma _1^2|`$, two pulses are uncorrelated, and otherwise they are correlated. Let us now suppose the time separation of two pulses is $`T`$, and note that the relevant graviton two-point function can be expressed as $`G_{xxxx}(t,x,0,0,0,t^{},x^{},0,0,nL)|_{tt^{}=\rho T}`$ (75) $`={\displaystyle \frac{1}{4\pi ^2}}[{\displaystyle \frac{1}{\beta ^3}}{\displaystyle \frac{8\rho ^2}{\beta (\rho ^2+n^2L^2)^2}}+{\displaystyle \frac{2}{\beta (\rho ^2+n^2L^2)}}`$ (76) $`+{\displaystyle \frac{16\beta \rho ^2}{(\rho ^2+n^2L^2)^3}}{\displaystyle \frac{4\beta }{(\rho ^2+n^2L^2)^2}}+{\displaystyle \frac{2n^2L^2}{\beta ^3(\rho ^2+n^2L^2)}}]`$ (77) $`{\displaystyle \frac{1}{4\pi ^2}}{\displaystyle \frac{1}{\beta ^3}}+{\displaystyle \frac{1}{4\pi ^2}}f(\rho ,n),`$ (78) where, $$\beta (n,\rho )=(n^2L^2+2\rho TT^2)^{1/2}.$$ (80) Utilizing the following integration relation $$_a^b𝑑x_a^b𝑑x^{}f(xx^{})=_0^r(r\rho )[f(\rho )+f(\rho )]𝑑\rho ,$$ (81) one finds that $$\sigma _1\sigma _1^{}=A+B,$$ (82) where $`A`$ $`=`$ $`{\displaystyle \frac{r^2}{16\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle _0^r}𝑑\rho (r\rho )\left[{\displaystyle \frac{1}{\beta (\rho ,n)^3}}+{\displaystyle \frac{1}{\beta (\rho ,n)^3}}\right]`$ (83) $`=`$ $`{\displaystyle \frac{r^2}{16\pi ^2L^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{L^4}{T^2}}\left(2\sqrt{n^2{\displaystyle \frac{T^2}{L^2}}}\sqrt{n^2+{\displaystyle \frac{2rTT^2}{L^2}}}\sqrt{n^2{\displaystyle \frac{2rT+T^2}{L^2}}}\right),`$ (84) and $`B`$ $`=`$ $`{\displaystyle \frac{r^2}{16\pi ^2}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle _0^r}𝑑\rho (r\rho )[f(\rho ,n)+f(\rho ,n)]`$ (86) $`=`$ $`{\displaystyle \frac{r^2}{16\pi ^2L^3}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}\mathrm{\hspace{0.17em}2}[{\displaystyle \frac{2\sqrt{n^2L^2T^2}}{n^2L^2}}+{\displaystyle \frac{\beta (r,n)+\beta (r,n)}{(n^2L^2+r^2)}}+{\displaystyle \frac{2}{nL}}\mathrm{ln}\left({\displaystyle \frac{nL+\sqrt{n^2L^2T^2}}{nL\sqrt{n^2L^2T^2}}}\right]`$ (88) $`{\displaystyle \frac{1}{nL}}\mathrm{ln}\left({\displaystyle \frac{n^2L^2+rT+nL\beta (r,n)}{n^2L^2+rTnL\beta (r,n)}}\right){\displaystyle \frac{1}{nL}}\mathrm{ln}\left({\displaystyle \frac{n^2L^2rT+nL\beta (r,n)}{n^2L^2rTnL\beta (r,n)}}\right)).`$ A few things are to be noticed here: (1) We need to drop those terms in $`A`$ and $`B`$ when the square root is imaginary. (2) It can be shown that the above expression for $`\sigma _1\sigma _1^{}`$ reduces to $`\sigma _1^2`$ when $`T=0`$, as it should. (3) The asymptotic behaviors of the summands when $`n\mathrm{}`$, are $`\frac{2}{n^3}`$ for $`A`$ and $`\frac{2}{n^5}`$ for $`B`$, hence both $`A`$ and $`B`$ converge. (4) If $`rT,L`$, then $`A`$ dominates over $`B`$, since the leading order of the summand in $`A`$ is $`\sqrt{r}`$ while that in $`B`$ is a constant independent of $`r`$ as $`r\mathrm{}`$. Thus $`\sigma _1\sigma _1^{}A`$. To proceed, let us now assume that $`rT`$ and $`rL`$, then $$p\sqrt{\frac{2rT+T^2}{L^2}}\sqrt{\frac{2rT}{L^2}}1$$ (90) is a huge number. Thus for the third term in Eq. (LABEL:eq:A), the sum should only start from $`n=p`$. We can now split the summation into two parts, i.e. terms with $`np`$ and those with $`n>p`$. Using the asymptotic form of the summand for the part with $`n>p`$ and defining $`m=[T/L]`$, where $`[]`$ denotes the integer part, one has $$\sigma _1\sigma _1^{}\frac{r^2}{16\pi ^2L^3}(\frac{2L^4}{T^2}\underset{n=m}{\overset{p}{}}\sqrt{n^2m^2}\frac{L^4}{T^2}\underset{n=1}{\overset{p}{}}\sqrt{n^2+\frac{2rTT^2}{L^2}}+\underset{p}{\overset{\mathrm{}}{}}\frac{2r^2}{n^3}.)$$ (91) Hence, it follows that $$|\sigma _1\sigma _1^{}|<\frac{r^2}{16\pi ^2L^3}(\frac{2L^4}{T^2}\underset{n=1}{\overset{p}{}}n+\frac{L^4}{T^2}\underset{n=1}{\overset{p}{}}\sqrt{n^2+\frac{2rTT^2}{L^2}}+\underset{p}{\overset{\mathrm{}}{}}\frac{2r^2}{n^3}.)$$ (92) Let us now evaluate the above expression term by term. One has, keeping in mind that $`p1`$, that $`{\displaystyle \underset{n=1}{\overset{p}{}}}\sqrt{n^2+{\displaystyle \frac{2rTT^2}{L^2}}}<{\displaystyle \underset{n=1}{\overset{p}{}}}\sqrt{n^2+p^2}=p{\displaystyle \underset{n=1}{\overset{p}{}}}\sqrt{n^2/p^2+1}`$ (93) $`p^2{\displaystyle _{1/p}^1}\sqrt{x^2+1}𝑑x{\displaystyle \frac{1}{2}}[\sqrt{2}+\mathrm{ln}(\sqrt{2}+1)]p^2,`$ (94) and $$\underset{p}{\overset{\mathrm{}}{}}\frac{r^2}{n^3}=\frac{r^2}{2}\mathrm{\Psi }(2,p)\frac{r^2}{2}\frac{1}{p^2}=\frac{1}{4}\frac{L}{r}\frac{L}{T}r^2.$$ (95) Here we have used Eq. (90) and the asymptotic expansion for $`\mathrm{\Psi }(2,x)`$ $$\mathrm{\Psi }(2,x)\frac{1}{x^2}\frac{1}{x^3}\frac{1}{2x^4}+O(1/x^6),$$ (96) where function $`\mathrm{\Psi }(n,x)`$ is defined as $$\mathrm{\Psi }(n,x)=\frac{d^n\psi (x)}{dx^n},\psi (x)=\frac{d}{dx}\mathrm{ln}\mathrm{\Gamma }(x),$$ (97) Noting that for $`p1`$, $$\underset{n=1}{\overset{p}{}}n\frac{1}{2}p^2,$$ (98) and letting $$A=1+\frac{1}{2}[\sqrt{2}+\mathrm{ln}(\sqrt{2}+1)],$$ (99) we finally find $$|\sigma _1\sigma _1^{}|<\frac{r^2}{16\pi ^2L^3}\left(A\frac{L^4}{T^2}p^2+\frac{r^2}{2}\frac{1}{p^2}\right)=\frac{r^2}{16\pi ^2L^3}\left(2A+\frac{1}{2}\right)\frac{L}{r}\frac{L}{T}r^2.$$ (100) Compare this result with $$|\sigma _1^2_R|\frac{r^2}{16\pi ^2L^3}\zeta (3)r^2.$$ (101) It is seen that two successive pulses separated by $`T`$ in time are uncorrelated ($`|\sigma _1\sigma _1^{}||\sigma _1^2_R|`$) provided that $$r\frac{L^2}{T},$$ (102) or equivalently $$T\frac{L^2}{r}=\frac{L}{r}L.$$ (103) However, if $`rL`$, one can show, by series expanding both $`A`$ and $`B`$, that $$|\sigma _1\sigma _1^{}|\frac{r^2}{16\pi ^2L^3}\underset{n=1}{\overset{\mathrm{}}{}}\frac{1}{(n^2\frac{T^2}{L^2})^{3/2}}.$$ (104) Clearly, in this case, $`|\sigma _1\sigma _1^{}||\sigma _1^2_R|`$, if $`TL`$, and $`|\sigma _1\sigma _1^{}||\sigma _1^2_R|`$, when $`T<L`$. A few comments are now in order about the physical picture behind our correlation results. It is natural to expect from the configuration that the dominant contributions to the light cone fluctuation come from the graviton modes with wavelengths of the order of $`L`$. In other words, the lightcone fluctuates on a typical time scale of $`1/L`$. If the travel distance, $`r`$, is less than $`L`$, successive pulses are uncorrelated only when their time separation is greater than the typical fluctuation time scale. Otherwise they are correlated because the quantum gravitational vacuum fluctuations are not significant enough in the interval between the pulses. However, if $`rL^2/T`$, then successive pulses are in general uncorrelated. Thus the correlation time for large $`r`$ is of order $`L^2/r`$, which is much smaller than the compactification scale $`L`$. We can understand this result as arising from the loss of correlation as the pulses propagate over an increasing distance. #### 4 Observational Limits Suppose that the experimental fractional resolution for a particular spectral line of period $`T`$ is $`\mathrm{\Gamma }`$. Then we must have $$\frac{\mathrm{\Delta }t}{T}\mathrm{\Gamma },$$ (105) which, with Eq. (67), leads to a bound on $`L`$ of $$L\frac{rt_{pl}}{\mathrm{\Gamma }T},$$ (106) assuming $`L`$ does not change over time. However, this bound can be trusted only when two successive wave crests are uncorrelated, when $`\mathrm{\Delta }t`$ is the expected variation in their arrival times. The most conservative constraint from this requirement is that $`L`$ is smaller than the wavelength of the spectral line, $`T`$. Namely, $$\frac{rt_{pl}}{\mathrm{\Gamma }T}T,$$ (107) yielding a restriction on the range of spectral lines that we should use to get the lower bound $$T\sqrt{\frac{rt_{pl}}{\mathrm{\Gamma }}}.$$ (108) If Eq. (108) is approximately an equality, then Eq. (106) becomes $$L\left(\frac{rt_{pl}}{\mathrm{\Gamma }}\right)^{\frac{1}{2}}=\left(\frac{rl_{pl}}{\mathrm{\Gamma }}\right)^{\frac{1}{2}}.$$ (109) Obviously, the optimal lower bound would be deduced from the spectral lines of distant galaxies, possibly of cosmological distance, with the highest observed spectral resolution. For astrophysical sources of cosmological distance, spectral lines which satisfy Eq. (108) will have wavelengths $`>1`$mm assuming a resolution of $`\mathrm{\Gamma }10^3`$ and a cosmological travel distance. The detection of CO(1$``$0) line emission at 2.6 mm from luminous infrared galaxies and quasars provides the type of data needed to get a bound. According to Ref. , the observed CO line resolution for the infrared quasar IRAS 07598+6508 , which is at a distance of 596 Mpc assuming $`H_0=75`$ km s<sup>-1</sup> Mpc<sup>-1</sup>, is about $`\mathrm{\Gamma }10^3`$, leading to $$L>10^1\mathrm{mm}.$$ (110) Here, the size of the extra dimension has to be macroscopically large in order not to contradict the astrophysical observation. The lower bound given here is within the sensitivity of the recently proposed experiments for possible deviations from Newtonian gravity. Note, however, this bound could be pushed to an even higher value if we do not require $`L<T`$. This is in fact legitimate, since for astrophysical sources one usually has both $`rL`$ and $`rT`$ satisfied, and the condition for two successive pulses to be uncorrelated is really $$r>\frac{L^2}{T}.$$ (111) Hence, we can use Eq. (106) to get a bound from experimental data as long as the resulting bound, $`L`$, obeys Eq. (111), $$L<\sqrt{rT}L_0.$$ (112) Thus, a stronger bound can be achieved, if we can find astrophysical sources of cosmological distance with observed spectra of much smaller wavelengths provided this condition is satisfied. The observation of $`\gamma `$ rays from astrophysical sources, such as gamma-ray bursters (GRBs) , provides such an opportunity. The use of these sources as probes of possible quantum gravity effects has been explored by a number of authors recently. Below we will select some of these $`\gamma `$ ray sources to calculate both the lower bound $`L_b`$ from Eq. (106) and $`L_0`$ from Eq. (112). To be conservative, we shall assume a resolution of the order of unity for all the gamma rays we are going to consider, since we should at least have this resolution before we can talk with confidence about the observed energies (or frequencies) of the gamma rays. In the following Table, we list the source names, the observed frequencies $`\nu `$, the source distances $`D`$, and the calculated $`L_b`$ and $`L_0`$. Table: Bounds on L from GRB sources | Source | D(Mpc) | $`\nu (Hz)`$ | $`L_b(mm)`$ | $`L_0(mm)`$ | | --- | --- | --- | --- | --- | | GRB 930229 | 791 | $`4.8\times 10^{19}`$ | $`6.1\times 10^4`$ | $`1.2\times 10^{10}`$ | | GRB 940217 | 385 | $`4.3\times 10^{24}`$ | $`2.6\times 10^9`$ | $`2.8\times 10^7`$ | | GRB 930131 | 260 | $`1.1\times 10^{23}`$ | $`4.6\times 10^7`$ | $`1.5\times 10^8`$ | | Mrk 421 | 112 | $`4.8\times 10^{26}`$ | $`8.7\times 10^{10}`$ | $`1.4\times 10^6`$ | | GRB 980703 | 1592 | $`1.2\times 10^{20}`$ | $`3.1\times 10^5`$ | $`1.0\times 10^{10}`$ | | GRB 980425 | 40 | $`5.4\times 10^{20}`$ | $`3.4\times 10^4`$ | $`8.3\times 10^8`$ | | GRB 990123 | 2400 | $`3.0\times 10^{20}`$ | $`1.2\times 10^6`$ | $`8.5\times 10^9`$ | From this Table, one can find that the largest lower bound for $`L`$ comes from GRB930131 and GRB990123, which is $$L>10^7\mathrm{mm}.$$ (113) For some sources, Mrk 421, for instance, one seems to get a much larger bound, which however can not be trusted, because the correlation condition $`L_b<L_0`$ is violated. The physical reason is that the frequencies for the gamma rays are so high that even travel over a cosmological distance does not wash out the correlation between successive wave crests. In principle, the above results only apply to the case where $`L`$ is fixed. When $`L`$ changes as the universe evolves, one should use Eq. (71), or Eq. (73), for the mean time deviation, $`\mathrm{\Delta }t`$. In this case, we can set either a bound on $`L`$ or a bound on the rate of change of $`L`$, if we can constrain either the rate of change of the size of extra dimensions over time or the present size $`L`$ from other considerations. Let $`L_f`$ be the size of the extra dimension at the present time and $`L_i`$ be that at the time of primordial nucleosynthesis, and write $$\frac{L_f}{L_i}=1+\delta .$$ (114) According to Refs. , the observational limits, obtained by investigating the effects on the primordial nucleosynthesis of $`{}_{}{}^{4}\mathrm{He}`$ as a consequence of the time variation of fundamental constants such as the electroweak, strong, and gravitational coupling constants, imply that $`\delta <0.01`$. However, stronger limits on $`\delta `$, which may be less reliable, arise from a detailed study of the events which took place $`1.8\times 10^9`$yr ago on the current site of an open-pit uranium mine at Oklo in the West African Republic of Gabon . This site gave rise to a natural nuclear reactor when it went critical for a period about $`1.8\times 10^9`$yr ago. The Oklo samples constrain the rate of change of extra spatial dimensions to satisfy the limits $$\left|\frac{\dot{L}}{L}\right|1.9\times 10^{19}\mathrm{yr}^1,$$ (115) which translates to $`\delta <10^{10}`$. In either case, $`\delta 1`$. Thus all the results obtained so far for a fixed $`L`$ hold for a changing $`L`$ if renormalization with respect to $`L\mathrm{}`$ is adopted (refer to Eq. (71)). However if one chooses the renormalization with respect to the current size $`L_f`$, then combination of Eq. (105) and Eq. (73) gives rise to $$L\sqrt{\delta ^2+2\delta }\left(\frac{rt_{pl}}{T\mathrm{\Gamma }}\right)\sqrt{\delta }\left(\frac{rt_{pl}}{T\mathrm{\Gamma }}\right).$$ (116) As a result, the following considerably smaller lower bounds can be deduced from Eq. (116) using the CO line data $`L>10^2\mathrm{mm},`$ $`\mathrm{for}\delta =0.01,`$ (117) $`L>10^6\mathrm{mm},`$ $`\mathrm{for}\delta =10^{10},`$ (118) and $`L>10^6\mathrm{mm},`$ $`\mathrm{for}\delta =0.01,`$ (119) $`L>10^2\mathrm{mm},`$ $`\mathrm{for}\delta =10^{10},`$ (120) using GRB data. On the other hand, one can also use our result to place a bound on the rate of change of $`L`$ if the present size of $`L`$ can be fixed by other considerations. In this respect, it is well-known that the five-dimensional Kaluza-Klein theory provides an explanation of the quantization of electric charge, in the sense that all electric charges are multiples of the elementary charge $$e=\frac{4\sqrt{\pi G}}{L}.$$ (121) The corresponding fine structure constant is then $$\alpha _f=\frac{4G}{L^2}.$$ (122) Setting $`\alpha _f`$ to its present value, $`1/137`$, we get an estimate of the size of the fifth dimension, $`L10^{31}`$ cm, in the original Kaluza-Klein model. For the case of renormalization with respect to $`L\mathrm{}`$, we can see from Eq. (105), Eq. (71) and Eq. (114) that we must have $`\delta <1`$, which is much weaker than the existing bounds on the change rate from primordial nucleosynthesis and the Oklo samples. To discuss the case of renormalization with respect to $`L_f`$, the current size, let us write Eq. (116) as $$\delta \left(\frac{LT\mathrm{\Gamma }}{rt_{pl}}\right)^2.$$ (123) Thus, using $`L10^{31}`$ cm and the same CO line data as before, we find the following limit for change of the extra dimension in the Kaluza-Klein model $$\delta 10^{58},$$ (124) or the rate of change by dividing $`\delta `$ by the travel time, which is $`10^9`$ yr $$\left|\frac{\dot{L}}{L}\right|10^{67}\mathrm{yr}^1.$$ (125) This is much stronger even than the strongest bound arising from the observational limits on the time evolution of fundamental constants. To conclude, we have demonstrated, in the case of one extra dimension, that the large quantum lightcone fluctuations due to the compactification of the extra dimension require either the size of the extra dimension to be macroscopically large or rate of change of the extra dimension to be extremely small. This result seems to rule out the five dimensional Kaluza-Klein theory, or at the very least, place strong limits on the rate of change of the extra dimension. We must point out that the rate of growth of $`\mathrm{\Delta }t`$ with $`r`$ depends crucially on the number of spatial dimensions. In four dimensions, $`\mathrm{\Delta }t\sqrt{r}`$, while in five dimensions $`\mathrm{\Delta }tr`$. One expects that in larger number of dimensions, there will be an effect of compactification, but its details need to be determined by explicit calculations for particular models. This is the topic for the next section. ### B Higher dimensional models There was no real reason to extend the Kaluza-Klein idea beyond five dimensions until the emergence of non-abelian gauge field theories which have had a profound impact on theoretical physics since their invention by Yang and Mills in 1954. It was suggested by DeWitt as early as in 1963 that a unification of Yang-Mills theories and gravitation could be achieved in a higher dimensional Kaluza-Klein framework. Nowadays, the possibility of unifying all the known interactions in nature in higher dimensional spacetimes has been actively pursued in the context of eleven dimensional supergravity and ten dimensional superstring theories. A necessary ingredient of all these higher dimensional models is the compactification of the extra dimensions to a very small size so as to leave the ordinary four-dimensional “large” world. In this section, we examine, case by case, the lightcone fluctuations arising from quantum gravitational vacuum fluctuations induced by the periodic compactification of extra, flat spatial dimensions in higher dimensional models up through 11 dimensions and make a conjecture about arbitrary dimensions. ##### a The case with $`n=2`$ This is the 6 dimensional spacetime with 2 extra dimensions. We obtain, after performing the differentiation in both $`F_{xx}`$ and $`H_{xxxx}`$ (See Eqs (LABEL:eq:Feven) and (LABEL:eq:Heven)). $$G_{xxxx}(t,x,\mathrm{𝟎},t^{},x^{},0,0,m_1L,m_2L)A_2(\rho )+B_2(\rho )$$ (126) with $$A_2(\rho )=\frac{1}{4\pi ^3}\left[\frac{54\rho ^6}{(\rho ^2+\alpha _2^2)^5}+\frac{102\alpha _2^2\rho ^4}{(\rho ^2+\alpha _2^2)^5}\frac{3\alpha _2^4\rho ^2}{2(\rho ^2+\alpha _2^2)^5}\right],$$ (127) and $`B_2(\rho )`$ $`=`$ $`{\displaystyle \frac{1}{4\pi ^3}}[{\displaystyle \frac{6\rho ^7}{(\rho ^2+\alpha _2^2)^{\frac{11}{2}}}}{\displaystyle \frac{27\alpha _2^2\rho ^5}{(\rho ^2+\alpha _2^2)^{\frac{11}{2}}}}+{\displaystyle \frac{27\alpha _2^4\rho ^3}{4(\rho ^2+\alpha _2^2)^{\frac{11}{2}}}}+{\displaystyle \frac{3\alpha _2^6\rho }{8(\rho ^2+\alpha _2^2)^{\frac{11}{2}}}}]\times `$ (129) $`\mathrm{ln}\left({\displaystyle \frac{\sqrt{\rho ^2+\alpha _2^2}+\rho }{\sqrt{\rho ^2+\alpha _2^2}\rho }}\right)^2,`$ where we have introduced a parameter $$\alpha _n^2=L^2\underset{i=1}{\overset{n}{}}m_i^2.$$ (130) Upon noting that $`G_{xxxx}`$ is an even function of $`\rho `$, it follows that, $$_a^b𝑑x_a^b𝑑x^{}G_{xxxx}^R(t,x,\mathrm{𝟎},t^{},x^{},\mathrm{𝟎})=\underset{i=1}{\overset{2}{}}\underset{m_i=\mathrm{}}{\overset{+\mathrm{}}{}}_0^r2(r\rho )(A_2+B_2)𝑑\rho .$$ (131) Performing the integration (integrate by parts for those terms involving logarithmic functions), we find $`{\displaystyle \underset{i=1}{\overset{2}{}}}{\displaystyle \underset{m_i=\mathrm{}}{\overset{+\mathrm{}}{}}}{\displaystyle _0^r}2(r\rho )(A_2+B_2)𝑑\rho `$ (132) $`={\displaystyle \frac{1}{2\pi ^3L^2}}{\displaystyle \underset{m_1=1}{\overset{\mathrm{}}{}}^{}}{\displaystyle \underset{m_2=1}{\overset{\mathrm{}}{}}^{}}[{\displaystyle \frac{\gamma ^2(m_1^2+m_2^2)}{(\gamma ^2+m_1^2+m_2^2)^3}}+{\displaystyle \frac{10\gamma ^4}{3(\gamma ^2+m_1^2+m_2^2)^3}}`$ (133) $`{\displaystyle \frac{8\gamma ^6}{3(m_1^2+m_2^2)(\gamma ^2+m_1^2+m_2^2)^3}}{\displaystyle \frac{8\gamma ^5+4(m_1+m_2^2)\gamma ^3+(m_1^2+m_2^2)^2\gamma }{2(m_1^2+m_2^2+\gamma ^2)^{7/2}}}`$ (134) $`\times \mathrm{ln}\left({\displaystyle \frac{\sqrt{m_1^2+m_2^2+\gamma ^2}+\gamma }{\sqrt{m_1^2+m_2^2+\gamma ^2}\gamma }}\right)],`$ (135) (136) where $`\gamma =r/L`$ is a dimensionless parameter. The above double summation is by no means easy to evaluate. However, we are interested in the case in which the travel distance $`r`$ is much greater than the size of the extra dimensions $`L`$, i.e., $`\gamma 1`$. It then follows that the summation is dominated, to the leading order, by $$\frac{1}{2\pi ^3L^2}\underset{m_1=1}{\overset{\mathrm{}}{}}^{}\underset{m_2=1}{\overset{\mathrm{}}{}}^{}\frac{8\gamma ^6}{3(m_1^2+m_2^2)(\gamma ^2+m_1^2+m_2^2)^3},$$ (137) which can be approximated by the following integral when $`\gamma 1`$ $$\frac{4}{3\pi ^3L^2}_{1/\gamma }^{\mathrm{}}𝑑x_1_{1/\gamma }^{\mathrm{}}𝑑x_2\frac{1}{(x_1^2+x_2^2)(x_1^2+x_2^2+1)^3}\frac{4}{3\pi ^3L^2}\mathrm{ln}(\gamma ),\mathrm{as}\gamma \mathrm{}.$$ (138) An easy way to see the above behavior is to note that the contribution to the integral is dominated by the region around $`(1/\gamma ,1/\gamma )`$ since the integrand dies away very quickly as $`x_1`$ and $`x_2`$ increase, and then change to polar coordinates to evaluate the integral while using the fact that around $`(1/\gamma ,1/\gamma )`$ the integrand is approximated by $`1/(x_1^2+x_2^2)`$. Therefore one finds that $$\sigma _1^2_R\frac{r^2}{6\pi ^3L^2}\mathrm{ln}(\gamma ),$$ (139) and the mean deviation from the classical propagation time due to the lightcone fluctuations $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\sqrt{\frac{32\pi G_6}{6\pi ^3L^2}}\sqrt{\mathrm{ln}(\gamma )}=\sqrt{\frac{16G_4}{3\pi ^2}}\sqrt{\mathrm{ln}(\gamma )}\frac{4t_{pl}}{\sqrt{3}\pi }\sqrt{\mathrm{ln}(r/L)}.$$ (140) Here we have used the fact that $`G_{4+n}=G_4L^n`$. Note that here the $`\mathrm{\Delta }t`$ dependence on $`\gamma `$ is square root of a logarithmic function which is quite different from linear dependence in the case with $`n=5`$ discussed above. So, the mean time deviation here grows much more slowly as the travel distance $`r`$ increases or the compactification scale decreases. As we shall see later, this seems to a general feature for $`n>5`$. ##### b The case with n=3 This is the 7 dimensional spacetime with 3 extra dimensions. The relevant two-point function can be shown to be $`G_{xxxx}(t,x,\mathrm{𝟎},t^{},x^{},0,0,m_1L,\mathrm{},m_3L)`$ (141) $`={\displaystyle \frac{1}{8\pi ^3}}\left[{\displaystyle \frac{24\rho ^8}{\alpha _3(\rho ^2+\alpha _3^2)^6}}{\displaystyle \frac{336\alpha _3\rho ^6}{(\rho ^2+\alpha _3^2)^6}}+{\displaystyle \frac{280\alpha _3^3\rho ^4}{2(\rho ^2+\alpha _2^3)^6}}\right],`$ (142) which leads to $`{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}G_{xxxx}^R(t,x,\mathrm{𝟎},t^{},x^{},\mathrm{𝟎})`$ (143) $`={\displaystyle \frac{1}{4\pi ^3L^3}}{\displaystyle \underset{i=1}{\overset{3}{}}}{\displaystyle \underset{m_1=1}{\overset{\mathrm{}}{}}^{}}({\displaystyle \frac{8\gamma ^8}{3(_{i=1}^3m_i^2)^{\frac{3}{2}}(\gamma ^2+_{i=1}^3m_i^2)^4}}`$ (144) $`+{\displaystyle \frac{56\gamma ^6}{3(_{i=1}^3m_i^2)^{\frac{1}{2}}(\gamma ^2+_{i=1}^3m_i^2)^4}}).`$ (145) The triple summation is dominated, to the leading order, by the first term when $`\gamma 1`$, which again can be approximated by integration to be $$\frac{2}{3\pi ^3L^3}_{1/\gamma }^{\mathrm{}}𝑑x_1_{1/\gamma }^{\mathrm{}}𝑑x_2_{1/\gamma }^{\mathrm{}}𝑑x_3\frac{1}{(_{i=1}^3x_i^2)^{\frac{3}{2}}(_{i=1}^3+1)^4}\frac{2}{3\pi ^3L^3}\mathrm{ln}(\gamma ),$$ (147) as $`\gamma \mathrm{}`$. Thus the mean deviation from the classical propagation time due to the lightcone fluctuations $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\sqrt{\frac{32\pi G_7}{12\pi ^3L^3}}\sqrt{\mathrm{ln}(\gamma )}=\sqrt{\frac{8G_4}{3\pi ^2}}\sqrt{\mathrm{ln}(\gamma )}\frac{2\sqrt{2}t_{pl}}{\sqrt{3}\pi }\sqrt{\mathrm{ln}(\gamma )}.$$ (148) ##### c The case with n=4 This is the 8 dimensional spacetime with 4 extra dimensions. The relevant two-point function is $$G_{xxxx}(t,x,\mathrm{𝟎},t^{},x^{},0,0,m_1L,\mathrm{},m_4L)A_4(\rho )+B_4(\rho )$$ (149) with $`A_4={\displaystyle \frac{1}{8\pi ^4}}`$ $`[`$ $`{\displaystyle \frac{912\rho ^6}{(\rho ^2+\alpha _4^2)^6}}+{\displaystyle \frac{902\alpha _4^2\rho ^4}{(\rho ^2+\alpha _4^2)^6}}+{\displaystyle \frac{201\alpha _4^4\rho ^2}{2(\rho ^2+\alpha _4^2)^6}}`$ (151) $`+{\displaystyle \frac{24\rho ^8}{\alpha _4^2(\rho ^2+\alpha _4^2)^6}}{\displaystyle \frac{12\alpha _4^6}{(\rho ^2+\alpha _4^2)^6}}],`$ and $`B_4`$ $`=`$ $`{\displaystyle \frac{1}{8\pi ^4}}[{\displaystyle \frac{240\rho ^7}{(\rho ^2+\alpha _4^2)^{\frac{13}{2}}}}{\displaystyle \frac{540\alpha _4^2\rho ^5}{(\rho ^2+\alpha _4^2)^{\frac{13}{2}}}}+{\displaystyle \frac{90\alpha _4^4\rho ^3}{(\rho ^2+\alpha _4^2)^{\frac{13}{2}}}}+{\displaystyle \frac{15\alpha _4^6\rho }{4(\rho ^2+\alpha _2^2)^{\frac{11}{2}}}}]\times `$ (153) $`\mathrm{ln}\left({\displaystyle \frac{\sqrt{\rho ^2+\alpha _4^2}+\rho }{\sqrt{\rho ^2+\alpha _4^2}\rho }}\right)^2.`$ One finds after carrying out the integration $`{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}G_{xxxx}^R(t,x,\mathrm{𝟎},t^{},x^{},\mathrm{𝟎})`$ (154) $`={\displaystyle \frac{1}{4\pi ^4L^4}}{\displaystyle \underset{i=1}{\overset{4}{}}}{\displaystyle \underset{m_i=1}{\overset{\mathrm{}}{}}^{}}[{\displaystyle \frac{64\gamma ^6}{3(_{i=1}^4m_i^2)(\gamma ^2+_{i=1}^4m_i^2)^4}}`$ (155) $`{\displaystyle \frac{10\gamma ^8}{3(_{i=1}^4m_i^2)^2(\gamma ^2+_{i=1}^4m_i^2)^4}}+{\displaystyle \frac{32\gamma ^4}{(\gamma ^2+_{i=1}^4m_i^2)^4}}+{\displaystyle \frac{15\gamma ^2(_{i=1}^4m_i^2)}{(\gamma ^2+_{i=1}^4m_i^2)^4}}`$ (156) $`{\displaystyle \frac{48\gamma ^5+16(_{i=1}^4m_i^2)\gamma ^3+3(_{i=1}^4m_i^2)^2\gamma }{2(_{i=1}^4m_i^2+\gamma ^2)^{9/2}}}\mathrm{ln}\left({\displaystyle \frac{\sqrt{\underset{i=1}{\overset{4}{}}m_i^2+\gamma ^2}+\gamma }{\sqrt{_{i=1}^4m_i^2+\gamma ^2}\gamma }}\right)].`$ (157) (158) The summation is seen to be dominated by the second term when $`\gamma 1`$ and in that case the summation turns out to be approximated by an integral as $$\frac{5}{6\pi ^4L^4}\underset{i=1}{\overset{4}{}}_{1/\gamma }^{\mathrm{}}𝑑x_i\frac{1}{(_{i=1}^5x_i^2)^2(_{i=1}^5x_i^2+1)^4}\frac{5}{6\pi ^4L^4}\mathrm{ln}(\gamma ).$$ (159) Therefore, one obtains $$\sigma _1^2_R\frac{5r^2}{48\pi ^4L^4}\mathrm{ln}(\gamma ),$$ (160) and $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\frac{\sqrt{10}t_{pl}}{\sqrt{3}\pi ^{\frac{3}{2}}}\sqrt{\mathrm{ln}(\gamma )}.$$ (161) ##### d The case with n=5 This is the 9 dimensional spacetime with 5 extra dimensions. One finds in this case that $`G_{xxxx}(t,x,\mathrm{𝟎},t^{},x^{},0,0,m_1L,\mathrm{},m_2L)`$ (162) $`={\displaystyle \frac{1}{16\pi ^3}}\left[{\displaystyle \frac{24\rho ^{10}}{\alpha _5^3(\rho ^2+\alpha _5^2)^7}}{\displaystyle \frac{648\rho ^8}{\alpha _5(\rho ^2+\alpha _5^2)^7}}{\displaystyle \frac{4536\alpha _5\rho ^6}{(\rho ^2+\alpha _5^2)^7}}+{\displaystyle \frac{2520\alpha _5^3\rho ^4}{2(\rho ^2+\alpha _5^3)^7}}\right],`$ (163) (164) and $`{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}G_{xxxx}^{(1)R}(t,x,0,0,t^{},x^{},0,m_iL)`$ (165) $`={\displaystyle \frac{1}{8\pi ^3L^5}}{\displaystyle \underset{i=1}{\overset{5}{}}}{\displaystyle \underset{m_i=1}{\overset{\mathrm{}}{}}^{}}({\displaystyle \frac{8\gamma ^{10}}{(_{i=1}^5m_i^2)^{\frac{5}{2}}(\gamma ^2+_{i=1}^5m_i^2)^5}}+{\displaystyle \frac{48\gamma ^8}{(_{i=1}^5m_i^2)^{\frac{3}{2}}(\gamma ^2+_{i=1}^5m_i^2)^5}}`$ (166) $`+{\displaystyle \frac{168\gamma ^6}{(_{i=1}^5m_i^2)^{\frac{1}{2}}(\gamma ^2+_{i=1}^5m_i^2)^5}}).`$ (167) The summation, to the leading order, can be approximated by integration $$\frac{1}{\pi ^4L^5}\underset{i=1}{\overset{5}{}}_{1/\gamma }^{\mathrm{}}𝑑x_i\frac{1}{(_{i=1}^5x_i^2)^{\frac{5}{2}}(_{i=1}^5x_i^2+1)^5}\frac{1}{\pi ^4L^5}\mathrm{ln}(\gamma ),$$ (168) as $`\gamma \mathrm{}`$. Thus the mean deviation from the classical propagation time due to the lightcone fluctuations $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\frac{2t_{pl}}{\pi ^{\frac{3}{2}}}\sqrt{\mathrm{ln}(\gamma )}.$$ (169) ##### e The case with n=6 This is the 10 dimensional spacetime with 4 extra dimensions as motivated by superstring theory. The relevant two-point function is given by $$G_{xxxx}(t,x,\mathrm{𝟎},t^{},x^{},0,0,m_1L,\mathrm{},m_6L)A_6(\rho )+B_6(\rho ),$$ (170) where $`A_6={\displaystyle \frac{1}{16\pi ^5}}`$ $`[`$ $`{\displaystyle \frac{15102\rho ^6}{(\rho ^2+\alpha _6^2)^7}}+{\displaystyle \frac{9102\alpha _6^2\rho ^4}{(\rho ^2+\alpha _6^2)^7}}+{\displaystyle \frac{4575\alpha _6^4\rho ^2}{2(\rho ^2+\alpha _6^2)^7}}`$ (172) $`+{\displaystyle \frac{816\rho ^8}{\alpha _6^2(\rho ^2+\alpha _6^2)^7}}{\displaystyle \frac{180\alpha _6^6}{(\rho ^2+\alpha _6^2)^7}}+{\displaystyle \frac{48\rho ^{10}}{\alpha _6^4(\rho ^2+\alpha _6^2)^7}}],`$ and $`B_6`$ $`=`$ $`{\displaystyle \frac{1}{16\pi ^5}}[{\displaystyle \frac{4200\rho ^7}{(\rho ^2+\alpha _6^2)^{\frac{15}{2}}}}{\displaystyle \frac{6300\alpha _6^2\rho ^5}{(\rho ^2+\alpha _6^2)^{\frac{15}{2}}}}+{\displaystyle \frac{1575\alpha _6^4\rho ^3}{2(\rho ^2+\alpha _6^2)^{\frac{15}{2}}}}+{\displaystyle \frac{105\alpha _6^6\rho }{4(\rho ^2+\alpha _6^2)^{\frac{11}{2}}}}]\times `$ (174) $`\mathrm{ln}\left({\displaystyle \frac{\sqrt{\rho ^2+\alpha _6^2}+\rho }{\sqrt{\rho ^2+\alpha _6^2}\rho }}\right)^2.`$ One finds after carrying out the integration $`{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}G_{xxxx}^R(t,x,0,0,t^{},x^{},0,m_iL)=`$ (175) $`{\displaystyle \frac{1}{8\pi ^5L^6}}{\displaystyle \underset{i=1}{\overset{6}{}}}{\displaystyle \underset{m_i=1}{\overset{\mathrm{}}{}}^{}}[{\displaystyle \frac{424\gamma ^6}{3(_{i=1}^6m_i^2)(\gamma ^2+_{i=1}^6m_i^2)^5}}{\displaystyle \frac{68\gamma ^8}{3(_{i=1}^6m_i^2)^2(\gamma ^2+_{i=1}^6m_i^2)^5}}`$ (176) $`+{\displaystyle \frac{390\gamma ^4}{(\gamma ^2+_{i=1}^6m_i^2)^5}}+{\displaystyle \frac{195\gamma ^2(_{i=1}^6m_i^2)}{(\gamma ^2+_{i=1}^6m_i^2)^5}}{\displaystyle \frac{4\gamma ^{10}}{3(_{i=1}^6m_i^2)^3(\gamma ^2+_{i=1}^6m_i^2)^5}}`$ (177) $`{\displaystyle \frac{400\gamma ^5+100(_{i=1}^6m_i^2)\gamma ^3+15(_{i=1}^4m_i^2)^2\gamma }{2(_{i=1}^4m_i^2+\gamma ^2)^{11/2}}}\mathrm{ln}\left({\displaystyle \frac{\sqrt{\underset{i=1}{\overset{6}{}}m_i^2+\gamma ^2}+\gamma }{\sqrt{_{i=1}^6m_i^2+\gamma ^2}\gamma }}\right)].`$ (178) (179) The summation is dominated by the second term when $`\gamma 1`$ and thus is approximated by an integral as $$\frac{1}{6\pi ^5L^6}\underset{i=1}{\overset{6}{}}_{1/\gamma }^{\mathrm{}}𝑑x_i\frac{1}{(_{i=1}^6x_i^2)^3(_{i=1}^6x_i^2+1)^5}\frac{1}{6\pi ^5L^6}\mathrm{ln}(\gamma ).$$ (180) Hence, we have $$\sigma _1^2_R\frac{r^2}{48\pi ^5L^6}\mathrm{ln}(\gamma ),$$ (181) and $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\frac{\sqrt{2}t_{pl}}{\sqrt{3}\pi ^2}\sqrt{\mathrm{ln}(\gamma )}.$$ (182) ##### f The case with n=7 This is the 11 dimensional spacetime with 7 extra dimensions . The relevant two-point function can be shown to be $`G_{xxxx}(t,x,\mathrm{𝟎},t^{},x^{},0,0,m_1L,\mathrm{},m_7L)`$ (183) $`={\displaystyle \frac{1}{16\pi ^5}}[{\displaystyle \frac{72\rho ^{12}}{\alpha _7^5(\rho ^2+\alpha _7^2)^8}}+{\displaystyle \frac{1056\rho ^{10}}{\alpha _7^3(\rho ^2+\alpha _7^2)^8}}{\displaystyle \frac{14256\rho ^8}{\alpha _7(\rho ^2+\alpha _7^2)^8}}`$ (184) $`{\displaystyle \frac{66528\alpha _7\rho ^6}{(\rho ^2+\alpha _7^2)^8}}+{\displaystyle \frac{27720\alpha _7^3\rho ^4}{2(\rho ^2+\alpha _7^3)^8}}].`$ (185) (186) Following the same steps, one finds for $`\gamma 1`$ $$\sigma _1^2_R\frac{5}{16\pi ^5L^7}\underset{i=1}{\overset{7}{}}_{1/\gamma }^{\mathrm{}}𝑑x_i\frac{1}{(_{i=1}^7x_i^2)^{\frac{7}{2}}(_{i=1}^7x_i^2+1)^6}\frac{5}{16\pi ^5L^7}\mathrm{ln}(\gamma ),$$ (187) and $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\frac{\sqrt{10}t_{pl}}{\pi ^4}\sqrt{\mathrm{ln}(\gamma )}.$$ (188) Thus we find that in all of these models in which there is more than one flat compactified extra dimension, the lightcone fluctuation effect grows only logarithmically with distance. Hence we are unable to derive any constraints on these models. We have not been able to prove that this behavior holds for any number of flat extra dimensions greater than one, but conjecture that this is the case. ## IV Parallel brane-worlds scenario In this section, we examine the lightcone fluctuations due to the presence of two 3+1 dimensional hyperplane boundaries, i.e., “ 3-branes”, living in extra dimensions separated from each other by some distance. This framework is motivated by the recent proposal to resolve the unnatural hierarchy between the weak and Planck scales . In this scenario, four-dimensional particle theory, such as the Standard Model, is confined to live in one of the branes, but gravity is free to propagate in the higher dimensional bulk. Therefore the bulk spacetime is dynamical, and the 3-branes can not be rigid, but must undergo quantum fluctuations in their positions, which we assume to be order of $`l_p`$, the Planck length in higher dimensions. To be more specific, we suppose one brane is located at the origin and the other at $`(0,0,0,z_1,z_2,\mathrm{},z_n)`$, and we shall examine the effect of lightcone fluctuations due to gravitons in the bulk by looking at a light ray traveling parallel to one of the boundaries, in the $`x`$axis, for example, but separated from it by a distance $`zl_p`$. This feature is to reflect the quantum fluctuations in the position of the brane . We consider both the Neumann and Dirichlet boundary conditions for the graviton field at each brane hyperplane. Once we have the graviton two-point functions without any boundary, those with two parallel plane boundaries can also be found by the method of image sum. If gravitons satisfy the Dirichlet boundary condition, then the renormalized graviton two-point function is given by the following multiple image sum $`G_{\mu \nu \rho \sigma }^R(t,z_i,t^{},z_i^{})`$ (189) $`={\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \underset{m_i=\mathrm{}}{\overset{+\mathrm{}}{}}^{}}G_{\mu \nu \rho \sigma }(t,z_i,t^{},z_i^{}+2m_iL){\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \underset{m_i=\mathrm{}}{\overset{+\mathrm{}}{}}}G_{\mu \nu \rho \sigma }(t,z_i,t^{},z_i^{}+2m_iL)`$ (190) $`={\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \underset{m_i=1}{\overset{\mathrm{}}{}}}[G_{\mu \nu \rho \sigma }(t,z_i,t^{},2m_iL+z_i^{})G_{\mu \nu \rho \sigma }(t,z_i,t^{},2m_iLz_i^{})`$ (191) $`+G_{\mu \nu \rho \sigma }(t,z_i,t^{},2m_iL+z_i^{})G_{\mu \nu \rho \sigma }(t,z_i,t^{},2m_iLz_i^{})]`$ (192) $`G_{\mu \nu \rho \sigma }(t,z_i,t^{},z_i^{})`$ (193) Again the prime denotes omitting the $`m_i=0`$ term in the summation. For the Neumann boundary, the renormalized graviton two-point function becomes $`G_{\mu \nu \rho \sigma }^R(t,z_i,t^{},z_i^{})`$ (195) $`={\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \underset{m_i=\mathrm{}}{\overset{+\mathrm{}}{}}^{}}G_{\mu \nu \rho \sigma }(t,z_i,t^{},z_i^{}+2m_iL)+{\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \underset{m_i=\mathrm{}}{\overset{+\mathrm{}}{}}}G_{\mu \nu \rho \sigma }(t,z_i,t^{},z_i^{}+2m_iL)`$ (196) $`={\displaystyle \underset{i=1}{\overset{n}{}}}{\displaystyle \underset{m_i=1}{\overset{\mathrm{}}{}}}[G_{\mu \nu \rho \sigma }(t,z_i,t^{},2m_iL+z_i^{})+G_{\mu \nu \rho \sigma }(t,z_i,t^{},2m_iLz_i^{})`$ (197) $`+G_{\mu \nu \rho \sigma }(t,z_i,t^{},2m_iL+z_i^{})+G_{\mu \nu \rho \sigma }(t,z_i,t^{},2m_iLz_i^{})]`$ (198) $`+G_{\mu \nu \rho \sigma }(t,z_i,t^{},z_i^{}).`$ (199) Let us examine a light ray propagating along the $`x`$-axis starting from $`(a,0,0,z_1,\mathrm{},z_n)`$ to $`(b,0,0,z_1,\mathrm{},z_n)`$. For simplicity, let us assume that $`z_1=\mathrm{}=z_n=zl_p/\sqrt{n}`$. Then the mean squared fluctuation in the geodesic interval function is $$\sigma _1^2=\frac{1}{8}(ba)^2_a^b𝑑x_a^b𝑑x^{}G_{xxxx}^R(t,x,0,0,z_i,t^{},x^{},0,0,z_i).$$ (200) ### A Five dimensional theory Here we have one extra dimension, and the relevant two-point function is $`G_{xxxx}(t,x,0,0,z_1,t^{},x^{},0,0,z_1^{})`$ $`=`$ $`{\displaystyle \frac{1}{4\pi ^2}}[{\displaystyle \frac{\rho ^6}{(\rho ^2+\mathrm{\Delta }z_1^2)^3|\mathrm{\Delta }z_1|^3}}{\displaystyle \frac{7\rho ^4}{(\rho ^2+\mathrm{\Delta }z_1^2)^3|\mathrm{\Delta }z_1|}}`$ (202) $`+{\displaystyle \frac{9\rho ^2|\mathrm{\Delta }z_1|}{(\rho ^2+\mathrm{\Delta }z_1^2)^3}}{\displaystyle \frac{|\mathrm{\Delta }z_1|^3}{(\rho ^2+\mathrm{\Delta }z_1^2)^3}}],`$ where $`\mathrm{\Delta }z_1=z_1z_1^{}`$. We then have $`{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}G_{xxxx}(t,x,0,0,z_1,t^{},x^{},0,0,z_1^{})`$ (203) $`={\displaystyle \frac{1}{\pi ^2|\mathrm{\Delta }z_1|}}\mathrm{ln}(1+{\displaystyle \frac{r^2}{\mathrm{\Delta }z_1^2}}){\displaystyle \frac{r^2}{4\pi ^2|\mathrm{\Delta }z_1|^3}}{\displaystyle \frac{r^2}{\pi ^2(r^2+\mathrm{\Delta }z_1^2)|\mathrm{\Delta }z_1|}}`$ (204) $`f(\mathrm{\Delta }z_1,r).`$ (205) If gravitons satisfy the Dirichlet boundary conditions, it follows that $`\sigma _1^2`$ $`=`$ $`{\displaystyle \frac{1}{8}}r^2{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}G_{xxxx}^R(t,x,0,0,z_i,t^{},x^{},0,z_i)`$ (206) $`=`$ $`{\displaystyle \frac{1}{8}}r^2f(2z,r)+{\displaystyle \frac{1}{8}}r^2{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}[2f(2mL,r)f(2z2mL,r)f(2z+2mL,r)].`$ (207) Here we are interested in the case in which $`rz`$ and $`Lz`$. One then finds that $$f(2z,r)\frac{1}{L}\frac{r^2}{32\pi ^2z^2}\frac{L}{z}=\frac{1}{L}\frac{\gamma ^2}{32\pi ^2}\left(\frac{L}{z}\right)^3,$$ (209) and $`{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}`$ $`[`$ $`2f(2mL,r)f(2z2mL,r)f(2z+2mL,r)]`$ (214) $`{\displaystyle \frac{1}{32\pi ^2}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\left[{\displaystyle \frac{2r^2}{m^3L^3}}+{\displaystyle \frac{r^2}{(mL+z)^3}}+{\displaystyle \frac{r^2}{(mLz)^3}}\right]`$ $`={\displaystyle \frac{1}{32\pi ^2L}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\left[{\displaystyle \frac{2\gamma ^2}{m^3}}+{\displaystyle \frac{\gamma ^2}{(mz/L)^3}}+{\displaystyle \frac{\gamma ^2}{(m+z/L)^3}}\right]`$ $`{\displaystyle \frac{\gamma ^2}{32\pi ^2L}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}\left[{\displaystyle \frac{12}{m^5}}\left({\displaystyle \frac{z}{L}}\right)^2+{\displaystyle \frac{30}{m^7}}\left({\displaystyle \frac{z}{L}}\right)^4+O((z/L)^6)\right]`$ $`{\displaystyle \frac{\gamma ^2}{32\pi ^2L}}{\displaystyle \underset{m=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{12}{m^5}}\left({\displaystyle \frac{z}{L}}\right)^2={\displaystyle \frac{3\gamma ^2\zeta (5)}{8\pi ^2L}}\left({\displaystyle \frac{z}{L}}\right)^2.`$ Here $`\zeta (5)`$ is the Riemann-zeta function. Consequently, we find $$\sigma _1^2=\frac{\gamma ^2r^2}{8^2\pi ^2L}[\frac{1}{4}(\frac{L}{z})^3+3\zeta (5)(\frac{z}{L})^2],$$ (215) thus, $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\gamma t_{pl}\sqrt{\frac{1}{2\pi }[\frac{1}{4}(\frac{L}{z})^3+3\zeta (5)(\frac{z}{L})^2]}\frac{1}{2\sqrt{2\pi }}(\frac{L}{z})^{\frac{3}{2}}\left(\frac{r}{L}\right)t_{pl}.$$ (216) If instead of the Dirichlet boundary condition, gravitons are forced to satisfy the Neumann boundary condition, one finds $$\sigma _1^2=\frac{\gamma ^2r^2}{8^2\pi ^2L}[\frac{1}{4}(\frac{L}{z})^3+\zeta (3)+3\zeta (5)(\frac{z}{L})^2],$$ (217) and $$\mathrm{\Delta }t=\frac{\sqrt{|\sigma _1^2_R|}}{r}\gamma t_{pl}\sqrt{\frac{1}{2\pi }[\frac{1}{4}(\frac{L}{z})^3+\zeta (3)+3\zeta (5)(\frac{z}{L})^2]}\frac{1}{2\sqrt{2\pi }}(\frac{L}{z})^{\frac{3}{2}}\left(\frac{r}{L}\right)t_{pl}.$$ (218) Here we find that the choice of different boundary conditions has very little effect on the growth of the lightcone fluctuations as long as $`Lz`$, although the fluctuations are slightly larger in the case of the Neumann boundary condition. Comparing this $`\mathrm{\Delta }t`$ with that in the case of periodical compactification Eq. (67), one can see that the effect of light cone fluctuation here is larger for a given size of extra dimensions, $`L`$. Another very distinctive feature is that here $`\mathrm{\Delta }t`$ increases as $`L`$ increases provided that $`L`$ and $`z`$ are independent of each other, which is in sharp contrast with the case of periodical compactification. As before, the fluctuation in the flight time of pulses, $`\mathrm{\Delta }t`$, can be applied to the successive wave crests of a plane wave. This leads to a broadening of spectral lines from luminous sources. In case of periodical compactification, we used the data from gamma ray bursters to derive a very strong lower bound on $`L`$. In the present framework, however, we have an intrinsic lower bound $`Lz`$. Recall that $`z`$ is the Planck length in higher dimensions which may be much larger than that in 4 dimensions. In the picture of Ref. , $`z`$ and $`L`$ are not independent, and may be expressed in terms of the Planck mass in $`4+n`$ dimensions, $`M_p^{(4+n)}`$, as $$l_p=2\times 10^{17}\left(\frac{1Tev}{M_p^{(4+n)}}\right)\mathrm{cm},$$ (219) $$L=10^{30/n17}\left(\frac{1Tev}{M_p^{(4+n)}}\right)^{1+2/n}\mathrm{cm}.$$ (220) One has for the ratio $`L/z`$ $$\frac{L}{z}=\frac{L}{l_p}=0.5\times 10^{30/n}\left(\frac{1Tev}{M_p^{(4+n)}}\right)^{2/n}.$$ (221) Express Eq. (216) or Eq. (218) as $$\mathrm{\Delta }t\frac{1}{2\sqrt{2\pi }}(\frac{L}{z})^{\frac{1}{2}}\left(\frac{r}{z}\right)t_{pl},$$ (222) or equivalently, $$\mathrm{\Delta }t10^{15}\left(\frac{1Tev}{M_p^{(4+n)}}\right)^{1/n1}\frac{r}{10^{17}cm}t_{pl}=10^{32}\left(\frac{1Tev}{M_p^{(4+n)}}\right)^{1/n1}\left(\frac{r}{1cm}\right)t_{pl}.$$ (223) Let $`M_p^{(4+n)}`$ be $`1`$Tev. Then for an astronomical source of cosmological distance, such as gamma-ray bursters with redshift $`1`$, or $`r10^{28}`$cm, we get the following estimate for the $`n=1`$ model $$\mathrm{\Delta }t10^{60}\times 10^{44}s=10^{16}s.$$ (224) That is far too large, so that the $`n=1`$ model can be ruled out completely. ### B Higher dimensional theories We shall only study the 6 dimensional case in some detail, but shall list results for other cases. It is not difficult to see that the relevant two-point function can be obtained from Eq. (126) by replacing $`\alpha _2`$ there with $`_{i=1}^2\mathrm{\Delta }z_i^2`$. Therefore one easily finds from Eq. (136) $`{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}G_{xxxx}(t,x,0,0,z_1,z_2,t^{},x^{},0,0,z_1^{},z_2^{})f_2(\mathrm{\Delta }z_1,\mathrm{\Delta }z_2,r)`$ (225) $`={\displaystyle \frac{1}{4\pi ^3}}[{\displaystyle \frac{r^2(\mathrm{\Delta }z_1^2+\mathrm{\Delta }z_2^2)}{(r^2+\mathrm{\Delta }z_1^2+\mathrm{\Delta }z_2^2)^3}}+{\displaystyle \frac{10r^4}{3(r^2+\mathrm{\Delta }z_1^2+\mathrm{\Delta }z_2^2)^3}}`$ (226) $`{\displaystyle \frac{8r^6}{3(\mathrm{\Delta }z_1^2+\mathrm{\Delta }z_2^2)(r^2+\mathrm{\Delta }z_1^2+\mathrm{\Delta }z_2^2)^3}}`$ (227) $`{\displaystyle \frac{8r^5+4(\mathrm{\Delta }z_1^2+\mathrm{\Delta }z_2^2)r^3+(\mathrm{\Delta }z_1^2+\mathrm{\Delta }z_2^2)^2r}{2(\mathrm{\Delta }z_1^2+\mathrm{\Delta }z_2^2+r^2)^{7/2}}}\mathrm{ln}\left({\displaystyle \frac{\sqrt{\mathrm{\Delta }z_1^2+\mathrm{\Delta }z_2^2+r^2}+r}{\sqrt{\mathrm{\Delta }z_1^2+\mathrm{\Delta }z_2^2+r^2}r}}\right)].`$ (228) Then it follows that $`\sigma _1^2`$ $`=`$ $`{\displaystyle \frac{1}{8}}r^2{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}G_{xxxx}^R(t,x,0,0,z,z,t^{},x^{},0,z,z)`$ (230) $`=`$ $`{\displaystyle \frac{1}{8}}r^2f(2z,2z,r)+{\displaystyle \frac{1}{8}}r^2{\displaystyle \underset{m_1=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m_2=1}{\overset{\mathrm{}}{}}}[2f(2m_1L,2m_2L,r)`$ (232) $`f(2z2m_1L,2zm_2L,r)f(2z+2m_1L,2z+2m_2L,r)].`$ Here we are interested in the case when $`rz`$ and $`Lz`$. One finds that $$f(2z,2z,r)\frac{1}{6\pi ^3L^2}\left(\frac{L}{z}\right)^2\frac{4}{\pi ^3L^2}\frac{\mathrm{ln}(r/z)}{\gamma ^2},$$ (234) and $`g(r,z,L)`$ $``$ $`{\displaystyle \underset{m_1=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m_2=1}{\overset{\mathrm{}}{}}}[\mathrm{\hspace{0.17em}2}f(2m_1L,2m_2L,r)f(2z2m_1L,2z2m_2L,r)`$ (236) $`f(2z+2m_1L,2m_2L,r)]`$ $``$ $`{\displaystyle \frac{1}{4\pi ^3L^2}}{\displaystyle \underset{m_1=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m_2=1}{\overset{\mathrm{}}{}}}({\displaystyle \frac{4\gamma ^6}{3(m_1^2+m_2^2)(\gamma +m_1^2+m_2^2)^3}}`$ (239) $`+{\displaystyle \frac{2\gamma ^6}{3[(m_1+z/L)^2+(m_2+z/L)^2][\gamma ^2+(m_1+z/L)^2+(m_2+z/L)^2]^3}}`$ $`+{\displaystyle \frac{2\gamma ^6}{3[(m_1z/L)^2+(m_2z/L)^2][\gamma ^2+(m_1z/L)^2+(m_2z/L)^2]^3}}).`$ If we assume the travel distance $`r`$ is much larger than the size of the extra dimensions $`L`$, i.e $`\gamma 1`$, $`g(r,z,L)`$ can be approximated by integrals as follows $`g(r,z,L)`$ $``$ $`{\displaystyle \frac{1}{6\pi ^3L^2}}{\displaystyle _{1/\gamma }^{\mathrm{}}}𝑑x_1{\displaystyle _{1/\gamma }^{\mathrm{}}}𝑑x_2{\displaystyle \frac{2}{(x_1^2+x_2^2)(x_1^2+x_2^2+1)^3}}`$ (243) $`+{\displaystyle \frac{1}{6\pi ^3L^2}}{\displaystyle _{\frac{1}{\gamma }(1z/L)}^{\mathrm{}}}𝑑x_1{\displaystyle _{\frac{1}{\gamma }(1z/L)}^{\mathrm{}}}𝑑x_2{\displaystyle \frac{1}{(x_1^2+x_2^2)(x_1^2+x_2^2+1)^3}}`$ $`+{\displaystyle \frac{1}{6\pi ^3L^2}}{\displaystyle _{\frac{1}{\gamma }(1+z/L)}^{\mathrm{}}}𝑑x_1{\displaystyle _{\frac{1}{\gamma }(1+z/L)}^{\mathrm{}}}𝑑x_2{\displaystyle \frac{1}{(x_1^2+x_2^2)(x_1^2+x_2^2+1)^3}}`$ $``$ $`{\displaystyle \frac{1}{6\pi ^3L^2}}\left[2\mathrm{ln}(\gamma )+\mathrm{ln}\left({\displaystyle \frac{1z/L}{\gamma }}\right)+\mathrm{ln}\left({\displaystyle \frac{1+z/L}{\gamma }}\right)\right]`$ (244) $`=`$ $`{\displaystyle \frac{1}{6\pi ^3L^2}}\mathrm{ln}\left[1(z/L)^2\right].`$ (245) Therefore, $`\sigma _1^2`$ $`=`$ $`{\displaystyle \frac{1}{8}}r^2{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}G_{xxxx}^R(t,x,0,0,z,z,t^{},x^{},0,z,z){\displaystyle \frac{r^2}{8}}f(2z,2z,r)`$ (247) $``$ $`{\displaystyle \frac{r^2}{48\pi ^3L^2}}\left({\displaystyle \frac{L}{z}}\right)^2+{\displaystyle \frac{2r^2}{3\pi ^3L^2}}{\displaystyle \frac{\mathrm{ln}(r/z)}{\gamma ^2}}.`$ (248) Hence, $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\sqrt{\frac{2}{3\pi ^2}}\left(\frac{L}{z}\right).$$ (249) Notice here that the leading term does not depend on the travel distance $`r`$ and moreover the next order corrections decrease as $`r`$ increases. Recall that $`\gamma =r/L`$. Interestingly, the lightcone fluctuation grows as the compactification scale increases. However, if we change to the Neumann boundary condition, the behavior will be different as we can see from the following analysis. Let us note that in this case $`\sigma _1^2`$ $`=`$ $`{\displaystyle \frac{1}{8}}r^2{\displaystyle _a^b}𝑑x{\displaystyle _a^b}𝑑x^{}G_{xxxx}^R(t,x,0,0,z,z,t^{},x^{},0,z,z)`$ (250) $`=`$ $`{\displaystyle \frac{1}{8}}r^2f(2z,2z,r)+{\displaystyle \frac{1}{8}}r^2{\displaystyle \underset{m_1=1}{\overset{\mathrm{}}{}}}{\displaystyle \underset{m_2=1}{\overset{\mathrm{}}{}}}[2f(2m_1L,2m_2L,r)`$ (252) $`+f(2z2m_1L,2zm_2L,r)+f(2z+2m_1L,2z+2m_2L,r)].`$ Notice the sign changes in the above expression as compared to Eq (LABEL:eq:sigmaD). Then it follows that $`\sigma _1^2`$ $``$ $`{\displaystyle \frac{r^2}{48\pi ^3L^2}}\left({\displaystyle \frac{L}{z}}\right)^2{\displaystyle \frac{2r^2}{3\pi ^3L^2}}{\displaystyle \frac{\mathrm{ln}(r/z)}{\gamma ^2}}+{\displaystyle \frac{r^2}{48\pi ^3L^2}}\left[2\mathrm{ln}(\gamma )\mathrm{ln}\left({\displaystyle \frac{1z^2/L^2}{\gamma ^2}}\right)\right]`$ (254) $``$ $`{\displaystyle \frac{r^2}{12\pi ^3L^2}}\mathrm{ln}(\gamma ){\displaystyle \frac{r^2}{48\pi ^3L^2}}\left({\displaystyle \frac{L}{z}}\right)^2,`$ (255) and $$\mathrm{\Delta }t=\frac{\sqrt{\sigma _1^2_R}}{r}\sqrt{\left|\frac{2}{3\pi ^2}\left(\frac{L}{z}\right)^2\frac{8}{3\pi ^2}\mathrm{ln}(r/L)\right|}t_{pl}.$$ (256) Here we have two different contributing terms to the mean deviation in time; one is independent of $`r`$ but increases as $`L`$ increases, the other grows logarithmically as $`r`$ increase or $`L`$ decreases. Similarly, we have calculated cases up to $`n=7`$ as motivated by string/M theory. The results are (all in units of $`t_{pl}`$, the Planck time in four dimensions): For n=3, $`\mathrm{\Delta }t\sqrt{{\displaystyle \frac{1}{6\pi ^3}}}\left({\displaystyle \frac{L}{z}}\right)^{\frac{3}{2}},\mathrm{for}\mathrm{Dirichlet}\mathrm{boundary}\mathrm{condition}`$ (257) $`\mathrm{\Delta }t\sqrt{\left|{\displaystyle \frac{1}{6\pi ^3}}\left({\displaystyle \frac{L}{z}}\right)^3{\displaystyle \frac{2}{3\pi ^3}}\mathrm{ln}(r/L)\right|},\mathrm{for}\mathrm{Neumann}\mathrm{boundary}\mathrm{condition}.`$ (258) For n=4, $`\mathrm{\Delta }t\sqrt{{\displaystyle \frac{5}{12\pi ^4}}}\left({\displaystyle \frac{L}{z}}\right)^2,\mathrm{for}\mathrm{Dirichlet}\mathrm{boundary}\mathrm{condition}`$ (260) $`\mathrm{\Delta }t\sqrt{\left|{\displaystyle \frac{5}{12\pi ^4}}\left({\displaystyle \frac{L}{z}}\right)^4{\displaystyle \frac{5}{3\pi ^4}}\mathrm{ln}(r/L)\right|},\mathrm{for}\mathrm{Neumann}\mathrm{boundary}\mathrm{condition}.`$ (261) For n=5, $`\mathrm{\Delta }t{\displaystyle \frac{1}{4\pi ^2}}\left({\displaystyle \frac{L}{z}}\right)^{\frac{5}{2}},\mathrm{for}\mathrm{Dirichlet}\mathrm{boundary}\mathrm{condition}`$ (263) $`\mathrm{\Delta }t\sqrt{\left|{\displaystyle \frac{1}{16\pi ^4}}\left({\displaystyle \frac{L}{z}}\right)^5{\displaystyle \frac{1}{4\pi ^4}}\mathrm{ln}(r/L)\right|},\mathrm{for}\mathrm{Neumann}\mathrm{boundary}\mathrm{condition}.`$ (264) For n=6, $`\mathrm{\Delta }t{\displaystyle \frac{1}{8\sqrt{3}\pi ^{\frac{5}{2}}}}\left({\displaystyle \frac{L}{z}}\right)^3,\mathrm{for}\mathrm{Dirichlet}\mathrm{boundary}\mathrm{condition}`$ (266) $`\mathrm{\Delta }t{\displaystyle \frac{1}{8}}\sqrt{\left|{\displaystyle \frac{1}{3\pi ^5}}\left({\displaystyle \frac{L}{z}}\right)^6{\displaystyle \frac{4}{3\pi ^5}}\mathrm{ln}(r/L)\right|},\mathrm{for}\mathrm{Neumann}\mathrm{boundary}\mathrm{condition}.`$ (267) (268) For n=7, $`\mathrm{\Delta }t{\displaystyle \frac{1}{8}}\sqrt{{\displaystyle \frac{5}{2\pi ^5}}}\left({\displaystyle \frac{L}{z}}\right)^{\frac{7}{2}},\mathrm{for}\mathrm{Dirichlet}\mathrm{boundary}\mathrm{condition}`$ (269) $`\mathrm{\Delta }t{\displaystyle \frac{1}{8}}\sqrt{\left|{\displaystyle \frac{5}{2\pi ^5}}\left({\displaystyle \frac{L}{z}}\right)^7{\displaystyle \frac{10}{\pi ^5}}\mathrm{ln}(r/L)\right|},\mathrm{for}\mathrm{Neumann}\mathrm{boundary}\mathrm{condition}.`$ (270) A few comments are now in order for all the cases we have examined: First, to the leading order $`\mathrm{\Delta }t`$ does not grow as $`r`$ increases for the Dirichlet boundary condition. Second, $`\mathrm{\Delta }t`$ behaves differently for different boundary conditions, but the logarithmic dependence associated with the Neumann boundary condition can be neglected as far as observation is concerned because logarithmic growth is extremely small. Third, our results seem to suggest a leading order behavior of $`(L/z)^{n/2}`$ for any dimensions. Finally, our results are quite different from those obtained by Ref. where only the contribution of the $`F`$ term was considered and only one dominant term in the sums computed. Our results are much smaller than those obtained by these authors because of cancellations among the various terms. We wish to gain an understanding of how large the effect of lightcone fluctuations can be in these higher dimensions and whether they might be observable. From Eq. (221) and the above results, one finds $$\mathrm{\Delta }t10^{15}t_{pl}\left(\frac{1Tev}{M_p^{(4+n)}}\right).$$ (272) This result reveals that the smaller is $`M_p^{(4+n)}`$, the larger is the effect of lightcone fluctuations. According to Ref. , $`M_p^{(4+n)}`$ may as low as the order of one Tev. This leads to $$\mathrm{\Delta }t10^{29}s.$$ (273) This is a tiny effect by conventional standards. However if we note that the above mean time deviation is equivalent to an uncertainty in position $$\mathrm{\Delta }x10^{21}m$$ (274) and the operation of gravity-wave interferometers is based upon the detection of minute changes in the positions of some test masses (relative to the position of a beam splitter), we can see that this effect might be testable in the next LIGO/VIRGO generation of gravity-wave interferometers . It constitutes an additional source of noise due to quantum gravity. Currently, the sensitivity of these gravity-wave interferometers has already reached the order of $`10^{19}`$m . Note that Amelino-Camelia and Ng and van Dam have also proposed rather different quantum gravity effects which might also be detectable by laser interferometers. A more complete discussion of the observability of the $`\mathrm{\Delta }t`$ given by Eq. (272) should involve a calculation of the correlation time, analogous to that performed in Sect. III A 3 for the five dimensional Kaluza-Klein model. This calculation has not yet been performed. However, it is reasonable to guess that the result will be of the order of or less than Tev scale which characterizes this model. Recall that in the five dimensional compactified model, the correlation time was found to be much smaller than the compactification scale when the travel distance is large. If this guess is correct, then the correlation time is much smaller than $`\mathrm{\Delta }t`$ itself. ## V Discussion and conclusions In this paper, we have examined the effects of compactified extra dimensions upon the propagation of light in the uncompactified dimensions. There are nontrivial effects that arise from quantum fluctuations of the gravitational field, induced by the compactification. These effects take the form of lightcone fluctuations, variations in the flight times of pulses between a source and a detector. The crucial quantity describing these fluctuations is $`\mathrm{\Delta }t`$, the expected variation in arrival times of two successive pulses which are separated by more than a correlation time. In Sect. II B, we gave a derivation of the formula for $`\mathrm{\Delta }t`$ using the geodesic deviation equation. This derivation allowed us to discuss issues of gauge and Lorentz invariance. In particular, it demonstrates that $`\mathrm{\Delta }t`$ is gauge invariant. All of the subsequent explicit calculations are performed in the transverse-tracefree gauge. As a prelude, we found the graviton two-point function in this gauge in Minkowski spacetime of arbitrary dimension. The two-point function in a higher dimensional flat compactified spacetime is given as an image sum. We then calculated $`\mathrm{\Delta }t`$ in the five-dimensional Kaluza-Klein model, five-dimensional flat spacetime with one periodic spatial dimension. We found that $`\mathrm{\Delta }t`$ grows linearly with increasing distance between the source and the detector, and is inversely proportional to the compactification length, $`L`$. This result differs from the square root dependence on distance that was found in four-dimensional flat spacetime with one periodic spatial dimension . This demonstrates that lightcone fluctuation effects are rather model-dependent. We also calculated the correlation time in the five-dimensional model and found that it is typically small compared to the compactification scale $`L`$. This allows us to place very tight constraints on the parameters of this model. We favor the viewpoint that the lightcone fluctuation effects should vanish only in the limit that $`L\mathrm{}`$. If one adopts this view, then the five-dimensional Kaluza-Klein model can be ruled out. Data from gamma ray burst sources imply a lower bound on $`L`$ which is larger than upper bounds obtained from other considerations. However, another logical possibility is that lightcone fluctuation effects happen to vanish ($`\sigma _1^2=0`$) at the present compactification scale. Even if one adopts this viewpoint, one still obtains very strong constraints on the fractional change in $`L`$ which can have occurred over a cosmological time scale. This in turn tightly constrains any five-dimensional Kaluza-Klein cosmology. We also examined an alternative five-dimensional model (the brane worlds scenario) in which gravitons satisfy Dirichlet or Neumann boundary conditions on a pair of parallel four-dimensional branes (one of which represents our world). In this model, we again find that the lightcone fluctuations are so large that the model can be ruled out. We next turned our attention to models with more than one extra dimension. In the case of two or more flat, periodically compactified dimensions, we found that $`\mathrm{\Delta }t`$ grows only logarithmically with distance, and that no constraints may be placed on these models. In the case of the brane worlds scenario with more than one extra dimension, we found that $`\mathrm{\Delta }t`$ approaches a constant which can be of the order of $`10^{29}s`$. This would produce a source of noise in laser interferometer detectors of gravity waves, which may eventually be within their range of sensitivity. In summary, compactified extra dimensions have the possibility to produce observable effects by enhancing the quantum fluctuations of the gravitational field. These effects might be used either to place constraints on theories with extra dimensions, or else possibly eventually to provide positive evidence for the existence of extra dimensions. Although our results for more than one flat compactified extra dimension are too weak for either of these purposes, the required calculations for models with curved extra dimensions have not yet been performed. Another model which has not yet been examined in this context is the Randall-Sundrum model , with an uncompactified fifth dimension. In this latter model, propagating graviton modes are effectively confined within a finite volume in the fifth dimension, so one might expect nonzero lightcone fluctuations. ###### Acknowledgements. We would like to thank Ken Olum for helpful discussions. This work was supported in part by the National Science Foundation under Grant PHY-9800965. ## Appendix ## VI Graviton two-point functions in spacetime with arbitrary number of extra dimensions Here we evaluate the functions $`D^n(x.x^{})`$, $`F_{ij}^n(x,x^{})`$ and $`H_{ijkl}^n(x,x^{})`$ defined in Eqs. (30), (29) and (31), respectively. Once these functions are given, the graviton two point functions are easy to obtain. Define $$R=|𝐱𝐱^{}|,\mathrm{\Delta }t=tt^{},k=|𝐤|=\omega ,$$ (A1) and assume $`n`$ extra dimensions, then $`D^n(x,x^{})`$ $`=`$ $`{\displaystyle \frac{Re}{(2\pi )^{3+n}}}{\displaystyle \frac{d^{(3+n)}𝐤}{2\omega }e^{i𝐤(𝐱𝐱^{})}e^{i\omega (tt^{})}}`$ (A2) $`=`$ $`{\displaystyle \frac{Re}{2(2\pi )^{3+n}}}{\displaystyle _0^{\mathrm{}}}k^{n+1}e^{ik\mathrm{\Delta }t}𝑑k{\displaystyle _0^\pi }𝑑\theta _1\mathrm{sin}^{1+n}\theta _1e^{ikR\mathrm{cos}\theta _1}`$ (A4) $`\times {\displaystyle _0^\pi }d\theta _2\mathrm{sin}^n\theta _2\mathrm{}{\displaystyle _0^\pi }d\theta _{n+1}\mathrm{sin}\theta _{n+1}{\displaystyle _0^{2\pi }}d\theta _{n+2}`$ $`=`$ $`{\displaystyle \frac{a_nRe}{2(2\pi )^{3+n}}}{\displaystyle _0^{\mathrm{}}}k^{n+1}e^{ik\mathrm{\Delta }t}𝑑k{\displaystyle _1^1}e^{ikRx}(1x^2)^{n/2}𝑑x`$ (A5) $`=`$ $`{\displaystyle \frac{a_nRe}{(2\pi )^{3+n}}}{\displaystyle _0^{\mathrm{}}}k^{n+1}e^{ik\mathrm{\Delta }t}𝑑k{\displaystyle _0^1}(1x^2)^{n/2}\mathrm{cos}(kRx)𝑑x`$ (A6) $`=`$ $`{\displaystyle \frac{a_n\sqrt{\pi }2^{\frac{n+1}{2}}\mathrm{\Gamma }(\frac{n}{2}+1)Re}{2(2\pi )^{3+n}}}{\displaystyle \frac{1}{R^{\frac{n+1}{2}}}}{\displaystyle _0^{\mathrm{}}}k^{\frac{n+1}{2}}J_{\frac{n+1}{2}}(kR)e^{ik\mathrm{\Delta }t}𝑑k`$ (A7) $`=`$ $`{\displaystyle \frac{a_n\sqrt{\pi }2^{\frac{n+1}{2}}\mathrm{\Gamma }(\frac{n}{2}+1)Re}{2(2\pi )^{3+n}}}{\displaystyle \frac{1}{R^{\frac{n+1}{2}}}}\underset{\alpha 0^++i\mathrm{\Delta }t}{lim}{\displaystyle _0^{\mathrm{}}}k^{\frac{n+1}{2}}J_{\frac{n+1}{2}}(kR)e^{\alpha k}𝑑k`$ (A8) $`=`$ $`{\displaystyle \frac{a_n2^n\mathrm{\Gamma }(\frac{n}{2}+1)^2}{(2\pi )^{3+n}}}{\displaystyle \frac{1}{(R^2\mathrm{\Delta }t^2)^{n/2+1}}}={\displaystyle \frac{\mathrm{\Gamma }(\frac{n}{2}+1)}{4\pi ^{\frac{n+4}{2}}}}{\displaystyle \frac{1}{(R^2\mathrm{\Delta }t^2)^{n/2+1}}}.`$ (A9) Here we have defined $$a_n=_0^\pi 𝑑\theta _2\mathrm{sin}^n\theta _2\mathrm{}_0^\pi 𝑑\theta _{n+1}\mathrm{sin}\theta _{n+1}_0^{2\pi }𝑑\theta _{n+2}=\frac{2\pi ^{\frac{n}{2}+1}}{\mathrm{\Gamma }(\frac{n}{2}+1)},$$ (A11) and used $$_0^1\mathrm{cos}(kRx)(1x^2)^{n/2}𝑑x=\frac{\sqrt{\pi }}{2}\mathrm{\Gamma }(\frac{n}{2}+1)\left(\frac{2}{kR}\right)^{\frac{n+1}{2}}J_{\frac{n+1}{2}}(kR),$$ (A12) and $$_0^{\mathrm{}}e^{\alpha x}J_\nu (\beta x)x^\nu 𝑑x=\frac{(2\beta )^\nu \mathrm{\Gamma }(\nu +1/2)}{\sqrt{\pi }(\alpha ^2+\beta ^2)^{\nu +\frac{1}{2}}},Re\nu >1/2.$$ (A13) When $`n`$ is odd, $`D^n(x,x^{})`$ should be taken to be zero when $`R^2<\mathrm{\Delta }t^2`$. Let us now turn our attention to the calculation of $`F_{ij}`$ and $`H_{jikl}`$. We find $`F_{ij}^n(x,x^{})`$ $`=`$ $`{\displaystyle \frac{Re}{(2\pi )^{3+n}}}{\displaystyle d^{3+n}𝐤\frac{k_ik_j}{2\omega ^3}e^{i𝐤(𝐱𝐱^{})}e^{i\omega (tt^{})}}`$ (A14) $`=`$ $`{\displaystyle \frac{Re}{2(2\pi )^{3+n}}}_i_j^{}{\displaystyle _0^{\mathrm{}}}k^{n1}e^{ik\mathrm{\Delta }t}𝑑k{\displaystyle _0^\pi }𝑑\theta _1\mathrm{sin}^{1+n}\theta _1e^{ikR\mathrm{cos}\theta _1}`$ (A16) $`\times {\displaystyle _0^\pi }d\theta _2\mathrm{sin}^n\theta _2\mathrm{}{\displaystyle _0^\pi }d\theta _{n+1}\mathrm{sin}\theta _{n+1}{\displaystyle _0^{2\pi }}d\theta _{n+2}`$ $`=`$ $`{\displaystyle \frac{a_nRe}{(2\pi )^{3+n}}}_i_j^{}{\displaystyle _0^{\mathrm{}}}k^{n1}e^{ik\mathrm{\Delta }t}𝑑k{\displaystyle _0^1}(1x^2)^{n/2}\mathrm{cos}(kRx)𝑑x`$ (A17) $`=`$ $`{\displaystyle \frac{a_n\sqrt{\pi }2^{\frac{n+1}{2}}\mathrm{\Gamma }(\frac{n}{2}+1)Re}{2(2\pi )^{3+n}}}_i_j^{}\left({\displaystyle \frac{1}{R^{\frac{n+1}{2}}}}{\displaystyle _0^{\mathrm{}}}k^{\frac{n3}{2}}J_{\frac{n+1}{2}}(kR)e^{ik\mathrm{\Delta }t}𝑑k\right)`$ (A18) $`=`$ $`{\displaystyle \frac{a_n\sqrt{\pi }2^{\frac{n+1}{2}}\mathrm{\Gamma }(\frac{n}{2}+1)}{2(2\pi )^{3+n}}}_i_j^{}\left({\displaystyle \frac{Re}{R^{\frac{n+1}{2}}}}{\displaystyle _0^{\mathrm{}}}k^{\frac{n3}{2}}J_{\frac{n+1}{2}}(kR)e^{ik\mathrm{\Delta }t}𝑑k\right)`$ (A19) $`=`$ $`{\displaystyle \frac{Re}{2(2\pi )^{\frac{3+n}{2}}}}_i_j^{}\left({\displaystyle \frac{1}{R^{\frac{n+1}{2}}}}{\displaystyle _0^{\mathrm{}}}k^{\frac{n3}{2}}J_{\frac{n+1}{2}}(kR)e^{ik\mathrm{\Delta }t}𝑑k\right)`$ (A20) $`=`$ $`{\displaystyle \frac{Re}{2(2\pi )^{\frac{3+n}{2}}}}_i_j^{}({\displaystyle \frac{n1}{R^2}}{\displaystyle \frac{1}{R^{\frac{n1}{2}}}}{\displaystyle _0^{\mathrm{}}}k^{\frac{n5}{2}}J_{\frac{n1}{2}}(kR)e^{ik\mathrm{\Delta }t}dk`$ (A22) $`{\displaystyle \frac{1}{R^{\frac{n+1}{2}}}}{\displaystyle _0^{\mathrm{}}}k^{\frac{n3}{2}}J_{\frac{n3}{2}}(kR)e^{ik\mathrm{\Delta }t}dk),`$ where we have utilized a recursive formula for Bessel functions $$zJ_{\nu 1}(z)+zJ_{\nu +1}(z)=2\nu J_\nu (z).$$ (A23) Similarly, one finds that $`H_{ijkl}^n(x,x^{})`$ $`={\displaystyle \frac{Re}{(2\pi )^{3+n}}}{\displaystyle d^{3+n}𝐤\frac{k_ik_jk_kk_l}{2\omega ^5}e^{i𝐤(𝐱𝐱^{})}e^{i\omega (tt^{})}}`$ (A26) $`={\displaystyle \frac{Re}{2(2\pi )^{\frac{3+n}{2}}}}_i_j^{}_k_l({\displaystyle \frac{n1}{R^2}}{\displaystyle \frac{1}{R^{\frac{n1}{2}}}}{\displaystyle _0^{\mathrm{}}}k^{\frac{n9}{2}}J_{\frac{n1}{2}}(kR)e^{ik\mathrm{\Delta }t}dk`$ $`{\displaystyle \frac{1}{R^2}}{\displaystyle \frac{1}{R^{\frac{n+1}{2}}}}{\displaystyle _0^{\mathrm{}}}k^{\frac{n7}{2}}J_{\frac{n3}{2}}(kR)e^{ik\mathrm{\Delta }t}dk).`$ To proceed further with the calculation, we need to deal with the cases when $`n`$ is odd or even separately. ### A The case of odd n Assume $`n=2m+1`$ and define $$S(m)=\frac{Re}{R^{m+1}}_0^{\mathrm{}}k^{m1}J_{m+1}(kR)e^{ik\mathrm{\Delta }t}𝑑k,m0,$$ (A27) $`T(m1)`$ $`=`$ $`{\displaystyle \frac{Re}{R^{m+1}}}{\displaystyle _0^{\mathrm{}}}k^{m1}J_{m1}(kR)e^{ik\mathrm{\Delta }t}𝑑k`$ (A28) $`=`$ $`{\displaystyle \frac{Re}{R^{m+1}}}\underset{\alpha 0^++i\mathrm{\Delta }t}{lim}{\displaystyle _0^{\mathrm{}}}k^{m1}J_{m1}(kR)e^{alphak}𝑑k`$ (A29) $`=`$ $`{\displaystyle \frac{2^{m1}\mathrm{\Gamma }(m1/2)}{\sqrt{\pi }}}{\displaystyle \frac{\sqrt{R^2\mathrm{\Delta }t^2}}{R^2(R^2\mathrm{\Delta }t^2)^m}},`$ (A30) $`=`$ $`{\displaystyle \frac{(2m1)!!}{(2m1)}}{\displaystyle \frac{\sqrt{R^2\mathrm{\Delta }t^2}}{R^2(R^2\mathrm{\Delta }t^2)^m}},m1,`$ (A31) where we have appealed to integral (6.623.1) in Ref. .The above result holds for $`R^2>\mathrm{\Delta }t^2`$, and $`T(m1)`$ is zero when $`R^2<\mathrm{\Delta }t^2`$. Then it follows from Eq (A22) that $$F_{ij}^{2m+1}=\frac{1}{2(2\pi )^{m+2}}_i_j^{}\left(S(m)\right),$$ (A32) and $$S(m)=\frac{2m}{R^2}S(m1)T(m1)$$ (A33) Using the recursive relation Eq (A33), we can show that $`S(m)`$ $`=`$ $`{\displaystyle \frac{(2m)!!}{R^{2m}}}S(0){\displaystyle \underset{k=1}{\overset{m}{}}}{\displaystyle \frac{(2m)!!}{(2k)!!}}{\displaystyle \frac{T(k1)}{R^{2m2k}}}`$ (A34) $`=`$ $`{\displaystyle \frac{(2m)!!}{R^{2m}}}S(0)\left[1{\displaystyle \underset{k=1}{\overset{m}{}}}{\displaystyle \frac{(2k1)!!}{(2k)!!(2k1)}}{\displaystyle \frac{R^{2k}}{(R^2\mathrm{\Delta }t^2)^k}}\right]`$ (A35) $`=`$ $`{\displaystyle \frac{(2m)!!}{R^{2m}}}S(0){\displaystyle \underset{k=0}{\overset{m}{}}}{\displaystyle \frac{(2k+1)!!}{(2k)!!(2k+1)(2k1)}}{\displaystyle \frac{R^{2k}}{(R^2\mathrm{\Delta }t^2)^k}}.`$ (A36) Here $$S(0)=\{\begin{array}{cc}\frac{\sqrt{R^2\mathrm{\Delta }t^2}}{R^2}\hfill & \text{ for}R^2>\mathrm{\Delta }t^2,\hfill \\ 0\hfill & \text{ for}R^2<\mathrm{\Delta }t^2,\hfill \end{array}$$ (A37) as reported in Ref. . For the sake of completeness, we give its derivation below $`S(0)`$ $`=`$ $`{\displaystyle \frac{Re}{R}}\underset{\alpha 0^++i\mathrm{\Delta }t}{lim}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{J_1(kR)}{k}}e^{\alpha k}𝑑k={\displaystyle \frac{1}{R}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{J_1(kR)\mathrm{cos}(k\mathrm{\Delta }t)}{k}}𝑑k`$ (A38) $`=`$ $`\{\begin{array}{cc}\frac{1}{R}\mathrm{cos}\left(\mathrm{arcsin}(\mathrm{\Delta }t/R)\right)=\frac{\sqrt{R^2\mathrm{\Delta }t^2}}{R^2}\hfill & \text{ for}R^2>\mathrm{\Delta }t^2,\hfill \\ 0\hfill & \text{ for}R^2<\mathrm{\Delta }t^2.\hfill \end{array}`$ (A41) If we define $$Q(m)=\frac{Re}{R^{m+1}}_0^{\mathrm{}}k^{m3}J_{m+1}(kR)e^{ik\mathrm{\Delta }t}𝑑k,m0,$$ (A42) then it is easy to see that $$H_{ijkl}^{2m+1}=\frac{1}{2(2\pi )^{m+2}}_i_j^{}_k_l^{}\left(Q(m)\right),$$ (A43) and $$Q(m)=\frac{2m}{R^2}Q(m1)\frac{1}{R^2}S(m2).$$ (A44) The above equation applies for $`m2`$. To use it to get a general expression, we need $`Q(0)`$, which is given by $`Q(0)`$ $`=`$ $`{\displaystyle \frac{1}{R}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{1}{k^3}}J_1(kR)\mathrm{cos}(k\mathrm{\Delta }t)𝑑k`$ (A45) $`=`$ $`\underset{\beta 0}{lim}{\displaystyle \frac{1}{R}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{k^1}{(k^2+\beta ^2)}}J_1(kR)\mathrm{cos}(k\mathrm{\Delta }t)𝑑k`$ (A46) $`=`$ $`\underset{\beta 0}{lim}{\displaystyle \frac{1}{R}}{\displaystyle \frac{e^{\beta \mathrm{\Delta }t}I_1(\beta R)}{\beta ^2}}={\displaystyle \frac{1}{2\beta }}{\displaystyle \frac{1}{2\mathrm{\Delta }t}}`$ (A47) This leads to a vanishing $`H_{ijkl}`$. We next need $`Q(1)`$, which can be calculated, using integral (6.693.5) in Ref. , as follows $`Q(1)`$ $`=`$ $`{\displaystyle \frac{1}{R^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{J_2(Rk)\mathrm{cos}(\mathrm{\Delta }tk)}{k^2}}𝑑k`$ (A49) $`=`$ $`\{\begin{array}{cc}\frac{1}{R^2}\left[\frac{R}{4}\mathrm{cos}(\mathrm{arcsin}(\mathrm{\Delta }t/R))+\frac{R}{12}\mathrm{cos}(3\mathrm{arcsin}(\mathrm{\Delta }t/R))\right]\hfill & \text{ for}R^2>\mathrm{\Delta }t^2,\hfill \\ 0\hfill & \text{ for}R^2<\mathrm{\Delta }t^2.\hfill \end{array}`$ (A52) $`=`$ $`\{\begin{array}{cc}\left(\frac{1}{3}\frac{\mathrm{\Delta }t^2}{3R^2}\right)\frac{\sqrt{R^2\mathrm{\Delta }t^2}}{R^2}=\left(\frac{1}{3}\frac{\mathrm{\Delta }t^2}{3R^2}\right)S(0)\hfill & \text{ for}R^2>\mathrm{\Delta }t^2,\hfill \\ 0\hfill & \text{ for}R^2<\mathrm{\Delta }t^2.\hfill \end{array}`$ (A55) In the above calculation, we have made use of the following trigonometric relations $$\mathrm{cos}(3x)=4\mathrm{cos}^3(x)3\mathrm{cos}(x),\mathrm{cos}(\mathrm{arcsin}x)=\sqrt{1x^2}.$$ (A57) Therefore one finds, using the recursive relation Eq (A44), $`Q(m)`$ $`=`$ $`{\displaystyle \frac{(2m)!!}{2R^{2m2}}}Q(1){\displaystyle \frac{1}{R^2}}{\displaystyle \underset{k=2}{\overset{m}{}}}{\displaystyle \frac{(2m)!!}{(2k)!!}}{\displaystyle \frac{S(k2)}{R^{2m2k}}}`$ (A58) $`=`$ $`{\displaystyle \frac{(2m)!!}{R^{2m2}}}[{\displaystyle \frac{1}{2}}Q(1)`$ (A60) $`+{\displaystyle \underset{k=2}{\overset{m}{}}}{\displaystyle \underset{j=0}{\overset{k2}{}}}{\displaystyle \frac{(2j+1)!!}{2k(2k2)(2j)!!(2j+1)(2j1)}}{\displaystyle \frac{R^{2j}}{(R^2\mathrm{\Delta }t^2)^j}}S(0)].`$ This expression can be simplified if we note that $$\underset{k=j+2}{\overset{m}{}}\frac{1}{k(k1)}=\underset{k=2}{\overset{m}{}}\frac{1}{k(k1)}\underset{k=2}{\overset{j+1}{}}\frac{1}{k(k1)}=\frac{mj1}{m(j+1)},$$ (A62) and $`{\displaystyle \underset{k=2}{\overset{m}{}}}{\displaystyle \underset{j=0}{\overset{k2}{}}}{\displaystyle \frac{(2j+1)!!}{2k(2k2)(2j)!!(2j+1)(2j1)}}{\displaystyle \frac{R^{2j}}{(R^2\mathrm{\Delta }t^2)^j}}S(0)]`$ (A63) $`={\displaystyle \underset{j=0}{\overset{m2}{}}}{\displaystyle \underset{k=j+2}{\overset{m}{}}}{\displaystyle \frac{(2j+1)!!}{2k(2k2)(2j)!!(2j+1)(2j1)}}{\displaystyle \frac{R^{2j}}{(R^2\mathrm{\Delta }t^2)^j}}S(0)]`$ (A64) $`={\displaystyle \underset{j=0}{\overset{m2}{}}}{\displaystyle \frac{(mj1)(2j+1)!!}{4m(j+1)(2j)!!(2j+1)(2j1)}}{\displaystyle \frac{R^{2j}}{(R^2\mathrm{\Delta }t^2)^j}}S(0)]`$ (A65) So, we have in this case $$D^{2m+1}=\{\begin{array}{cc}\frac{(2m+1)!!}{2(2\pi )^{m+2}}\frac{1}{(R^2\mathrm{\Delta }t^2)^{m+\frac{3}{2}}},\hfill & \text{ for}R^2>\mathrm{\Delta }t^2,\hfill \\ 0\hfill & \text{ for}R^2<\mathrm{\Delta }t^2,\hfill \end{array}$$ (A67) $$F_{ij}^{2m+1}=\frac{1}{2(2\pi )^{m+2}}_i_j^{}\left(\frac{(2m)!!}{R^{2m}}S(0)\underset{k=0}{\overset{m}{}}\frac{(2k+1)!!}{(2k)!!(2k+1)(2k1)}\frac{R^{2k}}{(R^2\mathrm{\Delta }t^2)^k}\right),$$ (A68) and $`H_{ijkl}^{2m+1}`$ $`=`$ $`{\displaystyle \frac{1}{2(2\pi )^{m+2}}}_i_j^{}_k_l^{}\{{\displaystyle \frac{(2m)!!}{R^{2m2}}}[{\displaystyle \frac{1}{2}}Q(1)`$ (A70) $`+{\displaystyle \underset{j=0}{\overset{m2}{}}}{\displaystyle \frac{(mj1)(2j+1)!!}{4m(j+1)(2j)!!(2j+1)(2j1)}}{\displaystyle \frac{R^{2j}}{(R^2\mathrm{\Delta }t^2)^j}}S(0)].`$ ### B The case of even n Let $`n=2m`$ with $`m=1,2,3\mathrm{}`$. The graviton two-point functions for $`m=0`$ corresponding to the usual 4 dimensional spacetime have been given previously . The analog of Eq. (A32) for this case is $$F_{ij}^{2m}=\frac{1}{2(2\pi )^{m+\frac{3}{2}}}_i_j^{}\left(S(m\frac{1}{2})\right).$$ (A72) Here $$S(m1/2)=\frac{2m1}{R^2}S(m\frac{3}{2})T(m\frac{3}{2}).$$ (A73) Using this recursive relation, we can express $`S(m1/2)`$ in terms of S(1/2) which is calculated, by employing $$J_{n+\frac{1}{2}}(z)=(1)^nz^{n+\frac{1}{2}}\sqrt{\frac{2}{\pi }}\frac{d^n}{(zdz)^n}\left(\frac{\mathrm{sin}z}{z}\right),$$ (A74) to be $`S(1/2)`$ $`={\displaystyle \frac{1}{R^{\frac{3}{2}}}}{\displaystyle _0^{\mathrm{}}}k^{1/2}J_{\frac{3}{2}}(Rk)\mathrm{cos}(\mathrm{\Delta }tk)𝑑k`$ (A78) $`=\sqrt{{\displaystyle \frac{2}{\pi }}}{\displaystyle \frac{1}{R^3}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{d}{dk}}\left({\displaystyle \frac{\mathrm{sin}(Rk)}{k}}\right)\mathrm{cos}(\mathrm{\Delta }tk)𝑑k`$ $`=\sqrt{{\displaystyle \frac{2}{\pi }}}\left({\displaystyle \frac{1}{R^3}}{\displaystyle \frac{\mathrm{sin}(Rk)\mathrm{cos}(\mathrm{\Delta }tk)}{k}}|_0^{\mathrm{}}+{\displaystyle \frac{\mathrm{\Delta }t}{R^3}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{\mathrm{sin}(Rk)\mathrm{sin}(\mathrm{\Delta }tk)}{k}}𝑑k\right)`$ $`=\sqrt{{\displaystyle \frac{2}{\pi }}}({\displaystyle \frac{1}{R^2}}{\displaystyle \frac{\mathrm{\Delta }t}{4R^3}}\mathrm{ln}\left({\displaystyle \frac{R+\mathrm{\Delta }t}{R\mathrm{\Delta }t}}\right)^2).`$ It then follows that $`S(m1/2)`$ $`={\displaystyle \frac{(2m1)!!}{R^{2m2}}}S(1/2){\displaystyle \underset{k=2}{\overset{m}{}}}{\displaystyle \frac{(2m1)!!}{(2k1)!!}}{\displaystyle \frac{T(k3/2)}{R^{2m2k}}}`$ (A80) $`={\displaystyle \frac{(2m1)!!}{R^{2m}}}\sqrt{{\displaystyle \frac{2}{\pi }}}[1{\displaystyle \frac{\mathrm{\Delta }t}{4R}}\mathrm{ln}\left({\displaystyle \frac{R\mathrm{\Delta }t}{R\mathrm{\Delta }t}}\right)^2{\displaystyle \frac{1}{R^2}}{\displaystyle \underset{k=2}{\overset{m}{}}}{\displaystyle \frac{2^{k2}\mathrm{\Gamma }(k1)}{(2k1)!!}}{\displaystyle \frac{R^{2k}}{(R^2\mathrm{\Delta }t^2)^{k1}}}].`$ Similarly, one has for $`H_{ijkl}^{2m}`$ $$H_{ijkl}^{2m}=\frac{1}{2(2\pi )^{m+\frac{3}{2}}}_i_j^{}_k_l^{}\left(Q(m\frac{1}{2})\right),$$ (A82) and $$Q(m1/2)=\frac{2m1}{R^2}Q(m\frac{3}{2})\frac{1}{R^2}S(m\frac{5}{2}).$$ (A83) Now the calculation becomes a little tricky. First, let us note that $`H_{ijkl}^0`$ has already been given and the recursive relation Eq (A83) can only be applied when $`m3`$. So, we need both $`H_{ijkl}^2`$ and $`H_{ijkl}^4`$ or $`Q(1/2)`$ and $`Q(3/2)`$ as our basis to use the recursive relation for a general expression. Because there is an infrared divergence in the $`Q(1/2)`$ integral, so, as we did in the 4 dimensional case, we will introduce a regulator $`\beta `$ in the denominator of the integrand and then let $`\beta `$ approach 0 after the integration is performed. Noting that $$J_{\frac{3}{2}}(z)=\sqrt{\frac{2}{\pi z}}\left(\frac{\mathrm{sin}z}{z}\mathrm{cos}z\right),$$ (A84) we obtain $`Q(1/2)`$ $`=`$ $`{\displaystyle \frac{1}{R^{\frac{3}{2}}}}{\displaystyle _0^{\mathrm{}}}k^{\frac{5}{2}}J_{\frac{3}{2}}(kR)\mathrm{cos}(kt)𝑑k`$ (A87) $`=\sqrt{{\displaystyle \frac{2}{\pi }}}\left({\displaystyle \frac{1}{R^3}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dk}{k^4}}\mathrm{sin}kR\mathrm{cos}k\mathrm{\Delta }t{\displaystyle \frac{1}{R^2}}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dk}{k^3}}\mathrm{cos}kR\mathrm{cos}k\mathrm{\Delta }t\right)`$ $`=\sqrt{{\displaystyle \frac{2}{\pi }}}\underset{\beta 0}{lim}({\displaystyle \frac{1}{R^3}}{\displaystyle \frac{1}{2\beta }}{\displaystyle \frac{}{\beta }}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{\mathrm{sin}kR\mathrm{cos}k\mathrm{\Delta }t}{k^2+\beta ^2}}dk`$ $`+{\displaystyle \frac{1}{R^2}}{\displaystyle \frac{1}{2\beta }}{\displaystyle \frac{}{\beta }}{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{k\mathrm{cos}kR\mathrm{cos}k\mathrm{\Delta }t}{k^2+\beta ^2}}dk).`$ (A88) We next use $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{\mathrm{sin}(ax)\mathrm{cos}(bx)}{\beta ^2+x^2}}𝑑x=`$ $`{\displaystyle \frac{1}{4\beta }}e^{a\beta }\{e^{b\beta }Ei[\beta (ab)]+e^{b\beta }Ei[\beta (a+b)]\}`$ (A91) $`{\displaystyle \frac{1}{4\beta }}e^{a\beta }\{e^{b\beta }Ei[\beta (a+b)]+e^{b\beta }Ei[\beta (ab)]\},`$ $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{x\mathrm{cos}(ax)\mathrm{cos}(bx)}{\beta ^2+x^2}}𝑑x=`$ $`{\displaystyle \frac{1}{4}}e^{a\beta }\{e^{b\beta }Ei[\beta (ab)]+e^{b\beta }Ei[\beta (a+b)]\}`$ (A93) $`{\displaystyle \frac{1}{4}}e^{a\beta }\{e^{b\beta }Ei[\beta (a+b)]+e^{b\beta }Ei[\beta (ab)]\},`$ where $`\mathrm{Ei}(\mathrm{x})`$ is the exponential-integral function, and the fact that, when $`x`$ is small, $$\mathrm{Ei}(\mathrm{x})\gamma +\mathrm{ln}|x|+x+\frac{1}{4}x^2+\frac{1}{18}x^3+O(x^4),$$ (A94) where $`\gamma `$ is the Euler constant. After expanding $`Q(1/2)`$ around $`\beta =0`$ to the order of $`\beta ^2`$, one finds $`Q(1/2)`$ $`=`$ $`\underset{\beta 0}{lim}\sqrt{{\displaystyle \frac{2}{\pi }}}({\displaystyle \frac{5}{18}}{\displaystyle \frac{1}{3}}\gamma {\displaystyle \frac{1}{3}}\mathrm{ln}(\beta ){\displaystyle \frac{1}{6}}\mathrm{ln}(R^2\mathrm{\Delta }t^2)`$ (A96) $`{\displaystyle \frac{\mathrm{\Delta }t^2}{6R^2}}+{\displaystyle \frac{\mathrm{\Delta }t}{8R}}({\displaystyle \frac{\mathrm{\Delta }t^2}{3R^2}}1\left)\mathrm{ln}\left({\displaystyle \frac{R+\mathrm{\Delta }t}{R\mathrm{\Delta }t}}\right)^2\right).`$ Note, however, that what we need is $`H_{ijkl}`$ which involves differentiation of $`Q(1/2)`$, therefore we can discard the constant and divergent terms in $`Q(1/2)`$ as far as $`H_{ijkl}`$ is concerned. To calculate $`Q(3/2)`$, let us recall that $$Q(3/2)=\frac{3}{R^2}Q(1/2)\frac{1}{R^2}S(1/2)$$ (A97) and note that $`S(1/2)`$ is given by $`\sqrt{2/\pi }`$ times Eq (A19) in Ref. Thus, we have $$Q(3/2)=\sqrt{\frac{2}{\pi }}(\frac{1}{6R^2}\frac{\mathrm{\Delta }t^2}{2R^4}+\frac{\mathrm{\Delta }t}{8R^3}(\frac{\mathrm{\Delta }t^2}{R^2}1\left)\mathrm{ln}\left(\frac{R+\mathrm{\Delta }t}{R\mathrm{\Delta }t}\right)^2\right).$$ (A98) With $`Q(3/2)`$ at hand, it is easy to show that for an arbitrary $`m3`$ $`Q(m1/2)`$ $`=`$ $`{\displaystyle \frac{(2m1)!!}{3R^{2m4}}}Q(3/2){\displaystyle \frac{1}{R^2}}{\displaystyle \underset{k=3}{\overset{m}{}}}{\displaystyle \frac{(2m1)!!}{(2k1)!!}}{\displaystyle \frac{S(k5/2)}{R^{2m2k}}}`$ (A99) $`=`$ $`{\displaystyle \frac{(2m1)!!}{R^{2m4}}}({\displaystyle \frac{1}{3}}Q(3/2){\displaystyle \underset{k=3}{\overset{m}{}}}{\displaystyle \frac{1}{(2k1)(2k3)}}S(1/2)`$ (A101) $`+{\displaystyle \frac{1}{R^4}}\sqrt{{\displaystyle \frac{2}{\pi }}}{\displaystyle \underset{k=3}{\overset{m}{}}}{\displaystyle \underset{j=2}{\overset{k2}{}}}{\displaystyle \frac{2^{j2}\mathrm{\Gamma }(j1)}{(2j1)!!(2k1)(2k3)}}{\displaystyle \frac{R^{2j}}{(R^2\mathrm{\Delta }t^2)^{j1}}})`$ $`=`$ $`{\displaystyle \frac{(2m1)!!}{R^{2m4}}}({\displaystyle \frac{1}{3}}Q(3/2){\displaystyle \underset{k=3}{\overset{m}{}}}{\displaystyle \frac{1}{(2k1)(2k3)}}S(1/2)`$ (A103) $`+{\displaystyle \frac{1}{R^4}}\sqrt{{\displaystyle \frac{2}{\pi }}}{\displaystyle \underset{j=2}{\overset{m2}{}}}{\displaystyle \frac{(mj1)2^{j2}\mathrm{\Gamma }(j1)}{(2m1)(2j+1)!!}}{\displaystyle \frac{R^{2j}}{(R^2\mathrm{\Delta }t^2)^{j1}}}).`$ Here in the last step, we have made use of the following results $`{\displaystyle \underset{k=j+2}{\overset{m}{}}}{\displaystyle \frac{1}{(2k1)(2k3)}}`$ $`=`$ $`{\displaystyle \underset{k=2}{\overset{m}{}}}{\displaystyle \frac{1}{(2k1)(2k3)}}{\displaystyle \underset{k=2}{\overset{j+1}{}}}{\displaystyle \frac{1}{(2k1)(2k3)}}`$ (A105) $`=`$ $`{\displaystyle \frac{mj1}{(2m1)(2j+1)}},`$ (A106) and $$\underset{k=3}{\overset{m}{}}\underset{j=2}{\overset{k2}{}}f(j)g(k)=\underset{k=4}{\overset{m}{}}\underset{j=2}{\overset{k2}{}}f(j)g(k)=\underset{j=2}{\overset{m2}{}}f(j)\underset{k=2+j}{\overset{m}{}}g(k).$$ (A107) Consequently, we obtain $$D^{2m}=\frac{2^mm!}{(2\pi )^{m+2}}\frac{1}{(R^2\mathrm{\Delta }t^2)^{m+1}},$$ (A108) $`F_{ij}^{2m}`$ $`=`$ $`{\displaystyle \frac{1}{(2\pi )^{m+2}}}_i_j^{}\{{\displaystyle \frac{(2m1)!!}{R^{2m}}}[1{\displaystyle \frac{\mathrm{\Delta }t}{4R}}\mathrm{ln}\left({\displaystyle \frac{R\mathrm{\Delta }t}{R\mathrm{\Delta }t}}\right)^2`$ (A110) $`{\displaystyle \frac{1}{R^2}}{\displaystyle \underset{k=2}{\overset{m}{}}}{\displaystyle \frac{2^{k2}\mathrm{\Gamma }(k1)}{(2k1)!!}}{\displaystyle \frac{R^{2k}}{(R^2\mathrm{\Delta }t^2)^{k1}}}]\},`$ and $`H_{ijkl}^{2m}`$ $`=`$ $`{\displaystyle \frac{1}{(2\pi )^{m+2}}}_i_j^{}_k_l^{}\{{\displaystyle \frac{(2m1)!!}{R^{2m4}}}[{\displaystyle \frac{\mathrm{\Delta }t}{24R^3}}({\displaystyle \frac{\mathrm{\Delta }t^2}{R^2}}1)\mathrm{ln}\left({\displaystyle \frac{R+\mathrm{\Delta }t}{R\mathrm{\Delta }t}}\right)^2{\displaystyle \frac{1}{18R^2}}`$ (A113) $`{\displaystyle \frac{\mathrm{\Delta }t^2}{6R^4}}{\displaystyle \underset{k=3}{\overset{m}{}}}{\displaystyle \frac{1}{(2k1)(2k3)}}({\displaystyle \frac{1}{R^2}}{\displaystyle \frac{\mathrm{\Delta }t}{4R^3}}\mathrm{ln}\left({\displaystyle \frac{R+\mathrm{\Delta }t}{R\mathrm{\Delta }t}}\right)^2)`$ $`+{\displaystyle \frac{1}{R^4}}{\displaystyle \underset{j=2}{\overset{m2}{}}}{\displaystyle \frac{(mj1)2^{j2}\mathrm{\Gamma }(j1)}{(2m1)(2j+1)!!}}{\displaystyle \frac{R^{2j}}{(R^2\mathrm{\Delta }t^2)^{j1}}}).`$
warning/0004/cond-mat0004034.html
ar5iv
text
# Novel type of orbital ordering: complex orbitals in doped Mott insulators ## Abstract An orbital ordering, often observed in Mott insulators with orbital degeneracy, is usually supposed to disappear with doping, e.g. in the ferromagnetic metallic phase of manganites. We propose that the orbital ordering of a novel type may exist in such situation: there may occur ferro. orbital ordering of complex orbitals (linear superposition of basic orbitals $`d_{x^2y^2}`$ and $`d_{z^2}`$ with complex coefficients). Despite the perfect orbital ordering, such state still retains cubic symmetry and thus would not induce any structural distortion. This novel state can resolve many problems in the physics of CMR manganites and can also exist in other doped Mott insulators with Jahn-Teller ions. Mott insulators often display various types of ordering connected with localized electrons. In particular, in the presence of an orbital degeneracy, very often an orbital ordering occurs besides the magnetic (spin) one. Classical examples are the systems with doubly-degenerate $`e_g`$-electrons, e.g. KCuF<sub>3</sub> or MnF<sub>3</sub> . At present special attention is paid to the manganites La$`{}_{1x}{}^{}M_{x}^{}`$MnO<sub>3</sub> ($`M=\mathrm{Ca},\mathrm{Sr}`$) with the colossal magnetoresistance (CMR). The undoped material LaMnO<sub>3</sub> is a Mott insulator with the orbital ordering. What becomes of this ordering in doped materials and which role the orbital degrees of freedom do play in the CMR phase, is a matter of active investigation and of hot debate. On the one hand it is usually assumed that the orbital ordering (at least the long-range one) disappears in a ferromagnetic metallic (FM) phase ($`0.16<x<0.5`$) displaying CMR. This is in particular concluded from the disappearance of lattice distortion typical for the orbitally ordered states. On the other hand the attempts to explain transport properties in this regime using only electron–spin scattering in the framework of the standard double-exchange model failed , which forced Millis and coworker to suggest that the orbital degrees of freedom are still important in this regime, e.g. producing Jahn-Teller polarons. Nevertheless the orbitals were still assumed to be disordered in the FM phase below $`T_c`$. Until now in all the numerous studies of the cooperative Jahn-Teller effect and orbital ordering only one class of solutions was considered: occupied orbitals were always assumed to be certain linear combinations of the basic orbitals $`d_{z^2}\frac{1}{\sqrt{6}}(2z^2x^2y^2)`$ and $`d_{x^2y^2}\frac{1}{\sqrt{2}}(x^2y^2)`$ of the form $$|\theta =\mathrm{cos}\frac{\theta }{2}|z^2+\mathrm{sin}\frac{\theta }{2}|x^2y^2,$$ (1) i.e. linear superposition with real coefficients. Such orbitals give an asymmetric (quadrupolar) distortion of the electron density at a given site. Consequently, any ordering involving orbitals of the type (1), be it of “ferro” or “antiferro” type, corresponds to a decrease of the symmetry of charge distribution and is accompanied (or driven) by the corresponding lattice distortion. There exists however yet another possibility: there may exist orbital ordering without any lattice distortion. From the basis set of doubly-degenerate orbitals we can also form a linear superposition of the type (1) but with complex coefficients. The simplest example is the normalized functions $$|\pm =\frac{1}{\sqrt{2}}\left(|z^2\pm i|x^2y^2\right).$$ (2) We can form an orbital ordering with the occupied orbitals of such type, e.g. the ferro. orbital ordering (FOO) with the orbitals $`|+`$ at each site. It can be easily checked that the electron charge distribution in this state is perfectly symmetric, it is the same along $`x`$, $`y`$ and $`z`$-directions, so that the net symmetry of a crystal would remain cubic. Correspondingly, the effective electron hopping matrix elements are isotropic, $`t^x=t^y=t^z=t^{}`$, as well as superexchange interaction $`J^x=J^y=J^z=2t^2/U`$. (Here we have in mind the traditional description of the strongly interacting electrons by the Hubbard-type model which in case of orbital degeneracy is actually a degenerate Hubbard model with the on-site repulsion $`U`$ and electron hopping between different orbitals, to be specified below.) We suggest that just such an ordering is realized in the FM phase of CMR manganites. If true, this would imply that both the spin and orbital degeneracies are lifted in the ground state in these systems which consequently are ferromagnetic in both the spin and orbital variables (but with “strange” orbitals (2)). Why have such states not been considered before? Apparently it is connected with two factors. On one hand, in most real materials with Jahn-Teller ions studied up to now there always occurred orbital ordering accompanied by lattice distortion. On the other hand, in the theoretical description of these phenomena only the operators corresponding to an ordering of conventional type appear in the effective Hamiltonian (see below), so that it seemed that other possibilities are never realized in practice . Most probably, indeed only the “real” solutions (1) can be realized in stoichiometric compounds with Jahn-Teller ions. The situation however may be different in doped materials — and that is what we propose here. It is well known that the electron (or hole) motion in systems with strongly correlated electrons is hindered by an antiferromagnetic background and is much facilitated if the background ordering becomes ferromagnetic (see e.g. ): in nondegenerate Hubbard model it gives rise to Nagaoka’s ferromagnetism , and in systems with two types of electrons — to ferromagnetism due to double exchange. The energy gain in such a ferro. state as compared to the antiferro. or paramagnetic (disordered) one is of the order of the electron bandwidth, or of the hopping matrix element $`t`$ per doped charge carrier: $`\mathrm{\Delta }E=ctx`$ ($`c`$ is some constant of order 1), whereas the energy loss is of the order of $`J`$ $`(t^2/U)(1x)`$. Thus, as argued already in , the saturated ferromagnetic state is realized for $`x>x_cJ/tt/U`$. For smaller $`x`$ the inhomogeneous phase separated state with ferromagnetic droplets in an antiferomagnetic matrix can be realized . In exact analogy to this case we should expect that due to the same mechanism a ferro. orbital ordering will be established in doped strongly interacting Mott insulators with orbital degeneracy. Which particular orbitals would be stabilized at that, may depend on the particular situation; we suggest that it may be the complex orbitals of the type (2). A convenient way to describe the orbital ordering of double-degenerate $`e_g`$-orbitals is to introduce pseudospin variables $`\tau _i`$ such that e.g. the orbital $`d_{z^2}`$ corresponds to $`\tau ^z=+\frac{1}{2}`$, and the orthogonal orbital $`d_{x^2y^2}`$ — to $`\tau ^z=\frac{1}{2}`$. The orbital states (1) considered until now are parametrized by the angle $`\theta `$ in the $`(\tau ^z,\tau ^x)`$-plane. Cubic symmetry is reflected in the $`\theta `$-plane in $`\frac{2\pi }{3}`$ symmetry: the state $`\theta =\frac{2\pi }{3}`$ corresponds to the orbital $`|x^2(2x^2y^2z^2)`$ and $`\theta =\frac{2\pi }{3}`$ — to $`|y^2(2y^2x^2z^2)`$; these are more or less the orbitals occupied in two sublattices in the undoped LaMnO<sub>3</sub>. When we dope the system with orbital ordering, the motion of the charge carriers would initially (for small $`x`$) occur on a background of this orbital ordering. Similar to the hole motion on an antiferromagnetic spin background, the “antiferro” orbital ordering would hinder the motion of a hole and would reduce its bandwidth. One can indeed check that, by making the orbital order ferromagnetic, e.g. occupying the same orbital at each site, for instance $`|z^2`$ or $`|z^2y^2`$, we would increase the bandwidth and correspondingly decrease kinetic energy of holes; the mathematical details are presented below. However such a FOO seems to contradict experimental observations. Thus e.g. $`|z^2`$-ferromagnetism would lead to a tetragonal distortion of the lattice with $`c/a>1`$ and to a strong anisotropy of transport properties; but nothing like that is observed experimentally. Probably that is why this possibility, which by analogy with the spin case seems quite natural, is never considered for CMR phase of manganites, and it is usually assumed that the orbital ordering is simply lost in the FM phase leading e.g. to an orbital liquid However if at such a FOO the complex orbitals (2) would be occupied — there would be no contradiction with the structural data, and still the motion of doped charge carriers would be unhindered. This is the main idea of the present paper. In the pseudospin language introduced above the state (2) is an eigenstate of the operator $`\tau ^y`$ which also exists in the algebra of $`\tau =\frac{1}{2}`$ operators. For strongly interacting electrons with one electron per site one can describe the effective spin and orbital interaction by the Hamiltonian which schematically has the form $$=J_s(\stackrel{}{s}_i\stackrel{}{s}_j)+J_\tau (\tau _i\tau _j)+J_{s\tau }(\stackrel{}{s}_i\stackrel{}{s}_j)(\tau _i\tau _j),$$ (3) where the first and the third terms are due to superexchange, $`J_{exch}J_sJ_{s\tau }t^2/U`$, and in the second term also the Jahn-Teller interaction of degenerate orbitals with the lattice contribute, $`J_\tau =J_{exch}+J_{\mathrm{JT}}`$ The interaction (3) is in general anisotropic with respect to $`\tau `$-operators and usually it does not contain terms with $`\tau ^y`$. Consequently $`\tau ^y`$-states do not appear in the mean-field approximation which is nearly always used to treat insulators with orbital degeneracy. But it does not mean that such states cannot appear in certain situations, e.g. for doped systems. Using the specific form of the $`e_g`$-orbitals and of the corresponding hopping integrals between different orbitals in different directions, one can easily calculate the one-electron band structure for different types of orbital ordering. Denoting the hopping between $`z^2`$-orbitals for a pair along $`z`$-direction as $`t`$, we have: $`t_{z^2,z^2}^z=t`$, $`t_{z^2,z^2}^{x,y}=t/4`$, $`t_{x^2y^2,x^2y^2}^{x,y}=3t/4`$, $`t_{z^2,x^2y^2}^{x,y}=\pm \sqrt{3}t/4`$, $`t_{z^2,x^2}^x=t_{x^2,y^2}^{x,y}=t/2`$, etc., see e.g. Generally speaking, in contrast to spin case (see for discussion and references), an antiferro. orbital ordering, e.g. that in LaMnO<sub>3</sub>, does not completely suppress the hole motion even if one ignores quantum effects: the hole can always hop to a neighbouring site without destroying orbital order along its trajectory. This is connected with the fact that the nondiagonal hopping in orbital channel is in general allowed, and pseudospin projection $`\tau ^z`$ is not conserved during hopping: the nondiagonal hopping integrals $`t_{z^2,x^2y^2}`$ are in general nonzero. Thus the hole can here move without leaving a “trace” of wrong spins and, consequently, there will be no “confinement”. However the bandwidth of this coherent motion without disturbing the background orbital ordering is reduced as compared to the bandwidth for FOO. Thus, using the values of hopping integrals $`t_{\alpha ,\beta }^{ij}`$ given above, one obtains for the undoped LaMnO<sub>3</sub> (alternation of $`x^2`$ and $`y^2`$ orbitals) the spectrum $$\epsilon (k)=2\left[\frac{1}{2}t(\mathrm{cos}k_x+\mathrm{cos}k_y)+\frac{1}{4}t\mathrm{cos}k_z\right]$$ (4) so that the minimum energy of the hole (the bottom of the band) will be $`\epsilon _{\mathrm{𝑚𝑖𝑛}}=\epsilon (k=0)=2.5t`$. If however we would make a FOO, e.g. occupying at each site $`z^2`$-orbital, the spectrum will be $$\epsilon ^{z^2\text{-ferro}}(k)=2t\left[\frac{1}{4}(\mathrm{cos}k_x+\mathrm{cos}k_y)+\mathrm{cos}k_z\right]$$ (5) and $`\epsilon _{\mathrm{𝑚𝑖𝑛}}^{z^2\text{-ferro}}=3t`$. Thus we indeed see that the energy gain in the FOO state is of order $`t`$. The same minimum energy is reached also for $`(x^2y^2)`$-ferro. ordering and for FOO with any orbital of the type (1). From (1) one can easily obtain that $$t_{\theta ,\theta }^z=\theta |\widehat{t}^z|\theta =t\mathrm{cos}^2\frac{\theta }{2};t_{\theta ,\theta }^{x/y}=\frac{t}{4}(\mathrm{cos}\frac{\theta }{2}\pm \sqrt{3}\mathrm{sin}\frac{\theta }{2})^2,$$ (6) And one obtains that the bottom of the spectrum with the hopping integrals given by (6) does not depend on $`\theta `$ and coincides with the value $`3t`$. But exactly the same $`\epsilon _{\mathrm{𝑚𝑖𝑛}}`$ is also reached for the $`\tau ^y`$-ferromagnetism, where the state (2) is occupied at each site. Using the values of $`t`$’s presented above, one obtains that $`t_{\tau ^y,\tau ^y}=\frac{1}{2}t`$ in all three directions, and $$\epsilon ^{\tau ^y\text{-ferro}}(k)=t(\mathrm{cos}k_x+\mathrm{cos}k_y+\mathrm{cos}k_z)$$ (7) so that the spectrum is indeed isotropic, and the minimum energy is also equal to $`3t`$. Note that, similarly to the nondegenerate Hubbard model, we can use the simple dispersion relations (4), (5) only for the electron motion which does not destroy the background ordering. The energy gain in the FOO state is again, similarly to the Nagaoka case , $`tx`$, and the energy loss is $`J`$ (in orbital sector $`J_\tau +J_{s\tau }`$). Here again we obtain the totally ferromagnetic state both in spins and in orbitals for $`x>x_cJ/t`$, and one can have phase separation for $`x<x_c`$. As we saw above, for small $`x`$ different FOO states are equivalent as to the gain in kinetic energy. On the other hand we can see that $`\tau ^y`$-ferro. state is favourable as to the loss of the effective exchange energy (3). Indeed, the effective Hamiltonian (3) contains the orbital operators in combinations $`\tau _i^z\tau _j^z`$ , $`\tau _i^x\tau _j^x`$ and $`\tau _i^z\tau _j^x`$ with positive (antiferromagnetic) coeficients $`J_\tau `$, $`J_{s\tau }`$ . In the ground state of the undoped system a certain spin and orbital order is realised which minimizes the total energy — and it will be an antiferro. ordering of real orbitals (1). If however we force our system to be ferromagnetic both in spins and in orbitals, e.g., as argued above, by doping, we strongly increase this exchange energy — and we can minimize this energy loss by chosing FOO of $`\tau ^y`$-type. One can check directly that the superexchange part of the energy is indeed smaller in the $`\tau ^y`$-FOO state as compared e.g. with the FOO state of real orbitals. Thus, taking the standard expression for the exchange integrals $`J=2t^2/U`$ and calculating exchange constants with the proper values of $`t`$, we obtain e.g. that for the $`z^2`$-FOO the energy of the ferromagnetic state (per site) in mean-field approximation can be written in the form: $$E_{\mathrm{𝑓𝑒𝑟𝑟𝑜}}^{z^2\text{-FOO}}=\frac{9}{4}\frac{t^2}{U}+C,$$ (8) where $`C=3\frac{t^2}{U}`$. The same value of $`E_{\mathrm{𝑓𝑒𝑟𝑟𝑜}}`$ we would obtain for $`(x^2y^2)`$-FOO and for all other FOO states with arbitrary state $`|\theta `$ of the type (1): again one can easily show that with the values of hopping integrals $`t_{\theta \theta }^\alpha `$ (6) the magnetic energy of both ferro- and antiferromagnetic states does not depend on $`\theta `$ and is given by (8). Similar treatment of the $`\tau ^y`$-FOO state shows however that the net exchange energy in this state is lower: $$E_{\mathrm{𝑓𝑒𝑟𝑟𝑜}}^{\tau ^y\text{-FOO}}=\frac{3}{2}\frac{t^2}{U}+C,$$ (9) where $`C`$ is the same as in Eq.(8). Similarly, the intersite Jahn-Teller energy in FOO state with real orbitals is positive, $`J_{\mathrm{JT}}`$, whereas it is zero for the $`\tau ^y`$-FOO state; again from this point of view the $`\tau ^y`$-FOO state wins against all other possible FOO states with real orbitals. As the electronic energy of doped charge carriers coincides for the states (1) and (2), and the magnetic energy of the latter is lower, one can conclude that the proposed ferromagnetic state with $`\tau ^y`$-FOO (occupation of complex orbitals (2)) will be the best among all possible ferro. ordered states — and these, we argue, may be stabilized by the usual Nagaoka mechanism. This state can possibly be realized in the ferromagnetic metallic phase of CMR manganites. The finite band filling would increase electron kinetic energy somewhat faster for $`\tau ^y`$-ordering as compared e.g. to $`z^2`$-FOO or $`(x^2y^2)`$-FOO (the energy spectrum in the latter cases is anisotropic and the density of states at the band edge is higher, cf. ). Consequently the $`\tau ^y`$-FOO may be destabilized at still higher doping levels. The state with $`\tau ^y`$-FOO need not be necessarily homogeneous: as in the pure spin case, for small doping there may exist a tendency towards phase separation due to an instability of the homogeneous canted spin state . But, as follows from the arguments given above, the structure of the hole-rich regions can well be of $`\tau ^y`$-FOO type. ¿From the point of view of symmetry the proposed $`\tau ^y`$-FOO corresponds to the $`A_{2g}`$ (or $`\mathrm{\Gamma }_2`$)-representation of the cubic group . The corresponding order parameter is $$T_{xyz}=𝒮l_xl_yl_z,$$ (10) where $``$ is the ground state average, $`𝒮`$ denotes symmetrization and $`l_x`$, $`l_y`$, $`l_z`$ are the corresponding components of the momentum operator $`\widehat{l}=2`$. This follows from the group symmetry, and we checked by direct calculation that it is the lowest nonzero average in our state. One can show that in the basis $`|z^2`$, $`|x^2y^2`$ $`T_{xyz}`$ is indeed proportional to the $`\tau ^y`$-Pauli matrix. From eq. (10) we see that the order parameter in our state is a magnetic octopole, cf. . It is in principle possible, although not easy, to observe such ordering experimentally. Probably the most promising would be the resonance experiments like NMR or circular X-Ray dichroism. One can also show that this state should have a piezomagnetism. It is not apriori clear whether this $`\tau ^y`$-FOO and the ferromagnetic spin ordering would occur at the same temperature. The direct product of corresponding irreducible representations $`A_2`$ and $`T_1`$ does not contain unit representation $`A_1`$ , so that there will be in general no linear coupling between these order parameters. However if the rhombohedral distortion $`T_{2g}`$, present in manganites at high temperatures (S. W. Cheong, private communication), would still exist at low temperatures, then there will exist linear coupling of these three types of ordering ($`A_2\times T_1\times T_2A_1`$) and the $`\tau ^y`$-FOO and the ferromagnetic ordering will occur at the same $`T_c`$. This could help to resolve the problem pointed out in Ref. , that the change of resistivity through $`T_c`$ in purely spin case is too small: one would get an extra scattering in a “para” phase — an orbital-disorder scattering — in addition to the spin-disorder one. (Alternatively one can speak of the increase of resistivity above $`T_c`$ not due to orbital fluctuations, but, again in analogy with the spin case , due to the effective narrowing of the band in the orbitally disordered phase.) Summarizing, we suggest in this paper that, similarly to magnetic ordering, doping of orbitally degenerate Mott insulators containing Jahn-Teller ions could stabilize the ferro. orbital ordering of special type, the occupied orbitals being “complex” — i.e. the linear superpositions of basic orbitals $`d_{x^2y^2}`$ and $`d_{z^2}`$ with complex coefficients. Such orbitals, despite perfect ordering, do not induce any structural distortion. At the same time the motion of charge carriers on such ordered background is completely free, and there is no extra band narrowing; it is just this factor that favours such ferro. orbital ordering. Thus the origin of the ferro. orbital ordering in doped systems is the same as that of the Nagaoka ferromagnetism in a partially-filled Hubbard model. The ferro. ordering of complex orbitals gives the same band energy as the ordering of conventional real ones but has lower exchange and Jahn-Teller energy; therefore the ferro. ordering of complex orbitals may be preferable. This idea may explain the main properties of the colossal magnetoresistance manganites in the most interesting ferromagnetic metallic concentration range: in this picture they are perfectly ordered with respect to both spin and orbitals, but with the complex orbitals occupied at each site. In particular, this could help to explain the sharp drop of resistivity below $`T_c`$, with the crystal structure remaining undistorted. Similar states may in principle exist also in other doped Mott insulators with Jahn-Teller ions which constitute a large interesting class of magnetic materials with very specific properties. I am grateful to G. A. Sawatzky and M. V. Mostovoy for useful discussions. I am also grateful do D. Cox who pointed out a possible importance of the rhombohedral distortion in manganites. This work was supported by the Netherlands Foundation FOM, by the European Network OXSEN and by the project INTAS–97 0963.
warning/0004/hep-lat0004019.html
ar5iv
text
# The 𝜋₀→𝛾⁢𝛾 decay and the chiral anomaly in the quark-composites approach to QCD ## 1 Introduction Unlike previous field theories, QCD does not contain the fields which describe the particles observed in the experiments. This motivated the introduction of effective Lagrangians like the chiral Lagrangians, written in terms of the phenomenological fields and based solely on general principles of invariance. The chiral Lagrangians encode the spontaneous breaking of the chiral symmetry which is generally believed to occur in QCD and the old results of current algebra and PCAC . But while these descriptions have been put on the safe ground of a consistent and predictive field theory, the determination from QCD of the parameters appearing in them as well as the proof that chiral symmetry is spontaneously broken in this theory are still lacking. The foundations of the chiral Lagrangians could, in principle, be completed by the use of numerical simulations in the lattice formulation of QCD, but, of course, an analytical approach is desirable also for a better understanding and a more clever way of using the numerical recipes. There have been, indeed, attempts in this direction, but they have been so far restricted to the strong coupling region in the gauge coupling constant , which is unfortunately far from the interesting continuum limit. It has also been proposed to extend QCD via the introduction of extra degrees of freedom . The corresponding fields, even though carry the quantum numbers of the chiral mesons, are supposed to decouple in the continuum limit, in order to avoid double counting. The idea behind the quark composites approach is that it should be possible to recover the interactions of the phenomenological fields by a change of variables in the partition function of QCD whereby the quark composites with the quantum numbers of the phenomenological fields of interest are assumed as new integration variables. The final goal is to unify the description of the spectrum properties and scattering processes in a framework consistent with the confinement of quarks. For technical reasons this program can be realized in this form only for the baryons . For the mesons instead of a change of variables we make recourse to auxiliary fields . This latter procedure, however, should not be confused with the quoted use of additional phenomenological fields in QCD. In practice, to implement our approach in a perturbative framework, is also necessary that, after our manipulations, the free actions of the composites emerge in the effective action. To this aim we make use of the arbitrariness in the definition of the regularized action and perform a suitable choice of irrelevant terms which help to construct the effective action of the composites. In our previous work we got a proof of the spontaneous breaking of the chiral symmetry in QCD in the framework of our perturbative approach. Exploratory applications of this approach have been performed in the study of the pion-nucleon interaction and the high temperature QCD phase transition . These works are limited by the fact that in the broken phase the chiral symmetry is not realized nonlinearly. This property has been included in where we proved that, in the absence of an explicit breaking due to the regularization, our approach generates the usual expansion of the chiral models in momenta and masses. In the present paper we evaluate the amplitude of the decay of the $`\pi _0`$ into two photons and show that our effective action correctly reproduces the chiral anomaly. These results are per se interesting, but they have in the present context a special relevance, related to the fact that in our effective action there is an explicit breaking of the chiral symmetry due to the regularization. This is because we are at the moment unable to treat the gluon composites analytically in analogy to the quark composites and therefore, to have access to the non-perturbative (in the gauge coupling constant) regime of the gluon field we are forced to define the theory on a lattice (but we emphasize that our approach is otherwise general). It is likely that in the near future we will have a chirally invariant form of this regularization also in the presence of non-Abelian fields . But for the time being to get rid of the spurious states of the fermions on a lattice we must introduce the so called Wilson term , which explicitly breaks the chiral invariance of our effective action. It is therefore very important to ascertain whether the Wilson parameter can be taken arbitrarily small in order to avoid this unpleasant consequence, but at the same time reproducing the chiral anomaly of QCD. This problem is made somewhat more intriguing by the fact that in our approach the quarks are perturbatively confined, because in the broken vacuum they acquire a large mass which does not allow their propagation to any finite order of our perturbative expansion. But if the quarks do not have any poles, why should we worry about the spurious ones and introduce the Wilson term? At the same time, if we omit this term, how can we reproduce the chiral anomaly? The present findings provide an answer to these questions. We find the standard results subject to a condition on the quark effective mass which is naturally satisfied in our approach . Then we can assume the Wilson parameter $`r`$ arbitrarily small and forget the Wilson term in our expansion in the strong sector, even though we must retain it in the electromagnetic amplitude which is non analytic in this parameter. In the next Section, for the convenience of the reader, we summarize the essential steps of the derivation of the effective action. In Section 3 we evaluate the electromagnetic decay amplitude of the $`\pi _0`$, in Section 4 we evaluate the anomaly and in Section 5 we present our conclusions. ## 2 The quark-composites approach We assume the modified partition function $$Z=[dV][d\overline{\lambda }d\lambda ]\mathrm{exp}[S_GS_QS_C],$$ (2.1) where $`S_G`$ is the action of the gauge fields, that is the Yang-Mills and Maxwell actions, $`S_Q`$ is the action of the quark fields and $`S_C`$ is a four fermions irrelevant operator which provides the kinetic terms for the quark composites with the quantum numbers of the chiral mesons. Therefore it will not have the form of the Nambu-Jona-Lasinio or Gross-Neveu models, that is of the so-called chirally extended QCD or $`\chi `$QCD (see also for example ). $`\lambda `$ is the quark field while the gluon and electromagnetic fields are associated to the link variables $`V_\mu `$. Differentials in square brackets are understood to be the product of the differentials over the lattice sites and the internal indices. All the fields live in an euclidian lattice of spacing $`a`$. We introduce the following notation for the sum over the lattice $$(f,g)=a^4\underset{x}{}f(x)g(x).$$ (2.2) In this notation the quark action is $$S_Q=(\overline{\lambda },Q\lambda )+m_q(\overline{\lambda },\lambda ).$$ (2.3) As already stated we will use the Wilson form of the quark wave operator $$Q=\gamma _\mu \overline{}_\mu a\frac{r}{2}\mathrm{}.$$ (2.4) The symmetric derivative $`\overline{}_\mu `$ and the Laplacian $`\mathrm{}`$ are covariant and are defined in terms of the right/left derivatives $$(_\mu ^\pm )_{xy}=\pm \frac{1}{a}\left(\delta _{x\pm \widehat{\mu },y}V_{\pm \mu }(x)\delta _{xy}\right)$$ (2.5) according to $`\overline{}_\mu `$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(_\mu ^++_\mu ^{}\right)`$ (2.6) $`\mathrm{}`$ $`=`$ $`{\displaystyle \underset{\mu }{}}_\mu ^+_\mu ^{}={\displaystyle \underset{\mu }{}}_\mu ^{}_\mu ^+={\displaystyle \underset{\mu }{}}{\displaystyle \frac{1}{a}}\left(_\mu ^+_\mu ^{}\right)`$ (2.7) We adopt the standard conventions $`V_\mu (x)`$ $`=`$ $`\mathrm{exp}\left[EaA_\mu ^{em}\left(x+{\displaystyle \frac{1}{2}}\widehat{\mu }\right)+gaA_\mu ^{YM}\left(x+{\displaystyle \frac{1}{2}}\widehat{\mu }\right)\right]`$ (2.8) $`V_\mu (x)`$ $`=`$ $`V_\mu ^{}(x\widehat{\mu }),`$ (2.9) where $`E`$ is a charge matrix. The chiral composites are the pions and the sigma $$\stackrel{}{\widehat{\pi }}=ik_\pi a^2\overline{\lambda }\gamma _5\stackrel{}{\tau }\lambda ,\widehat{\sigma }=k_\pi a^2\overline{\lambda }\lambda .$$ (2.10) $`\gamma _5`$ is assumed hermitian, the $`\stackrel{}{\tau }`$’s are the Pauli matrices and a factor of dimension (length)<sup>2</sup>, necessary to give the composites the dimension of a scalar, has been written in the form $`a^2k_\pi `$. Since for massless quarks the QCD action is chirally invariant, the action of the chiral mesons must be, apart from a linear breaking term, $`O(4)`$ invariant. It must then have the form $$S_C=\frac{1}{4}(\widehat{\mathrm{\Sigma }}^{},C\widehat{\mathrm{\Sigma }})\frac{1}{4a^2}(\chi ^{},\widehat{\mathrm{\Sigma }})+(\widehat{\mathrm{\Sigma }}^{},\chi ),$$ (2.11) where $$\widehat{\mathrm{\Sigma }}=\widehat{\sigma }i\stackrel{}{\tau }\stackrel{}{\widehat{\pi }},\chi =si\stackrel{}{\tau }\stackrel{}{p},A=\text{tr}^{isospin}A.$$ (2.12) We introduced the sources $`s`$ and $`\stackrel{}{p}`$ of the sigma and pion fields (their coupling to the quarks differ by a factor $`k_\pi `$ from the notation of ). Heuristic considerations, based on experience with simple, solvable models, lead to the following form of the wave operator of the chiral composites $$C=\frac{\rho ^4}{a^4}\frac{1}{\mathrm{}+\rho ^2/a^2},$$ (2.13) where $`\rho `$ is a dimensionless parameter. The irrelevance by power counting of $`S_C`$ requires that in the continuum limit $`\rho `$ do not to vanish and $`k_\pi `$, as well as the product $`k_\pi \rho `$, do not diverge. Under these conditions operators of dimension higher than 4 are accompanied by the appropriate powers of the cut-off. The operator $`C`$ behaves, as a function of the distance, as the propagator of a particle with a mass divergent at least as the cutoff. Therefore, even though strictly speaking it is nonlocal, its departure from locality is very mild in general, and very small with our present choice of $`\rho 1/a`$ which makes the mass appearing in $`C`$ divergent as the square of the cutoff. Ultimately, however, its irrelevance can be proven by showing that its local approximation $`C(\rho ^2/a^2)\mathrm{}`$ yields the same results (at the cost of more involved calculations). This proof is at present not complete. We replace the chiral composites by the auxiliary fields $$\mathrm{\Sigma }=\mathrm{\Sigma }_0i\stackrel{}{\tau }\stackrel{}{\mathrm{\Sigma }}$$ (2.14) by means of the Stratonovich-Hubbard transformation . Ignoring, as we will systematically do in the sequel, field independent factors, the partition function can be written $`Z`$ $`=`$ $`{\displaystyle [dV][d\overline{\lambda }d\lambda ]\left[\frac{d\mathrm{\Sigma }}{\sqrt{2\pi }}\right]\mathrm{exp}\left[S_GS_0+(\overline{\lambda },(DQ)\lambda )\right]}`$ (2.15) $`=`$ $`{\displaystyle [dV]\left[\frac{d\mathrm{\Sigma }}{\sqrt{2\pi }}\right]\mathrm{exp}\left[S_GS_0+\text{Tr}\mathrm{ln}(DQ)\right]},`$ where “Tr” is the trace over both space and internal degrees of freedom and we introduced the functions of the auxiliary fields $`S_0`$ $`=`$ $`{\displaystyle \frac{1}{4}}\rho ^4(\mathrm{\Sigma }^{},(a^4C)^1\mathrm{\Sigma }),`$ (2.16) $`D`$ $`=`$ $`{\displaystyle \frac{1\gamma _5}{2}}k_\pi \left[\rho ^2\mathrm{\Sigma }+\chi \right]+{\displaystyle \frac{1+\gamma _5}{2}}k_\pi \left[\rho ^2\mathrm{\Sigma }^{}+\chi ^{}\right].`$ (2.17) Eventually we will set $`s=m`$, the breaking parameter of the chiral symmetry, which must be distinguished from the quark mass $`m_q`$. The scaling with the lattice spacing that we will assume for $`m`$ will render the corresponding term irrelevant. The quark mass is absorbed in the breaking parameter $`m`$ according to $$mm\frac{1}{k_\pi }m_q.$$ (2.18) Since $`1/k_\pi `$ will play the role of an expansion parameter, the contribution from $`m_q`$ will be sub-leading in our expansion. The derivative nature of the couplings of the pions is exhibited after the transformation $$\mathrm{\Sigma }=RU$$ (2.19) where $`U`$ is an element of SU(2), and $$R^2=\mathrm{\Sigma }_0^2+\stackrel{}{\mathrm{\Sigma }}^2.$$ (2.20) The volume element $$\left[\frac{d\mathrm{\Sigma }}{\sqrt{2\pi }}\right]=\left[\frac{dR}{\sqrt{2\pi }}\right][dU]\mathrm{exp}\underset{x}{}3\mathrm{ln}R$$ (2.21) provides the Haar measure $`[dU]`$ over the group. We get the effective action $`\stackrel{~}{S}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle \underset{\mu }{}}(_\mu (RU^{}),^\mu (RU))+{\displaystyle \frac{\rho ^2}{2a^2}}(R,R)\text{Tr }\mathrm{ln}R{\displaystyle \underset{x}{}}3\mathrm{ln}R`$ (2.22) $`{\displaystyle \frac{1}{2}}\text{Tr }\mathrm{ln}\left(1{\displaystyle \frac{1}{\rho ^2R}}\chi ^{}U\right){\displaystyle \frac{1}{2}}\text{Tr }\mathrm{ln}\left(1+{\displaystyle \frac{1}{\rho ^2R}}U^{}\chi \right)`$ $`\text{Tr }\mathrm{ln}\left(1+D^1Q\right),`$ where $`_\mu `$ is the right or left covariant derivative. After our manipulations the partition function becomes $$Z=[dV]\left[\frac{dR}{\sqrt{2\pi }}\right]\left[dU\right]\mathrm{exp}\left[S_G\stackrel{~}{S}\right].$$ (2.23) The minimum of $`\stackrel{~}{S}`$ is obtained for $$U=1,\overline{R}=\sqrt{\mathrm{\Omega }}\frac{1}{a\rho }\left[1\frac{am}{2\rho \sqrt{\mathrm{\Omega }}}\right],$$ (2.24) where $`\mathrm{\Omega }`$ is the total number of quark components. In our case, by collecting the spinorial, colour and flavour indices we get $$\mathrm{\Omega }=24.$$ (2.25) The expansion around the minimum is naturally organized as a series in $`1/\sqrt{\mathrm{\Omega }}`$. In this framework we have a realization of the spontaneous breaking of the chiral invariance in QCD. The dominant part of the Lagrangian density, neglecting the fluctuations of $`R`$ and terms arising from the expansion of $`\overline{R}`$ with respect to $`m`$ is identical to that of the chiral models $$_2=\frac{1}{4}f_\pi ^2_\mu U^{}^\mu U2B\left(\chi ^{}U+U^{}\chi \right),$$ (2.26) with the identifications $$f_\pi =\frac{\sqrt{\mathrm{\Omega }}}{a\rho },B=\frac{1}{2a^2f_\pi }.$$ (2.27) If we confine ourselves to this leading term, we must assume $`f_\pi =92\text{Mev}`$. After these positions we recognize that the expansion in $`1/\sqrt{\mathrm{\Omega }}`$ is equivalent to that one in $`1/f_\pi `$ . Note that the above definition implies that $`\rho 1/a`$. Other scalings with the lattice spacing are possible, but will not be considered here. If we introduce the pion field $`\pi `$ according to $$U=\mathrm{exp}\left(\frac{i}{f_\pi }\stackrel{}{\tau }\stackrel{}{\pi }\right),$$ (2.28) we have for the pion mass $`m_\pi `$ and the chiral condensate the relations $$m_\pi ^2=2mB,k_\pi 0|\overline{\lambda }\lambda |0=2f_\pi ^2B.$$ (2.29) The presence of the factor $`k_\pi `$ is due to the fact that the source $`s`$ has a coupling to the quark fields that differs by this factor from the conventions of . It should also be noted that in the present case $`m`$ vanishes while $`B`$ diverges in the continuum limit. Let us examine the mass of the quarks and the $`\sigma `$ in the broken vacuum. According to eq. (2.15) the quark effective mass $$M_Q=k_\pi \rho ^2\overline{R}=k_\pi \rho ^2f_\pi ,$$ (2.30) is $`O(k_\pi f_\pi )`$, and therefore the quarks are perturbatively confined. Whether their mass is or is not divergent in the continuum limit, depends on how the product $`k_\pi \rho `$ scales with the lattice spacing. The $`\sigma `$ instead has a mass $`\sqrt{2}\rho /a`$ which is always divergent in the continuum limit. ## 3 The $`\pi _0\gamma \gamma `$ decay In this Section all the functions are in momentum space. We will perform the calculations to leading order in all the couplings, namely to order $`1/f_\pi `$, $`e^2/(2\pi )^2`$, while the Yang-Mills fields will be suppressed. The gauge fields $`A`$ therefore, are only photon fields. The amputated amplitude $`_{\alpha \beta }`$ for the decay of the $`\pi _0`$ into two photons is related to the three-point function $`\pi _0(q)A_\alpha (k_1)A_\beta (k_2)`$ according to $`\pi _0(q)A_\alpha (k_1)A_\beta (k_2)=`$ $`=_{\alpha \beta }(k_1,k_2)G_\pi (q^2)G_A(k_1^2)G_A(k_2^2)(2\pi )^4\delta ^4(q+k_1+k_2),`$ where the $`G`$ are the free propagators of the pion and the photons. $`_{\alpha \beta }`$ can be obtained by taking functional derivatives of the effective action with respect to the Maxwell and the pion fields $`(2\pi )^4\delta ^4(q+k_1+k_2)_{\alpha \beta }(k_1,k_2)=`$ $`=`$ $`{\displaystyle \frac{\delta ^2}{\delta A_\alpha (k_1)\delta A_\beta (k_2)}}\text{Tr }\left[{\displaystyle \frac{1}{DQ}}{\displaystyle \frac{\delta D}{\delta \pi _3(q)}}\right]|_{\stackrel{}{\pi }=A=0}`$ $`=`$ $`{\displaystyle \frac{\delta ^2}{\delta A_\alpha (k_1)\delta A_\beta (k_2)}}\text{Tr }\left[\left(1+{\displaystyle \frac{1}{M_QQ}}Q\right){\displaystyle \frac{1}{M_Q}}{\displaystyle \frac{\delta D}{\delta \pi _3(q)}}\right]|_{\stackrel{}{\pi }=A=0}`$ $`=`$ $`{\displaystyle \frac{\delta ^2}{\delta A_\alpha (k_1)\delta A_\beta (k_2)}}\text{Tr }\left[{\displaystyle \frac{1}{M_QQ}}{\displaystyle \frac{1}{2}}\{Q,{\displaystyle \frac{1}{M_Q}}{\displaystyle \frac{\delta D}{\delta \pi _3(q)}}\}\right]|_{\stackrel{}{\pi }=A=0}.`$ It is easy to relate the above expression the anomalous triangle graph identifying vertices and propagators. Let $$\left(\frac{1}{M_QQ}\right)(k_1,k_2)|_{\stackrel{}{\pi }=A=0}=\delta _{k_1,k_2}S(k_1).$$ (3.3) $`S`$ is the free propagator of the quark in the broken vacuum $$S(k)=\left(i\overline{k/}+W(k)\right)^1$$ (3.4) with $`W(k)`$ $`=`$ $`a{\displaystyle \frac{r}{2}}\widehat{k}^2+M_Q`$ (3.5) $`\overline{k}_\alpha `$ $`=`$ $`{\displaystyle \frac{1}{a}}\mathrm{sin}ak_\alpha `$ (3.6) $`\widehat{k}_\alpha `$ $`=`$ $`{\displaystyle \frac{1}{a}}\mathrm{\hspace{0.17em}2}\mathrm{sin}a{\displaystyle \frac{k_\alpha }{2}}.`$ (3.7) Having in mind the continuum limit, we neglected $`\sqrt{\mathrm{\Omega }}k_\pi m`$ with respect to $`M_Q`$. Next we identify the vertices. The electromagnetic vertex is $$\frac{\delta Q(k_1,k_2)}{\delta A_\alpha (k)}|_{\stackrel{}{\pi }=A=0}=EV_\alpha (k_1,k_2)(2\pi )^4\delta ^4(k_1+k_2+k)$$ (3.8) with $$V_\alpha (k_1,k_2)=V_\alpha \left(\frac{k_1+k_2}{2}\right)$$ (3.9) and $$V_\alpha (k)=\frac{}{k_\alpha }S^1(k).$$ (3.10) The quark-charge matrix $`E`$ is defined as $$E=\frac{e}{6}(1+3\tau _3)$$ (3.11) where $`e`$ is the electric charge of the proton. The anomalous vertex is $$\frac{1}{2}\{Q,\frac{1}{M_Q}\frac{D(k_1,k_2)}{\pi _3(q)}\}|_{\stackrel{}{\pi }=A=0}=i\frac{1}{f_\pi }\gamma _5\tau _3W(q)(2\pi )^4\delta ^4(k_1+k_2+q).$$ (3.12) Using the above equations the decay amplitude becomes $`_{\alpha \beta }(k_1,k_2)`$ $`=`$ $`i{\displaystyle \frac{1}{f_\pi }}{\displaystyle }\left({\displaystyle \frac{dk}{2\pi }}\right)^4\text{tr }[E^2\tau _3\gamma _5W(k)S(k+k_1)`$ (3.13) $`V_\alpha (k+k_1,k)S(k)V_\beta (k,kk_2)S(kk_2)],`$ and after the sum over colour and isospin indices $`_{\alpha \beta }(k_1,k_2)=2i{\displaystyle \frac{e^2}{f_\pi }}{\displaystyle \left(\frac{dk}{2\pi }\right)^4W(k)}`$ (3.14) $`\text{tr}^{spin}\left[\gamma _5S(k+k_1)V_\alpha (k+k_1,k)S(k)V_\beta (k,kk_2)S(kk_2)\right].`$ Let us develop the expression (3.14) in series of the photon momenta $`k_1`$ and $`k_2`$. By using the Ward identity (3.10) in the form $$SV_\alpha S=S_\alpha S^1S=_\alpha S,$$ (3.15) where all the functions are evaluated at the same momentum $`k`$, one can reduce the number of $`\gamma `$-matrices appearing in the trace. It is easy to show in this way that the first non-vanishing contribution comes at order $`k_1k_2`$ and is given by $`_{\alpha \beta }(k_1,k_2)`$ $`=`$ $`i{\displaystyle \frac{e^2}{f_\pi }}k_{1\mu }k_{2\nu }{\displaystyle }\left({\displaystyle \frac{dk}{2\pi }}\right)^4\text{tr}^{spin}\{\gamma _5W`$ (3.16) $`[{\displaystyle \frac{1}{2}}S_\mu V_\alpha S_\nu V_\beta S+2_\mu SV_\alpha SV_\beta _\nu S`$ $`+_\mu SV_\alpha S_\nu V_\beta S+S_\mu V_\alpha SV_\beta _\nu S]\}`$ which, after a few integrations by parts, becomes $`_{\alpha \beta }(k_1,k_2)`$ $`=`$ $`i{\displaystyle \frac{e^2}{f_\pi }}k_{1\mu }k_{2\nu }{\displaystyle }\left({\displaystyle \frac{dk}{2\pi }}\right)^4\text{tr}^{spin}\{\gamma _5W`$ (3.17) $`[_\mu SV_\alpha _\nu SV_\beta S+SV_\alpha _\mu SV_\beta _\nu S]\}.`$ By using the explicit expressions of $`S`$ and $`V`$ we get $`_{\alpha \beta }(k_1,k_2)`$ $`=`$ $`2i{\displaystyle \frac{e^2}{f_\pi }}k_{1\mu }k_{2\nu }{\displaystyle \left(\frac{dt}{2\pi }\right)^4\text{tr}^{spin}[\gamma _5\gamma _\mu \gamma _\alpha \gamma _\nu \gamma _\beta ]\frac{W}{d^3}}`$ (3.18) $`\mathrm{cos}t_\alpha \mathrm{cos}t_\nu \mathrm{cos}t_\beta \left[W\mathrm{cos}t_\mu 4_\mu W\mathrm{sin}t_\mu \right]`$ where $$d=\overline{t}^2+W^2(t).$$ (3.19) We assume, as usual, that in the continuum limit $$\underset{a0}{lim}\frac{aM_Q}{r}=0,$$ (3.20) a condition which can be naturally satisfied in our approach . Then we see that the integral takes its contribution only from the pole and gives the well known result $`_{\alpha \beta }(k_1,k_2)`$ $`=`$ $`8i{\displaystyle \frac{e^2}{f_\pi }}ϵ_{\mu \nu \alpha \beta }k_{1\mu }k_{2\nu }{\displaystyle \left(\frac{dt}{2\pi }\right)^4\mathrm{cos}t_2\mathrm{cos}t_3\mathrm{cos}t_4_1\frac{\mathrm{sin}t_1}{d^2}}`$ (3.21) $`=`$ $`i{\displaystyle \frac{1}{f_\pi }}\left({\displaystyle \frac{e}{2\pi }}\right)^2ϵ_{\mu \nu \alpha \beta }k_{1\mu }k_{2\nu }.`$ ## 4 The chiral anomaly Apart from the explicit breakings, that is the source and the Wilson terms, our partition function is exactly invariant under the symmetry transformations $$Ug_RUg_L^{},$$ (4.1) with $`(g_R,g_L)`$ SU(2)$`\times `$SU(2). While the presence of the source term is necessary to provide a mass term to the pions, the Wilson term is a residue of the doubling problem of lattice fermions which induces a departure from the general structure of the chiral Lagrangians. Nonetheless this term is responsible for the correct chiral anomaly in lattice QCD, which, as it is well known, is deeply related to the electromagnetic decay of the $`\pi _0`$. Let us derive the general Ward identities. For the sake of simplicity hereafter we shall restrict the field $`R`$ at his extremal value and we neglect its fluctuations. For an arbitrary infinitesimal transformation $$UU+\delta U$$ (4.2) with $$\delta U=w_RUUw_L.$$ (4.3) The variation of the first term of $`\stackrel{~}{S}`$ is $`{\displaystyle \frac{1}{2}}\delta \text{Tr }{\displaystyle \underset{\mu }{}}\left[_\mu ^+U^{}_\mu ^+U\right]=`$ $`=`$ $`\delta \text{Tr }{\displaystyle \underset{\mu }{}}\left[U_{x+\mu }^{}U_x+U_x^{}U_{x+\mu }\right]`$ $`=`$ $`\text{Tr }{\displaystyle \underset{\mu }{}}\left[U_{x+\mu }^{}\delta U_{x+\mu }U_{x+\mu }^{}U_x+U_{x+\mu }^{}\delta U_xU_x^{}\delta U_xU_x^{}U_{x+\mu }+U_x^{}\delta U_{x+\mu }\right]`$ $`=`$ $`\text{Tr }{\displaystyle \underset{\mu }{}}\left[U_xU_{x+\mu }^{}+U_{x+\mu }U_x^{}\right]\left[\delta U_{x+\mu }U_{x+\mu }^{}\delta U_xU_x^{}\right]`$ $`=`$ $`\text{Tr }{\displaystyle \underset{\mu }{}}\left[U_xU_{x+\mu }^{}+U_{x+\mu }U_x^{}\right]_\mu ^+\left[\delta U_xU_x^{}\right]`$ $`=`$ $`\text{Tr }{\displaystyle \underset{\mu }{}}\left\{_\mu ^{}\left[U_xU_{x+\mu }^{}+U_{x+\mu }U_x^{}\right]\right\}\left[\delta U_xU_x^{}\right]`$ $`=`$ $`\text{Tr }{\displaystyle \underset{\mu }{}}\left\{_\mu ^{}\left[U\left(_\mu ^+U^{}\right)+\left(_\mu ^+U\right)U^{}\right]\right\}\left[\delta UU^{}\right]`$ $`=`$ $`\text{Tr }{\displaystyle \underset{\mu }{}}\left\{_\mu ^{}\left[U\left(_\mu ^+U^{}\right)+\left(_\mu ^+U\right)U^{}\right]\right\}\left[w_RUw_LU^{}\right]`$ $`=`$ $`\text{Tr }{\displaystyle \underset{\mu }{}}\left\{_\mu ^{}\left[U\left(_\mu ^+U^{}\right)+\left(_\mu ^+U\right)U^{}\right]\right\}w_R`$ $`+\text{Tr }{\displaystyle \underset{\mu }{}}\left\{_\mu ^{}\left[U^{}\left(_\mu ^+U\right)+\left(_\mu ^+U^{}\right)U\right]\right\}w_L`$ which has the form of the divergence of a current. More precisely for a vectorial transformation, that is when $`g_L`$ $`=`$ $`g_R`$ (4.5) $`w_L`$ $`=`$ $`w_R=w_V`$ (4.6) we get the vector current $$\stackrel{}{𝒱}_\mu =\frac{1}{2}\stackrel{}{\tau }\left\{[_\mu ^+U,U^{}]+[_\mu ^+U^{},U]\right\}.$$ (4.7) While for an axial transformation $`g_L`$ $`=`$ $`g_R^{}`$ (4.8) $`w_L`$ $`=`$ $`w_R=w_A`$ (4.9) we get $$\stackrel{}{𝒜}_\mu =\frac{1}{2}\stackrel{}{\tau }\left[\{_\mu ^+U,U^{}\}\{_\mu ^+U^{},U\}\right].$$ (4.10) Let us consider now the explicit breaking at linear level $$\delta \text{Tr }\left(\chi U^{}+U\chi ^{}\right)=\text{Tr }\left[\left(U\chi ^{}\chi U^{}\right)w_R\left(\chi ^{}UU^{}\chi \right)w_L\right].$$ (4.11) From this expression in the case of a vector transformation we get $$\delta \text{Tr }\left(\chi U^{}+U\chi ^{}\right)=\text{Tr }\left\{[U,\chi ^{}]+[U^{},\chi ]\right\}w_V$$ (4.12) while for an axial transformation $$\delta \text{Tr }\left(\chi U^{}+U\chi ^{}\right)=\text{Tr }\left[\{U,\chi ^{}\}\{U^{},\chi \}\right]w_A.$$ (4.13) From these expressions we get, in the absence of the fermionic determinant, when $`\chi =m`$ $`{\displaystyle \underset{\mu }{}}_\mu ^{}\stackrel{}{𝒱}_\mu `$ $`=`$ $`0`$ (4.14) $`{\displaystyle \underset{\mu }{}}_\mu ^{}\stackrel{}{𝒜}_\mu `$ $`=`$ $`m_\pi ^2\stackrel{}{\tau }\left(UU^{}\right)`$ (4.15) which is nothing but the classical equation of motion. Let us now evaluate the variation of the fermionic determinant. Since we have already taken the linear contribution of the breaking term we will set now for simplicity $`\chi =0`$. Then $$D=M_Q\left[\frac{1\gamma _5}{2}U+\frac{1+\gamma _5}{2}U^{}\right]$$ (4.16) and its variation is $`\delta D`$ $`=`$ $`M_Q\left[{\displaystyle \frac{1\gamma _5}{2}}\delta U+{\displaystyle \frac{1+\gamma _5}{2}}\delta U^{}\right]`$ $`=`$ $`M_Q\left[{\displaystyle \frac{1\gamma _5}{2}}\delta U{\displaystyle \frac{1+\gamma _5}{2}}U^{}\delta UU^{}\right].`$ Therefore $`\delta \text{Tr }\mathrm{ln}(DQ)=\text{Tr }(DQ)^1\delta D=`$ $`=`$ $`\text{Tr }(DQ)^1M_Q\left[{\displaystyle \frac{1\gamma _5}{2}}\delta U{\displaystyle \frac{1+\gamma _5}{2}}U^{}\delta UU^{}\right]`$ $`=`$ $`\text{Tr }M_Q\left[(DQ)^1{\displaystyle \frac{1\gamma _5}{2}}U+{\displaystyle \frac{1+\gamma _5}{2}}U^{}(DQ)^1\right]w_L+`$ $`\text{Tr }M_Q\left[{\displaystyle \frac{1\gamma _5}{2}}U(DQ)^1(DQ)^1{\displaystyle \frac{1+\gamma _5}{2}}U^{}\right]w_R.`$ If we specialize to a vector transformation $`\delta \text{Tr }\mathrm{ln}(DQ)`$ $`=`$ $`M_Q\text{Tr }[(DQ)^1,{\displaystyle \frac{1\gamma _5}{2}}U+{\displaystyle \frac{1+\gamma _5}{2}}U^{}]w_V`$ (4.19) $`=`$ $`\text{Tr }[(DQ)^1,D]w_V`$ $`=`$ $`\text{Tr }[(DQ)^1,Q]w_V`$ $`=`$ $`\text{Tr }(DQ)^1[Q,w_V].`$ While for an axial transformation $`\delta \text{Tr }\mathrm{ln}(DQ)`$ $`=`$ $`M_Q\text{Tr }\{(DQ)^1,{\displaystyle \frac{1\gamma _5}{2}}U{\displaystyle \frac{1+\gamma _5}{2}}U^{}\}w_A`$ (4.20) $`=`$ $`\text{Tr }\{(DQ)^1,\gamma _5D\}w_A`$ $`=`$ $`\text{Tr }\gamma _5\{(DQ)^1,D\}w_A`$ $`=`$ $`\text{Tr }\gamma _5\{(DQ)^1,Q\}w_A`$ $`=`$ $`\text{Tr }(DQ)^1\left(Q\gamma _5w_A+w_A\gamma _5Q\right)`$ $`=`$ $`\text{Tr }(DQ)^1\left(\gamma _5[Q,w_A]+\{Q,\gamma _5\}w_A\right)`$ We arrive at the equations $`{\displaystyle \underset{\mu }{}}_\mu ^{}\stackrel{}{𝒱}_\mu `$ $`=`$ $`{\displaystyle \frac{2}{f_\pi ^2}}\text{Tr }\stackrel{}{\tau }[Q,(DQ)^1]`$ (4.21) $`{\displaystyle \underset{\mu }{}}_\mu ^{}\stackrel{}{𝒜}_\mu `$ $`=`$ $`m_\pi ^2\stackrel{}{\tau }\left(UU^{}\right){\displaystyle \frac{2}{f_\pi ^2}}\text{Tr }\stackrel{}{\tau }[\gamma _5Q,(DQ)^1]`$ (4.22) $`{\displaystyle \frac{2}{f_\pi ^2}}\text{Tr }\stackrel{}{\tau }(DQ)^1\{Q,\gamma _5\}.`$ where the new terms with respect to (4.15) correspond to the fermionic contributions to the currents. In particular the last term with the anti-commutator originates the anomalous vertex which entered in the computation of the decay rate of the $`\pi _0`$ of the previous section. It is therefore this term which is responsible for the breaking, at the quantum level, of the axial symmetry in the continuum limit. The anomaly of the underlying QCD is correctly reproduced by our effective chiral lagrangian. ## 5 Conclusions We evaluated the decay amplitude of the electromagnetic decay of the pion by an effective action derived from QCD in the quark composites approach. This allowed us to treat in a unified way the anomalous and the electromagnetic vertices. As a consequence of formula (3.21), to leading order the decay rate turns out to take the standard value of $`7.63\text{eV}`$ (see for example ), surprisingly close to the experimental rate $`(7.37\pm 1.5)\text{eV}`$. It would be interesting to check whether the strong corrections are sufficiently small in our theory. The present results allow us to establish a close relation between our expansion for the chiral mesons and the chiral models. Since the chiral anomaly is independent of the value of the Wilson parameter $`r`$, provided it is different from zero, we can then assume $`r=O(\mathrm{\Omega }^n)`$, with $`n`$ arbitrarily large. Now there is no reason to believe that the amplitudes in the strong sector are not analytic in $`r`$, and therefore studying the strong interactions in the framework of our $`1/\sqrt{\mathrm{\Omega }}1/f_\pi `$ expansion, we can forget the Wilson term. We showed that in this case our theory, under the standard condition (3.20) which is naturally satisfied, generates only terms of the chiral models . In the electromagnetic sector, on the contrary, the amplitudes are not analytic in $`r`$, as we have seen in the previous section, and the Wilson term must be retained to get correct results. We think that the quark-composites approach might prove useful also in numerical simulations. One can consider the action of eq. (2.15) as an improved lattice QCD action, where the chiral limit is obtained in the simple limit of zero breaking term, where the pion mass vanishes (rather than by a fine tuning). There is a price to pay because of the inclusion of the auxiliary fields, but this should be rewarded by a simpler evaluation of the quark determinant because of the dominance of the diagonal contribution in configuration space. There are indeed indications in this sense in the work presented in . ## 6 Acknowledgments It is a pleasure to thank Andrea Pelissetto for useful discussions.
warning/0004/math0004170.html
ar5iv
text
# An Infinite Suite of Links–Gould Invariants ## 1 Overview In 1992, Jon Links and Mark Gould described a method for constructing link invariants from quantum superalgebras. That work stopped short of evaluations of the invariants due to want of an efficient computational method. In 1999, the author, in collaboration with Jon Links and Louis Kauffman , first evaluated a two-variable example of one these invariants, using a state model. We used the $`(0,0|\alpha )`$ representations of $`U_q[gl(2|1)]`$, and labeled our resulting $`(1,1)`$-tangle invariant $`LG`$, ‘*the* Links–Gould invariant’. In that paper, and subsequently in , we showed that whilst $`LG`$ would detect neither inversion nor mutation, it was still able to distinguish all prime knots of up to $`10`$ crossings, making it more powerful than the HOMFLY and Kauffman invariants. Here, we generalise the notation, denoting $`LG^{m,n}`$ as “the Links–Gould invariant associated with the $`(\dot{0}_m|\dot{\alpha }_n)`$ representation of $`U_q[gl(m|n)]`$”. For the case $`n=1`$, we will write $`LG^mLG^{m,1}`$, so our previous invariant $`LG`$ was in fact $`LG^2`$. This generalisation is motivated by the automation of a procedure to construct the appropriate R matrices ; previously, we were limited to the $`m=2`$ case, for which the R matrix had been calculated by hand. We explicitly demonstrate the construction of state model parameters for $`LG^{m,n}`$, illustrating our results for $`LG^m`$, for the cases $`m=1,2,3,4`$. Further, we describe some of the gross properties of these invariants, and provide a limited set of evaluations of them. Although these invariants $`LG^{m,n}`$ are not more powerful in their gross properties than $`LG^2`$ (they can detect neither inversion nor mutation), each one is expected to distinguish many more knots $`K`$ as the degree of the polynomials $`LG_K^{m,n}`$ increases rapidly with $`m`$ and $`n`$. Perhaps more significantly, the development of the current formalism points the way towards automation of the evaluation of more general classes of quantum link invariants; a discussion of this is provided. ## 2 Quantum superalgebra state models Corresponding to each finite dimensional highest weight representation of each quantum superalgebra, there exists a quantum link invariant ($`LG`$), originally described in . These invariants are similar to those associated with the usual (i.e. ungraded) quantum algebras (e.g. ), although there are some technical differences. Here, we describe the construction of parameters for state models for evaluating a class of these invariants. Specifically, we will define $`LG^{m,n}`$ to be the quantum link invariant associated with the representation $`\pi \pi _\mathrm{\Lambda }`$ of highest weight $`\mathrm{\Lambda }=(\dot{0}_m|\dot{\alpha }_n)`$, of the quantum superalgebra $`U_q[gl(m|n)]`$. To do so, we first broadly introduce the algebraic structures, then we briefly review the terminology used to describe state model parameters, and finally, we look at the construction of specific state model parameters for our particular class of representations. ### 2.1 The quantum superalgebra $`U_q[gl(m|n)]`$ $`U_q[gl(m|n)]`$ is a unital super (i.e. $`_2`$-graded) algebra with free parameter $`q`$. In the limit $`q1`$, it degenerates to the ordinary Lie superalgebra $`gl(m|n)`$. Here, we provide a broad outline of $`U_q[gl(m|n)]`$ in terms of generators and relations, for readers not familiar with it. This material is largely abstracted from the fuller description contained in (see also ). #### 2.1.1 $`U_q[gl(m|n)]`$ generators A set of generators for $`U_q[gl(m|n)]`$ is: $`\{\begin{array}{ccc}\hfill K_a,& 1am+n\hfill & \mathrm{Cartan}\hfill \\ \hfill E_{}^{a}{}_{b}{}^{},& 1a<bm+n\hfill & \mathrm{raising}\hfill \\ \hfill E_{}^{b}{}_{a}{}^{},& 1a<bm+n\hfill & \mathrm{lowering}\hfill \end{array}\}.`$ An equivalent notation for $`K_a`$ is $`q_a^{E_{}^{a}{}_{a}{}^{}}`$, where we have introduced the notation $`q_aq^{()^{[a]}}`$. For any power $`N`$, we may write $`q_a^N`$, and hence $`K_a^N`$, thus for $`M,N`$: $`K_a^MK_a^N=K_a^{M+N}\mathrm{where}K_a^0\mathrm{Id},`$ where $`\mathrm{Id}`$ is the $`U_q[gl(m|n)]`$ identity element. Using the following $`_2`$ grading on the $`gl(m|n)`$ *indices*: $`[a]\{\begin{array}{ccc}0\hfill & \mathrm{if}1am\hfill & \mathrm{even}\hfill \\ 1\hfill & \mathrm{if}m+1am+n\hfill & \mathrm{odd},\hfill \end{array}`$ we may define a natural $`_2`$ grading on the generators: $`[K_a^N]0,[E_{}^{a}{}_{b}{}^{}][a]+[b](\mathrm{mod}\mathrm{\hspace{0.33em}2}),`$ and we use the terms “even” and “odd” for generators in the same manner as we do for indices. Elements of $`U_q[gl(m|n)]`$ are said to be *homogeneous* if they are linear combinations of generators of the same grading. The product $`XY`$ of homogeneous $`X,YU_q[gl(m|n)]`$ has grading: $`[XY][X]+[Y](\mathrm{mod}\mathrm{\hspace{0.33em}2}).`$ Within the full set of generators, we have the $`U_q[gl(m|n)]`$ *simple* generators: $`\{\begin{array}{ccc}\hfill K_a,& 1am+n\hfill & \mathrm{Cartan}\hfill \\ \hfill E_{}^{a}{}_{a+1}{}^{},& 1a<m+n\hfill & \mathrm{simple}\mathrm{raising}\hfill \\ \hfill E_{}^{a+1}{}_{a}{}^{},& 1a<m+n\hfill & \mathrm{simple}\mathrm{lowering}\hfill \end{array}\},`$ such that the remaining nonsimple generators may be expressed in terms of these \[25, p1238, (2)\]. The fact that there are $`m+n1`$ simple raising generators indicates that $`U_q[gl(m|n)]`$ has rank $`m+n1`$. #### 2.1.2 $`U_q[gl(m|n)]`$ relations The *graded commutator* $`[,]:U_q[gl(m|n)]\times U_q[gl(m|n)]U_q[gl(m|n)]`$, is defined for homogeneous $`X,YU_q[gl(m|n)]`$ by: $`[X,Y]XY()^{[X][Y]}YX,`$ and extended by linearity. With this, we have the following $`U_q[gl(m|n)]`$ relations: 1. The Cartan generators all commute: $`K_a^MK_b^N=K_b^NK_a^M,M,N.`$ 2. The Cartan generators commute with the simple raising and lowering generators in the following manner: $`K_aE_{}^{b}{}_{b\pm 1}{}^{}=q_a^{(\delta _b^a\delta _{b\pm 1}^a)}E_{}^{b}{}_{b\pm 1}{}^{}K_a.`$ 3. The squares of the odd simple generators are zero: $`(E_{}^{m}{}_{m+1}{}^{})^2=(E_{}^{m+1}{}_{m}{}^{})^2=0.`$ (This implies that the squares of nonsimple odd generators are also zero.) 4. The non-Cartan generators satisfy the following commutation relations: $`[E_{}^{a}{}_{a+1}{}^{},E_{}^{b+1}{}_{b}{}^{}]=\delta _b^a{\displaystyle \frac{K_aK_{a+1}^1K_a^1K_{a+1}}{q_a\overline{q}_a}},`$ where we have written $`\overline{q}q^1`$ for brevity. We also have, for $`|ab|>1`$, the commutations: $`E_{}^{a}{}_{a+1}{}^{}E_{}^{b}{}_{b+1}{}^{}=E_{}^{b}{}_{b+1}{}^{}E_{}^{a}{}_{a+1}{}^{}\mathrm{and}E_{}^{a+1}{}_{a}{}^{}E_{}^{b+1}{}_{b}{}^{}=E_{}^{b+1}{}_{b}{}^{}E_{}^{a+1}{}_{a}{}^{}.`$ 5. Lastly, we have the $`U_q[gl(m|n)]`$ *Serre relations*; their inclusion ensures that the algebra is reduced enough to be *simple*. We omit these for brevity; they are not required below. #### 2.1.3 $`U_q[gl(m|n)]`$ as a Hopf superalgebra When equipped with an appropriate<sup>2</sup><sup>2</sup>2The details of these structures are not required here; the reader can find them in . coproduct $`\mathrm{\Delta }`$, counit $`\epsilon `$ and antipode $`S`$, we may regard $`U_q[gl(m|n)]`$ as a quasitriangular Hopf superalgebra. This means that it possesses an R matrix $`\stackrel{ˇ}{R}`$, an operator on the tensor product $`U_q[gl(m|n)]U_q[gl(m|n)]`$, satisfying the quantum Yang–Baxter equation (QYBE) in the form: $$(\stackrel{ˇ}{R}I)(I\stackrel{ˇ}{R})(\stackrel{ˇ}{R}I)=(I\stackrel{ˇ}{R})(\stackrel{ˇ}{R}I)(I\stackrel{ˇ}{R}),$$ (4) immediately recognisable as the braid relation: $$\sigma _1\sigma _2\sigma _1=\sigma _2\sigma _1\sigma _2.$$ (5) ### 2.2 State model parameters The following comments briefly describe in what is well-documented in the literature (see, e.g. ). They are included so as to introduce our particular notation. A *state model* $`(\sigma ,C)`$ for a link invariant consists of two parameters: $`\sigma `$, an invertible rank $`4`$ tensor representing the braid generator (i.e. a positive crossing), and $`C`$, an invertible rank $`2`$ tensor representing a *positive handle*, that is an anticlockwise-oriented, vertical open arc, used to close a one string of a braid. From these, we may immediately define the representations corresponding to negative crossings (viz $`\overline{\sigma }\sigma ^1`$) and negative handles (viz $`\overline{C}C^1`$).<sup>3</sup><sup>3</sup>3We frequently use the notation $`\overline{X}`$ to mean $`X^1`$, in particular, writing $`\overline{\sigma }\sigma ^1`$ and $`\overline{C}C^{}`$ allows us to omit superfluous “$`+`$” signs, viz we write $`CC^+`$ for the positive handle. Our current collection of arcs is shown in Figure 1. Let $`\beta `$ be a braid corresponding to a link $`L\widehat{\beta }`$, formed from the vertical closure of $`\beta `$. The diagram components corresponding to $`\sigma `$ (and $`\overline{\sigma }`$) are sufficient to construct $`\beta `$, and those corresponding to $`C`$ (we don’t need $`\overline{C}`$) are then sufficient to construct $`\widehat{\beta }`$ from $`\beta `$. When the pair $`(\sigma ,C)`$ is chosen to satisfy the Reidemeister moves (below, we write R1, R2 and R3), we may form a link invariant from the contraction over the free indices of the tensors corresponding to the diagram components. It may happen that the algebraic structures underlying the model mean that this invariant will be zero on closed links (i.e. $`(0,0)`$-tangles), however, we may still form an invariant of $`(1,1)`$-tangles , by contracting over all but one free index, and obtain an invariant which is not necessarily trivial. Our invariants $`LG`$ are based on typical $`U_q[gl(m|n)]`$ representations, for which the appropriate supertrace is zero, hence we define our invariants to be $`(1,1)`$-tangle invariants. ### 2.3 State model parameters for $`U_q[gl(m|n)]`$ representations $`\mathrm{\Lambda }`$ Here, we integrate the materials of §2.1 and §2.2, allowing us to describe the construction of state model parameters corresponding to arbitrary $`U_q[gl(m|n)]`$ representations $`\mathrm{\Lambda }`$. Below, in §2.4, we perform some extra necessary calculations. After that, in §3, we specialise this material to the case $`\mathrm{\Lambda }=(\dot{0}_m|\dot{\alpha }_n)`$. So, how do we construct state model parameters $`(\sigma ,C)`$ corresponding to an invariant associated with an arbitrary $`U_q[gl(m|n)]`$ representation $`\pi `$? Firstly, the tensor product representation $`\stackrel{ˇ}{R}(\pi \pi )\stackrel{ˇ}{R}`$ necessarily satisfies the QYBE in the form (4), and hence the braid relation (5). This means that abstract tensors built from $`\stackrel{ˇ}{R}`$ are invariant under R2 and R3, hence we may construct representations of arbitrary braids from $`\stackrel{ˇ}{R}`$. Thus $`\sigma \kappa _\sigma \stackrel{ˇ}{R}`$ (for any scalar constant $`\kappa _\sigma `$) realises a representation of the braid generator. A technical point distinguishes the quantum superalgebra situation from that of the quantum algebra. Quantum superalgebra R matrices are in fact *graded*, and actually satisfy a *graded* QYBE. It is however, a simple matter to strip out this grading (i.e. apply an automorphism ),<sup>4</sup><sup>4</sup>4The $`\stackrel{ˇ}{R}`$ supplied in are normalised such that $`lim_{q1}\stackrel{ˇ}{R}`$ is a (graded) permutation matrix. Scaling by $`\kappa _\sigma `$ does not change that. yielding $`\stackrel{ˇ}{R}`$ satisfying the usual, ungraded QYBE. Secondly, to ensure that our invariant is an invariant of ambient isotopy, we must select $`C`$ to ensure that abstract tensors built from $`\sigma `$ and $`C`$ are also invariant under R1. To this end, we apply (a grading-stripped version of) the following result \[19, Lemma 2\] (see also ): $`(I\mathrm{str})[(Iq^{2h_\rho })\sigma ]=KI,`$ where the Cartan element $`q^{h_\rho }`$ is defined in §2.4, $`\mathrm{str}`$ is the supertrace, and $`K`$ is some constant depending on the normalisations of $`\sigma `$ and $`q^{h_\rho }`$. Writing $`S\pi (q^{2h_\rho })`$ for convenience; for any scalar constant $`\kappa _C`$, setting $`C\kappa _CS`$ allows us to represent positive handles. It remains to choose $`(\kappa _S,\kappa _C)`$ to satisfy R1. Thus, we demonstrate how to select $`\kappa _\sigma `$ and $`\kappa _C`$ such that the abstract tensor associated with removal of an isolated loop is invariant under R1. Figure 2 shows that for $`\sigma `$ and $`C`$ to satisfy R1, they must satisfy (Einstein summation convention): $$C_c^d\sigma _{db}^{ca}=\delta _b^a=C_c^d\overline{\sigma }_{db}^{ca},$$ (6) where the definitions of $`\kappa _\sigma `$ and $`\kappa _C`$ yield: $`\overline{\sigma }=\kappa _\sigma ^1\stackrel{ˇ}{R}^1`$, and $`\overline{C}=\kappa _C^1S^1`$. So, if we have established $`S`$ and $`\stackrel{ˇ}{R}`$, we may determine $`\kappa _\sigma `$ and $`\kappa _C`$ by solving the following equations: $`\kappa _C\kappa _\sigma S_c^d(\stackrel{ˇ}{R})_{db}^{ca}=\kappa _C\kappa _\sigma ^1S_c^d(\stackrel{ˇ}{R}^1)_{db}^{ca}=\delta _b^a.`$ Setting $`a=b=1`$ and using the fact that $`S`$ is diagonal, we thus have: $`\kappa _\sigma =X_1^{\frac{1}{2}}X_2^{+\frac{1}{2}},\kappa _C=X_1^{\frac{1}{2}}X_2^{\frac{1}{2}},`$ where $`X_1S_c^c(\stackrel{ˇ}{R})_{c1}^{c1}`$ and $`X_2S_c^c(\stackrel{ˇ}{R}^1)_{c1}^{c1}`$. Note that reflecting the diagrams of Figure 2 about a vertical axis yields exactly the same constraints on $`\kappa _\sigma `$ and $`\kappa _C`$. To see this, the constraints obtained by reflecting the diagrams in a vertical axis are: $$\overline{\sigma }_{bd}^{ac}\overline{C}_c^d=\delta _b^a=\sigma _{bd}^{ac}\overline{C}_c^d,$$ (7) however, we have: $`\overline{\sigma }_{bd}^{ac}=(\sigma _{db}^{ca})|_{q\overline{q}}`$ and $`\overline{C}_c^d=(C_c^d)|_{q\overline{q}}`$. Replacing $`q\overline{q}`$ in (7) and applying these equivalences recovers (6). Similarly, reversing the orientations of the strings in Figure 2 yields no new constraints. What is significant in the above is that we have explicit formulae for automatically scaling from $`(\stackrel{ˇ}{R},S)`$ to $`(\sigma ,C)`$, something apparently absent in the literature. Variations on these formulae should hold for a much wider class of representations and algebraic structures. We write them up as a little lemma: ###### Lemma 1 Let $`\pi `$ be a finite-dimensional highest weight $`U_q[gl(m|n)]`$ representation, for which we have computed $`\stackrel{ˇ}{R}(\pi \pi )\stackrel{ˇ}{R}`$ and $`S\pi (q^{2h_\rho })`$. Then the state model parameters $`(\sigma ,C)`$ for the corresponding link invariant of ambient isotopy may be obtained from $`(\stackrel{ˇ}{R},S)`$ by the scalings $`\sigma =(X_1^1X_2)^{\frac{1}{2}}\stackrel{ˇ}{R}`$ and $`C=(X_1^1X_2^1)^{\frac{1}{2}}S`$, where $`X_1S_c^c(\stackrel{ˇ}{R})_{c1}^{c1}`$ and $`X_2S_c^c(\stackrel{ˇ}{R}^1)_{c1}^{c1}`$. #### 2.3.1 Negative Handles, Caps and Cups Demanding that our model parameters satisfy R0 (ambient isotopy in the plane) allows us to determine appropriate values for negative handles, caps and cups (see ). Although we *can* evaluate our invariants without these, we describe them here for completeness and backwards compatibility. Firstly, the negative handle $`\overline{C}`$ is simply $`C|_{q\overline{q}}`$. Secondly, although there is some flexibility in the choice of suitable caps $`\mathrm{\Omega }^\pm `$ and cups $`\mathrm{}^\pm `$, in fact it is natural to choose them to be the square roots of the handles $`C^\pm `$: $$\mathrm{\Omega }^\pm =\mathrm{}^\pm =(C^\pm )^{\frac{1}{2}},$$ (8) taking the positive square root by convention. Note that these choices further improve those of our previous work by increasing the symmetries between the diagram components. Satisfaction of R0 is described in Figure 3, that is, we demand: $$\overline{\mathrm{\Omega }}_{bc}\mathrm{}^{ca}=\delta _b^a=\overline{\mathrm{}}^{ac}\mathrm{\Omega }_{cb}.$$ (9) The definition (8) ensures that (9) is satisfied. In fact, the LHS and RHS of (9) are actually equivalent, hence one is redundant. Again, reversing the orientations of the strings in Figure 3 yields no new constraints. ### 2.4 $`q^{h_\rho }`$ for $`U_q[gl(m|n)]`$ Here, we determine the form that $`q^{h_\rho }`$ takes in $`U_q[gl(m|n)]`$, in terms of Cartan generators. Recall that for any particular representation $`\pi `$, our state model requires (a grading-stripped version of) $`S=\pi (q^{2h_\rho })`$, and this may be obtained by substitution of the appropriate matrix elements into the expression for $`q^{2h_\rho }`$. Initially, we shall work with $`gl(m|n)`$. To this end, let $`H`$ be the Cartan subalgebra of $`gl(m|n)`$, with dual the root space $`H^{}`$. A basis for $`H^{}`$ is given by the *fundamental weights* $`\{\epsilon _i\}_{i=1}^{m+n}`$, which are elementary unit vectors of $`m+n`$ components, with $`1`$ in position $`i`$ and $`0`$ elsewhere. On $`H^{}`$, we have the following invariant bilinear form $`(,):H^{}\times H^{}`$: $$(\epsilon _i,\epsilon _j)()^{[i]}\delta _{ij},$$ (10) and as $`H`$ and $`H^{}`$ are dual, we of course have the form: $$E_{}^{j}{}_{j}{}^{}(\epsilon _i)\delta _{ij},$$ (11) for $`gl(m|n)`$ Cartan generators $`E_{}^{j}{}_{j}{}^{}`$, $`j=1,\mathrm{},m+n`$. To the $`gl(m|n)`$ root $`\epsilon _i\epsilon _j`$, there corresponds a $`gl(m|n)`$ Chevalley generator $`E_{}^{i}{}_{j}{}^{}`$, and we assign a grading and a sign to the roots in accordance with those of these generators. In terms of these, $`gl(m|n)`$ has the following simple, positive roots: $$\alpha _i\epsilon _i\epsilon _{i+1},i=1,\mathrm{},m+n1,$$ (12) in the sense that these form a basis for $`H^{}`$. Apart from the single odd root $`\alpha _m`$, the simple positive roots are all even. (Of various choices for superalgebra root systems, this *distinguished root system* is unique in containing only one odd root.) Where $`\mathrm{\Delta }^+`$ is the set of *all* positive roots, and $`\gamma `$ denotes the grading of the root $`\gamma `$, we define $`\rho `$ as the graded half sum of all positive roots: $`\rho \frac{1}{2}_{\gamma \mathrm{\Delta }^+}()^{[\gamma ]}\gamma `$. Explicitly, for $`gl(m|n)`$, we have \[7, p6207\]: $`\rho =\frac{1}{2}{\displaystyle \underset{i=1}{\overset{m}{}}}(mn2i+1)\epsilon _i+\frac{1}{2}{\displaystyle \underset{i=m+1}{\overset{m+n}{}}}(3m+n2i+1)\epsilon _i,`$ although we will not actually require this form. We are actually interested in $`h_\rho gl(m|n)`$, defined to satisfy: $$h_\rho (\alpha _i)(\rho ,\alpha _i),\alpha _i,$$ (13) where we intend (11) on the LHS and (10) on the RHS. From the definition of $`\rho `$: $`(\rho ,\alpha _i)=\frac{1}{2}(\alpha _i,\alpha _i)\stackrel{(\text{12})}{=}\frac{1}{2}(\epsilon _i\epsilon _{i+1},\epsilon _i\epsilon _{i+1})\stackrel{(\text{10})}{=}\frac{1}{2}[()^{[i]}+()^{[i+1]}].`$ (14) As $`h_\rho `$ is a Cartan element of $`gl(m|n)`$, we may express it as a linear combination of Cartan generators $`E_{}^{i}{}_{i}{}^{}`$; viz for some undetermined scalar coefficients $`\beta _i`$, we may set: $`h_\rho =_{i=1}^{m+n}\beta _iE_{}^{i}{}_{i}{}^{}`$. Substituting this into (13) yields: $`h_\rho (\alpha _i)={\displaystyle \underset{j=1}{\overset{m+n}{}}}\beta _jE_{}^{j}{}_{j}{}^{}(\epsilon _i\epsilon _{i+1})\stackrel{(\text{12},\text{11})}{=}\beta _i\beta _{i+1}.`$ (15) Substituting (14) and (15) into (13), we have: $$\beta _i\beta _{i+1}=\{\begin{array}{cc}\hfill +1& i=1,\mathrm{},m1\hfill \\ \hfill 0& i=m\hfill \\ \hfill 1& i=m+1,\mathrm{},m+n1.\hfill \end{array}$$ (16) For symmetry, selecting $`\beta _m=\theta `$ and substituting backwards and forwards yields: $`\beta _i=\theta +\{\begin{array}{cc}mi\hfill & i=1,\mathrm{},m\hfill \\ i(m+1)\hfill & i=m+1,\mathrm{},m+n,\hfill \end{array}`$ therefore: $`h_\rho `$ $`=`$ $`{\displaystyle \underset{i=1}{\overset{m+n}{}}}\beta _iE_{}^{i}{}_{i}{}^{}={\displaystyle \underset{i=1}{\overset{m}{}}}(\theta +mi)E_{}^{i}{}_{i}{}^{}+{\displaystyle \underset{i=m+1}{\overset{m+n}{}}}(\theta +i(m+1))E_{}^{i}{}_{i}{}^{}`$ $`=`$ $`\theta C_1+{\displaystyle \underset{i=1}{\overset{m}{}}}(mi)E_{}^{i}{}_{i}{}^{}+{\displaystyle \underset{i=m+1}{\overset{m+n}{}}}(i(m+1))E_{}^{i}{}_{i}{}^{},`$ where $`C_1_{i=1}^{m+n}E_{}^{i}{}_{i}{}^{}`$ is the first-order Casimir element of $`gl(m|n)`$. This shows us that $`h_\rho `$ is only determined up to an additive constant.<sup>5</sup><sup>5</sup>5For $`sl(m|n)`$ and $`sl(n)`$, $`h_\rho `$ is actually unique. $`C_1`$ also satisfies $`C_1(\alpha _i)=0,\alpha _i`$. In passing from $`gl(m|n)`$ to $`U_q[gl(m|n)]`$, we pass from $`h_\rho `$ to $`q^{h_\rho }`$, hence we have: $`q^{h_\rho }`$ $`=`$ $`q^{\theta C_1}q^{_{i=1}^m(mi)E_{}^{i}{}_{i}{}^{}}q^{_{i=m+1}^{m+n}(i(m+1))E_{}^{i}{}_{i}{}^{},}`$ $`=`$ $`(q^{C_1})^\theta {\displaystyle \underset{i=1}{\overset{m}{}}}\left(q^{E_{}^{i}{}_{i}{}^{}}\right)^{mi}{\displaystyle \underset{i=m+1}{\overset{m+n}{}}}\left(q^{E_{}^{i}{}_{i}{}^{}}\right)^{i(m+1)}`$ $`=`$ $`(q^{C_1})^\theta {\displaystyle \underset{i=1}{\overset{m}{}}}K_i^{mi}{\displaystyle \underset{i=m+1}{\overset{m+n}{}}}K_i^{(m+1)i},`$ where we have reminded ourselves of the definition $`K_iq^{()^{[i]}E_{}^{i}{}_{i}{}^{}}`$. Thus, $`q^{h_\rho }`$ is only determined up to an arbitrary multiplicative constant. Selecting $`\theta =0`$, we declare the resulting product to be the standard $`q^{h_\rho }`$. For arbitrary $`m,n`$, we have: $$q^{h_\rho }=K_1^{m1}K_2^{m2}\mathrm{}K_{m1}^1K_m^0K_{m+1}^0K_{m+2}^1\mathrm{}K_{m+n}^{(n1)},$$ (18) where of course $`K_i^0`$ is the $`U_q[gl(m|n)]`$ identity element. For our state models we require $`S=\pi (q^{2h_\rho })`$. To construct $`S`$, it suffices to compute matrix elements for the $`U_q[gl(m|n)]`$ Cartan generators $`K_i`$, and insert (appropriate powers of) these into (18), finally stripping the grading from $`S`$. In , we described the automation of the construction of $`\stackrel{ˇ}{R}`$ corresponding to the $`U_q[gl(m|1)]`$ representations $`(\dot{0}_m|\alpha )`$, for arbitrary $`m`$, and obtained explicit $`\stackrel{ˇ}{R}`$ for $`m=1,2,3,4`$. Explicit matrix elements for the $`K_i`$ are obtained as a byproduct of that construction, facilitating the evaluation of $`S`$. What is particularly interesting about this work is that the entire process, from the construction of the underlying representations , to the scaling of the state model parameters, to the final evaluations of the polynomials, has been automated. This represents a step forward in computational power in knot theory. ## 3 The quantum link invariants $`LG^{m,n}`$ Having described the construction of state models for arbitrary finite dimensional highest weight $`U_q[gl(m|n)]`$ representations $`\mathrm{\Lambda }`$, we now restrict our attention to the case: $`\mathrm{\Lambda }=(\dot{0}_m|\dot{\alpha }_n)(0,\mathrm{},0|\alpha ,\mathrm{},\alpha ),`$ and the resulting invariants $`LG^{m,n}`$. Evaluation of $`LG^{m,n}`$ for any particular link follows from that for $`LG^2`$, described in our previous work . Below, we make a few comments on the properties of $`LG^{m,n}`$, before describing in §4 some computational issues and evaluations for $`LG^3`$ and $`LG^4`$. ### 3.1 Checking the QYBE and applying the Matveev $`\mathrm{\Delta }`$$``$ test To be certain that we have made no errors in our computations, we check that our braid generator $`\sigma `$ satisfies the (quantum) Yang–Baxter equation. The code used to construct the tensors $`Z_K`$ is immediately adaptable to such a test. If $`Z`$ is the same for the braids $`\sigma _1\sigma _2\sigma _1`$ and $`\sigma _2\sigma _1\sigma _2`$, then our braid generator satisfies the QYBE. This is depicted in Figure 4. The same framework allows us to carry out a simple sufficiency check to determine if a link invariant associated with some R matrix solution of the QYBE will be trivial.<sup>6</sup><sup>6</sup>6This test is known to be a sufficient (but perhaps not a necessary) test of triviality – it doesn’t even guarantee the *existence* of an invariant. Matveev (see also ) introduced a ‘delta unknotting operation’ (which we call the Matveev $`\mathrm{\Delta }`$$``$ test), and proved that any knot can be transformed to the unknot by using only this operation. In our tensor language, if $`Z`$ fails to distinguish $`\sigma _1\overline{\sigma }_2\sigma _1`$ and $`\sigma _2\overline{\sigma }_1\sigma _2`$, then the associated invariant will be trivial, as a series of exchanges of crossings of this form is always sufficient to convert any links to the unknot. Matveev’s test is depicted in Figure 5. Both these tests have been satisfactorily carried out for our various braid generators $`\sigma `$, viz each $`\sigma `$ satisfies the QYBE and the invariant built from it is not necessarily trivial. ### 3.2 Behaviour of $`LG^{m,n}`$ under inversion of $`q`$ Let $`K^{}`$ denote the reflection of a link $`K`$. In , we showed that $`LG_K^{}^2=LG_K^2|_{q\overline{q}}`$. This result immediately carries over to $`LG^{m,n}`$, and means that if $`LG_K^{m,n}`$ is palindromic in $`q`$ (i.e. invariant under the inversion $`q\overline{q}`$), then $`LG^{m,n}`$ cannot distinguish the chirality of $`K`$. Examples illustrating that $`LG^2`$ can distinguish the chirality of all prime knots of up to $`10`$ crossings demonstrate that $`LG^{m,n}`$ can indeed sometimes distinguish chirality, although counterexamples are expected to exist. ### 3.3 $`LG^{m,n}`$ doesn’t detect mutation Theorem 5 of shows that quantum invariants based on R matrices where the orthogonal decomposition of $`VV`$ contains no multiplicities will not distinguish mutants. The extension of this result to quantum superalgebras is straightforward, and as our invariants $`LG^{m,n}`$ are indeed based on representations of this type , they will not distinguish mutants. ### 3.4 Behaviour of $`LG^{m,n}`$ under representation duality In \[6, Proposition 3.2\], we showed that link invariants derived from irreducible representations of quantum (super)algebras are unable to detect knot inversion, as such invariants are necessarily equivalent to invariants constructed from their dual representations. Let us determine what this means for $`LG^{m,n}`$. Let $`VV_\mathrm{\Lambda }`$ be the module associated with $`\pi \pi _\mathrm{\Lambda }`$, viz $`V`$ has a highest weight vector $`v_+`$, of weight $`\mathrm{\Lambda }`$. The corresponding lowest weight vector of $`V`$ is obtained by the combined action of all the odd lowering operators: $`E_{}^{m+j}{}_{i}{}^{}`$ on $`v_+`$, where the product is over all $`i=1,\mathrm{},m`$ and $`j=1,\mathrm{},n`$. The action of $`E_{}^{m+j}{}_{i}{}^{}`$ on a weight vector lowers its weight by $`\epsilon _i\epsilon _{m+j}`$, viz: $`(0,\mathrm{},0,1,0,\mathrm{},0|\mathrm{\hspace{0.33em}0},\mathrm{},0,1,0,\mathrm{},0).`$ The resulting lowest weight vector of $`V`$ thus has weight $`\overline{\mathrm{\Lambda }}`$: $`\overline{\mathrm{\Lambda }}`$ $`=`$ $`\mathrm{\Lambda }_{ij}(\epsilon _i\epsilon _{m+j})=\mathrm{\Lambda }n_{i=1}^m\epsilon _im_{j=1}^n\epsilon _{m+j}`$ $`=`$ $`(n,\mathrm{},n|\alpha +m,\mathrm{},\alpha +m).`$ The dual of $`V`$ is labeled $`V^{}`$, and naturally has highest weight $`\overline{\mathrm{\Lambda }}`$: $`\overline{\mathrm{\Lambda }}=(n,\mathrm{},n|\alpha m,\mathrm{},\alpha m),`$ but $`V^{}`$ is equivalent to the module of highest weight $`\mathrm{\Lambda }^{}`$: $`\mathrm{\Lambda }^{}=(0,\mathrm{},0|\alpha (mn),\mathrm{},\alpha (mn)),`$ hence we may regard the representations $`\mathrm{\Lambda }`$ and $`\mathrm{\Lambda }^{}`$ as duals. Thus, at least up to a scalar multiple, we expect $`LG^{m,n}`$ to be invariant under the transformation $`\alpha \alpha (mn)`$, equivalently in the more symmetric form: $`\alpha +\frac{mn}{2}\alpha \frac{mn}{2}`$, viz: $`LG^{m,n}(q,q^{\alpha +\frac{mn}{2}})=LG^{m,n}(q,q^{\alpha \frac{mn}{2}}).`$ Thus, if we define: $$pq^{\alpha +\frac{mn}{2}},$$ (19) we have, again, up to a scalar multiple, the symmetry: $$LG^{m,n}(q,p)=LG^{m,n}(q,\overline{p}).$$ (20) Experiments show that the scalar multiple is always $`\pm 1`$, and, for knots, always $`1`$. Where $`\overline{K}`$ is the inverse of a knot $`K`$, inspection of diagram components shows that $`LG_{\overline{K}}^{m,n}(q,p)=LG_K^{m,n}(q,\overline{p})`$, hence (20) shows that $`LG^{m,n}`$ is unable to detect the inversion of knots. Experimentally (setting $`n=1`$), we find that the only time the “$``$” sign actually appears is for odd $`m`$ and links of $`2`$ components. That this is true for case $`m=1`$ (which is in fact the Alexander–Conway polynomial) is well-known . These results are exemplified in our previous work for $`LG^2`$. Lastly, we often wish to eliminate $`\alpha `$ from expressions of the form $`q^{x\alpha +y}`$, to express them in terms of $`p`$ and $`q`$ alone. Using (19), we have: $`q^{x\alpha +y}=p^xq^{yx(\frac{mn}{2})}.`$ ### 3.5 $`LG^{m,n}`$ of split links Recall that we define $`LG_K^{m,n}`$ as a $`(1,1)`$ tangle invariant, obtained for a link $`K`$ as the first component of the diagonal tensor (scalar multiple of the identity) $`T_K`$. We do this as the closed form (i.e. the $`(0,0)`$ tangle form) always evaluates to zero (cf. the ADO invariant ). To see this, begin by observing that the value of our state model on $`0_1`$ (i.e. the unknot, an isolated loop) as a $`(0,0)`$ tangle is zero, as $`_aC_a^a=0`$. This follows from the fact that for the $`U_q[gl(m|n)]`$ superalgebras, the $`q`$-superdimension of typical representations (defined by $`\mathrm{str}[\pi (q^{2h_\rho })]`$) is always identically zero . As $`S`$ is a grading-stripped version of the exponential of the Cartan element $`\pi (q^{2h_\rho })`$, we necessarily have $`\mathrm{tr}(S)=0`$, hence $`\mathrm{tr}(C)=0`$. Multiplying these results by the scalar in $`T_K`$ yields the result. Now, let $`K=K_1K_2`$ be the split (i.e. disconnected, separated) union of links $`K_1`$ and $`K_2`$, and say that we are trying to evaluate the $`(1,1)`$ tangle form using a string of $`K_1`$. The construction of $`LG_K^{m,n}`$ means that at some point of contracting $`Z_K`$ to $`T_K`$, we close the final string of $`K_2`$, and at this stage our tensor becomes zero throughout, thus $`T_K`$ is zero. Thus, as disconnected multicomponent links represented by $`(1,1)`$ tangles necessarily include a closed component, we have proven: ###### Theorem 2 $`LG_K^{m,n}=0`$ for disconnected multicomponent links $`K`$. ## 4 Computational issues in evaluating $`LG^{m,n}`$ ### 4.1 Various sets of computational variables The representation of the braid generator $`\sigma `$ obtained from the representation theory contains algebraic expressions in variables $`q`$ and $`\alpha `$, including many $`q`$ brackets. This form is readable to human eyes, but can be improved upon for machine consumption. We shall call $`\{q,\alpha \}`$ the rep(resentation) variables. From (19), we see that our link invariants are naturally expressed in terms of $`q`$ and $`pq^{\alpha +\frac{mn}{2}}`$; so we initially make this change of variables in the internal representation of the braid generator and the positive handle. This action replaces all the $`q`$ brackets, which contained $`\alpha `$. The resulting braid generator contains rational expressions in variables $`q^{\frac{1}{2}}`$ and $`p`$. To simplify the vulgar fractions within the exponents, we define a new variable to be used internally: $`Qq^{\frac{1}{2}}`$. In some sense, the resulting braid generator is now optimally literate, and we use this form to accrete tensors to build $`Z_K`$, and also to check the QYBE and the Matveev $`\mathrm{\Delta }`$$``$ test. We shall call $`\{Q,p\}`$ the int(ernal) variables, and to convert from rep to int variables, we shall invoke *in order* the following rules: $`\left\{q^{x\alpha +y}p^xq^{yx(\frac{mn}{2})},qQ^2\right\},`$ where $`x,y`$. We occasionally have an interest in the inverse transformation to convert from int to rep variables, and for this we shall invoke *in order* the following rules: $`\left\{Qq^{\frac{1}{2}},pq^{\alpha +\frac{mn}{2}},(q^{x\alpha +y}\overline{q}^{x\alpha +y})(q\overline{q})[x\alpha +y]_q\right\},`$ where handling the last of these rules typically requires some care. Sometimes, we must invert the int variables, for example in computing the inverse braid generator $`\overline{\sigma }`$. We have the rules: $`\left\{Q\overline{Q},p\overline{p}\right\}.`$ Finally, extracting the first component of $`T_K`$ thus yields an expression int variables. We must then expand $`Qq^2`$. Furthermore, we discover that $`LG_K^{m,n}`$ is actually an invariant in $`p^2`$ not just $`p`$, so we define $`P=p^2`$ to reduce things a little. We shall call $`\{q,P\}`$ the L(ink) I(nvariant) variables, and to convert from int to LI variables, we shall invoke the following rules: $`\left\{Qq^{\frac{1}{2}},pP^{\frac{1}{2}}\right\}.`$ Parameters used for the state models for $`LG^m`$ for $`m=1,2,3,4`$ are presented in Appendix A. ### 4.2 Explicit construction of $`S`$ From (18), we have for $`U_q[gl(m|n)]`$ that $`S\pi (q^{2h_\rho })`$ is: $`S`$ $`=`$ $`\pi (K_1)^{2(m1)}\pi (K_2)^{2(m2)}\mathrm{}\pi (K_{m1})^2`$ (21) $`\pi (K_{m+2})^2\pi (K_{m+3})^4\mathrm{}\pi (K_{m+n})^{2(n1)}.`$ Setting $`n=1`$ in (21), we have: $`S=\pi _\mathrm{\Lambda }(K_1)^{2(m1)}\pi _\mathrm{\Lambda }(K_2)^{2(m2)}\mathrm{}\pi _\mathrm{\Lambda }(K_{m1})^2.`$ To illustrate, for the $`U_q[gl(2|1)]`$ case, we have $`h_\rho =E_{}^{1}{}_{1}{}^{}`$, hence $`q^{h_\rho }=K_1`$, so $`S=\pi (K_1)^2`$. This contrasts with the choice of $`\theta =1`$ made in , which yields $`h_\rho =E_{}^{2}{}_{2}{}^{}E_{}^{3}{}_{3}{}^{}`$, viz $`q^{h_\rho }=q^{E_{}^{2}{}_{2}{}^{}}q^{E_{}^{3}{}_{3}{}^{}}=K_2^1K_3`$, so $`S=\pi (K_2)^2\pi (K_3)^2`$. ### 4.3 Illustrative examples of $`LG^m`$ At present, we are able to compute state model parameters for $`LG^{m,1}LG^m`$ only, as we have not yet computed $`\stackrel{ˇ}{R}`$ or matrix elements for the $`K_i`$ for cases $`n1`$. For the cases $`m=1,2,3,4`$, we are able to make the following comments. * $`LG^1`$ is the Alexander–Conway polynomial in variable $`Pq^{2\alpha }`$. This is a well-known result, cf. . * Evaluations for $`LG^2`$ for all prime knots of up to $`10`$ crossings have been reported in . In that paper, we claimed that $`LG^m`$ for $`m>2`$ was essentially incomputable due to vast memory requirements of the tensors $`Z_K`$; but we have since made some headway in this by adapting our code to recognise the sparsity of these tensors; doing the symbolic equivalent of what is called “sparse matrix multiplication” in numerical linear algebra. This change comes at a cost of more lines of interpreted code, but is still an improvement in algorithmic efficiency. It also results in an increase in the speed of computation for $`LG^2`$, and facilitates its evaluation from braid presentations of $`6`$ strings, something not previously feasible. * Evaluations for $`LG^3`$ and $`LG^4`$ for various links are presented in Appendix B. Those lists are quite brief, and only include some links of braid index at most $`3`$. Our current computational method requires too much memory for us to extend our tables of polynomials any further. Of some interest is the rate of growth in exponent of the polynomials with $`m`$ for a particular link. For example, we have the following results for the trefoil knot $`3_1`$ and the figure eight knot $`4_1`$: $`\begin{array}{ccc}\hfill LG_{3_1}^1& =& (1)\hfill \\ & & +(\overline{P}^1+P^1)(+1)\hfill \\ & & \\ \hfill LG_{3_1}^2& =& (1+2q^2)\hfill \\ & & (\overline{P}^1+P^1)(q+q^3)\hfill \\ & & +(\overline{P}^2+P^2)(q^2)\hfill \\ & & \\ \hfill LG_{3_1}^3& =& (q^2+2q^4+3q^6+q^8)\hfill \\ & & +(\overline{P}^1+P^1)(q^2+2q^4+2q^6+q^8)\hfill \\ & & (\overline{P}^2+P^2)(q^4+q^6+q^8)\hfill \\ & & +(\overline{P}^3+P^3)(q^6)\hfill \\ & & \\ \hfill LG_{3_1}^4& =& (q^4+2q^6+4q^8+4q^{10}+5q^{12}+2q^{14}+q^{16})\hfill \\ & & (\overline{P}^1+P^1)(q^5+2q^7+4q^9+4q^{11}+3q^{13}+2q^{15})\hfill \\ & & +(\overline{P}^2+P^2)(q^6+2q^8+2q^{10}+3q^{12}+q^{14}+q^{16})\hfill \\ & & (\overline{P}^3+P^3)(q^9+q^{11}+q^{13}+q^{15})\hfill \\ & & +(\overline{P}^4+P^4)(q^{12})\hfill \end{array}`$ $`\begin{array}{ccc}\hfill LG_{4_1}^1& =& (3)\hfill \\ & & (\overline{P}^1+P^1)(1)\hfill \\ & & \\ \hfill LG_{4_1}^2& =& (2\overline{q}^2+7+2q^2)\hfill \\ & & (\overline{P}^1+P^1)(3\overline{q}+3q)\hfill \\ & & +(\overline{P}^2+P^2)(1)\hfill \\ & & \\ \hfill LG_{4_1}^3& =& (5\overline{q}^4+9\overline{q}^2+17+9q^2+5q^4)\hfill \\ & & (\overline{P}^1+P^1)(2\overline{q}^4+8\overline{q}^2+10+8q^2+2q^4)\hfill \\ & & +(\overline{P}^2+P^2)(3\overline{q}^2+3+3q^2)\hfill \\ & & (\overline{P}^3+P^3)(1)\hfill \end{array}`$ ## 5 Further work The current work is part of a larger program to automate the construction of more general quantum link invariants. A few comments on the direction of this program are in order. * In this paper, the limits of our method of evaluation have been reached,<sup>7</sup><sup>7</sup>7The material has also been applied to the evaluation of ‘$`N`$-Jones’ polynomials $`V^N`$. These are the quantum link invariants associated with the $`N`$ dimensional representations of $`U_q[sl(2)]`$. In the language of , they are monochromatic versions of coloured Jones polynomials of order $`N`$. The limit to computation for these invariants for prime knots of up to $`10`$ crossings is around $`N=4`$, although we can calculate $`V_{3_1}^{13}`$. and a more efficient method of evaluation is required. A promising candidate involves chasing through braids one crossing at a time, accumulating only an $`N\times N`$ matrix (where $`N=\mathrm{dim}(\mathrm{V})`$) of polynomials at each step. That method requires foreknowledge of the decomposition of $`\stackrel{ˇ}{R}`$ into the canonical form $`\stackrel{ˇ}{R}=_ia_ib_i`$, and this is already available for $`U_q[sl(2)]`$ and $`U_q[gl(1|1)]`$. It is applicable to links of any number of crossings and components, and is really only limited by $`N`$, although much less strongly than our current method. In particular, it is not dependent on the string index of braid presentations. * Moreover, the construction of more general quantum link invariants requires a more general approach to construction of underlying R matrices. The current method exploits explicit knowledge of the decomposition of the tensor product of the underlying module, but this is *not* generally known. Alternatively, it is also possible to construct explicit R matrices from knowledge of the universal (i.e. representation-independent) R matrix and the matrix elements of the underlying representation. As we have to hand details of the universal R matrices for arbitrary quantum (super)algebras (albeit in a somewhat abstract form), and some knowledge of a process to construct the matrix elements, it is eminently possible to construct many more R matrices. * Lastly, we are limited by our use of braids, for which we have systematic tables only for the first $`249`$ prime knots of up to $`10`$ crossings. As of 1998, Dowker codes for all the $`1,701,936`$ prime knots of up to $`16`$ crossings have been enumerated , and our not being able to access them is a sad thing. As we don’t have the implementation of an algorithm that allows us to map these Dowker codes to braids, it is attractive to try to adapt new material to accept Dowker codes as input. The converse to this is that our new invariants $`LG^{m,n}`$ are well suited to extending those tables, as they distinguish many more knots than other polynomial invariants. ## Acknowledgements I am grateful to Jon Links of The University of Queensland, Australia, for advice on how the deduce the natural choice of variables to describe $`LG^{m,n}`$, and other worthwhile discussions. My research at Kyoto University is funded by a Postdoctoral Fellowship for Foreign Researchers (# P99703), provided by the Japan Society for the Promotion of Science. Dōmo arigatō gozaimashita! ## Appendix A State model parameters Below, we list state model parameters for $`LG^m`$, for $`m=1,2,3,4`$. To improve literacy, we have written $`[X]`$ for $`[X]_q`$, $`\overline{X}`$ for $`X^1`$, for various $`X`$, and $`\mathrm{\Delta }=q\overline{q}`$. Horizontal lines divide tensor components into symmetry classes. ### Parameters for $`LG^1`$ The braid generator $`\sigma `$ has $`5`$ nonzero components: $`q^\alpha \left\{\begin{array}{c}e_{11}^{11}\end{array}\right\},q^\alpha \left\{\begin{array}{c}e_{22}^{22}\end{array}\right\},\mathrm{\Delta }[\alpha ]\left\{\begin{array}{c}e_{21}^{21}\end{array}\right\},1\left\{\begin{array}{c}e_{21}^{12}\\ e_{12}^{21}\end{array}\right\},`$ and the left handle $`C`$ has $`2`$ diagonal components: $`C=q^\alpha \left\{\begin{array}{c}+e_1^1\hfill \\ e_2^2\hfill \end{array}\right\},`$ using the scaling factors: $`\kappa _\sigma =q^\alpha ,\kappa _C=q^\alpha .`$ ### Parameters for $`LG^2`$ The braid generator $`\sigma `$ has $`26`$ nonzero components: $`q^{2\alpha }\left\{\begin{array}{c}e_{11}^{11}\end{array}\right\},1\left\{\begin{array}{c}e_{22}^{22},e_{33}^{33}\end{array}\right\},q^{2\alpha +2}\left\{\begin{array}{c}e_{44}^{44}\end{array}\right\},`$ $`\mathrm{\Delta }q^\alpha [\alpha ]\left\{\begin{array}{c}e_{21}^{21},e_{31}^{31}\end{array}\right\},\mathrm{\Delta }^2q[\alpha ][\alpha +1]\left\{\begin{array}{c}e_{41}^{41}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +1}[\alpha +1]\left\{\begin{array}{c}e_{42}^{42},e_{43}^{43}\end{array}\right\},\mathrm{\Delta }q\left\{\begin{array}{c}e_{32}^{32}\end{array}\right\},`$ $`1\left\{\begin{array}{c}e_{41}^{14}\\ e_{14}^{41}\end{array}\right\},q\left\{\begin{array}{c}e_{32}^{23}\\ e_{23}^{32}\end{array}\right\},q^\alpha \left\{\begin{array}{c}e_{21}^{12},e_{31}^{13}\\ e_{12}^{21},e_{13}^{31}\end{array}\right\},q^{\alpha +1}\left\{\begin{array}{c}e_{42}^{24},e_{43}^{34}\\ e_{24}^{42},e_{34}^{43}\end{array}\right\},`$ $`\mathrm{\Delta }q[\alpha ]^{\frac{1}{2}}[\alpha +1]^{\frac{1}{2}}\{\overline{q}^{\frac{1}{2}}\left\{\begin{array}{c}e_{41}^{23}\\ e_{23}^{41}\end{array}\right\},+q^{\frac{1}{2}}\left\{\begin{array}{c}e_{41}^{32}\\ e_{32}^{41}\end{array}\right\}\},`$ and the left handle $`C`$ has $`4`$ diagonal components: $`C=q^{2\alpha 1}\left\{\begin{array}{c}+q\left\{e_1^1\right\}\hfill \\ \{qe_2^2,\overline{q}e_3^3\}\hfill \\ +\overline{q}\left\{e_4^4\right\}\hfill \end{array}\right\},`$ using the scaling factors: $`\kappa _\sigma =q^{2\alpha },\kappa _C=q^{2\alpha }.`$ ### Parameters for $`LG^3`$ The braid generator $`\sigma `$ has $`139`$ nonzero components: $`q^{3\alpha }\left\{\begin{array}{c}e_{11}^{11}\end{array}\right\},q^\alpha \left\{\begin{array}{c}e_{22}^{22},e_{33}^{33},e_{44}^{44}\end{array}\right\},q^{\alpha +2}\left\{\begin{array}{c}e_{55}^{55},e_{66}^{66},e_{77}^{77}\end{array}\right\},q^{3\alpha +6}\left\{\begin{array}{c}e_{88}^{88}\end{array}\right\},`$ $`\mathrm{\Delta }q^{2\alpha }[\alpha ]\left\{\begin{array}{c}e_{21}^{21},e_{31}^{31},e_{41}^{41}\end{array}\right\},\mathrm{\Delta }^2q^{\alpha +1}[\alpha ][\alpha +1]\left\{\begin{array}{c}e_{51}^{51},e_{61}^{61},e_{71}^{71}\end{array}\right\},`$ $`\mathrm{\Delta }q^{2\alpha +4}[\alpha +2]\left\{\begin{array}{c}e_{87}^{87},e_{86}^{86},e_{85}^{85}\end{array}\right\},\mathrm{\Delta }^2q^{\alpha +3}[\alpha +1][\alpha +2]\left\{\begin{array}{c}e_{84}^{84},e_{83}^{83},e_{82}^{82}\end{array}\right\},`$ $`\mathrm{\Delta }q[\alpha +1]\left\{\begin{array}{c}e_{52}^{52},e_{53}^{53},e_{62}^{62},e_{64}^{64},e_{73}^{73},e_{74}^{74}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +1}\left\{\begin{array}{c}e_{32}^{32},e_{42}^{42},e_{43}^{43}\end{array}\right\},\mathrm{\Delta }q^{\alpha +3}\left\{\begin{array}{c}e_{65}^{65},e_{75}^{75},e_{76}^{76}\end{array}\right\},`$ $`\mathrm{\Delta }q^2[\alpha +1]\left\{\begin{array}{c}\mathrm{\Delta }e_{63}^{63},q(\overline{q}^2q^2)e_{72}^{72}\end{array}\right\},\mathrm{\Delta }^3q^3[\alpha ][\alpha +1][\alpha +2]\left\{\begin{array}{c}e_{81}^{81}\end{array}\right\},`$ $`q^{2\alpha }\left\{\begin{array}{c}e_{21}^{12},e_{31}^{13},e_{41}^{14}\\ e_{12}^{21},e_{13}^{31},e_{14}^{41}\end{array}\right\},q^\alpha \left\{\begin{array}{c}e_{51}^{15},e_{61}^{16},e_{71}^{17}\\ e_{15}^{51},e_{16}^{61},e_{17}^{71}\end{array}\right\},`$ $`q^{2\alpha +4}\left\{\begin{array}{c}e_{58}^{85},e_{68}^{86},e_{78}^{87}\\ e_{85}^{58},e_{86}^{68},e_{87}^{78}\end{array}\right\},q^{\alpha +2}\left\{\begin{array}{c}e_{28}^{82},e_{38}^{83},e_{48}^{84}\\ e_{82}^{28},e_{83}^{38},e_{84}^{48}\end{array}\right\},`$ $`q^{\alpha +1}\left\{\begin{array}{c}e_{32}^{23},e_{42}^{24},e_{43}^{34}\\ e_{23}^{32},e_{24}^{42},e_{34}^{43}\end{array}\right\},q^{\alpha +3}\left\{\begin{array}{c}e_{65}^{56},e_{75}^{57},e_{76}^{67}\\ e_{56}^{65},e_{57}^{75},e_{67}^{76}\end{array}\right\},`$ $`q\left\{\begin{array}{c}e_{52}^{25},e_{62}^{26},e_{53}^{35},e_{73}^{37},e_{64}^{46},e_{74}^{47}\\ e_{25}^{52},e_{26}^{62},e_{35}^{53},e_{37}^{73},e_{46}^{64},e_{47}^{74}\end{array}\right\},q^2\left\{\begin{array}{c}e_{72}^{27},e_{54}^{45},e_{63}^{36}\\ e_{27}^{72},e_{45}^{54},e_{36}^{63}\end{array}\right\},1\left\{\begin{array}{c}e_{81}^{18}\\ e_{18}^{81}\end{array}\right\},`$ $`\mathrm{\Delta }q^2\{+1\left\{\begin{array}{c}e_{63}^{45}\\ e_{45}^{63}\end{array}\right\},q\left\{\begin{array}{c}e_{72}^{45}\\ e_{45}^{72}\end{array}\right\},+1\left\{\begin{array}{c}e_{72}^{36}\\ e_{36}^{72}\end{array}\right\}\},`$ $`\mathrm{\Delta }q^3[\alpha +1]\{+\overline{q}\left\{\begin{array}{c}e_{63}^{54}\\ e_{54}^{63}\end{array}\right\},1\left\{\begin{array}{c}e_{72}^{54}\\ e_{54}^{72}\end{array}\right\},+q\left\{\begin{array}{c}e_{72}^{63}\\ e_{63}^{72}\end{array}\right\}\},`$ $`\mathrm{\Delta }q^2[\alpha ]^{\frac{1}{2}}[\alpha +2]^{\frac{1}{2}}\{\overline{q}\left\{\begin{array}{c}e_{81}^{27}\\ e_{27}^{81}\end{array}\right\},+1\left\{\begin{array}{c}e_{81}^{36}\\ e_{36}^{81}\end{array}\right\},q\left\{\begin{array}{c}e_{81}^{45}\\ e_{45}^{81}\end{array}\right\}\},`$ $`\mathrm{\Delta }q^{\alpha +1}[\alpha ]^{\frac{1}{2}}[\alpha +1]^{\frac{1}{2}}\{\overline{q}^{\frac{1}{2}}\left\{\begin{array}{c}e_{51}^{23},e_{61}^{24},e_{71}^{34}\\ e_{23}^{51},e_{24}^{61},e_{34}^{71}\end{array}\right\},+q^{\frac{1}{2}}\left\{\begin{array}{c}e_{51}^{32},e_{61}^{42},e_{71}^{43}\\ e_{32}^{51},e_{42}^{61},e_{43}^{71}\end{array}\right\}\},`$ $`\mathrm{\Delta }q^{\alpha +3}[\alpha +1]^{\frac{1}{2}}[\alpha +2]^{\frac{1}{2}}\{\overline{q}^{\frac{1}{2}}\left\{\begin{array}{c}e_{82}^{56},e_{83}^{57},e_{84}^{67}\\ e_{56}^{82},e_{57}^{83},e_{67}^{84}\end{array}\right\},+q^{\frac{1}{2}}\left\{\begin{array}{c}e_{82}^{65},e_{83}^{75},e_{84}^{76}\\ e_{65}^{82},e_{75}^{83},e_{76}^{84}\end{array}\right\}\},`$ $`\mathrm{\Delta }^2q^3[\alpha ]^{\frac{1}{2}}[\alpha +1][\alpha +2]^{\frac{1}{2}}\{+\overline{q}\left\{\begin{array}{c}e_{81}^{54}\\ e_{54}^{81}\end{array}\right\},1\left\{\begin{array}{c}e_{81}^{63}\\ e_{63}^{81}\end{array}\right\},+q\left\{\begin{array}{c}e_{81}^{72}\\ e_{72}^{81}\end{array}\right\}\},`$ and the left handle $`C`$ has $`8`$ diagonal components: $`C=q^{3\alpha 3}\left\{\begin{array}{c}+q^3\left\{e_1^1\right\}\hfill \\ q\{q^2e_2^2,e_3^3,\overline{q}^2e_4^4\}\hfill \\ +\overline{q}\{\overline{q}^2e_7^7,e_6^6,q^2e_5^5\}\hfill \\ \overline{q}^3\left\{e_8^8\right\}\hfill \end{array}\right\},`$ using the scaling factors: $`\kappa _\sigma =q^{3\alpha },\kappa _C=q^{3\alpha }.`$ ### Parameters for $`LG^4`$ The reader will have by now appreciated the recurring patterns in the components of our R matrices. To save space, we introduce a little more notation, which eliminates the $`q`$ brackets altogether. To whit, we write: $`A_i^z`$ $``$ $`[\alpha +i]_{q}^{}{}_{}{}^{z},\mathrm{where}z\{\frac{1}{2},1\},`$ and $`i\{0,1,2,3\}`$. With this notation, the braid generator $`\sigma `$ has $`758`$ nonzero components: $`q^{4\alpha }\left\{\begin{array}{c}e_{1,1}^{1,1}\end{array}\right\},q^2\left\{\begin{array}{c}e_{6,6}^{6,6},e_{7,7}^{7,7},e_{8,8}^{8,8},e_{9,9}^{9,9},e_{10,10}^{10,10},e_{11,11}^{11,11}\end{array}\right\},q^{4\alpha +12}\left\{\begin{array}{c}e_{16,16}^{16,16}\end{array}\right\},`$ $`q^{2\alpha }\left\{\begin{array}{c}e_{2,2}^{2,2},e_{3,3}^{3,3},e_{4,4}^{4,4},e_{5,5}^{5,5}\end{array}\right\},q^{2\alpha +6}\left\{\begin{array}{c}e_{15,15}^{15,15},e_{14,14}^{14,14},e_{13,13}^{13,13},e_{12,12}^{12,12}\end{array}\right\},`$ $`\mathrm{\Delta }q^{3\alpha }A_0\left\{\begin{array}{c}e_{2,1}^{2,1},e_{3,1}^{3,1},e_{4,1}^{4,1},e_{5,1}^{5,1}\end{array}\right\},\mathrm{\Delta }q^{3\alpha +9}A_3\left\{\begin{array}{c}e_{16,12}^{16,12},e_{16,13}^{16,13},e_{16,14}^{16,14},e_{16,15}^{16,15}\end{array}\right\},`$ $`\mathrm{\Delta }q^3\left\{\begin{array}{c}e_{7,6}^{7,6},e_{8,6}^{8,6},e_{8,7}^{8,7},e_{9,6}^{9,6},e_{9,7}^{9,7},e_{10,6}^{10,6},e_{10,8}^{10,8},e_{10,9}^{10,9},e_{11,7}^{11,7},e_{11,8}^{11,8},e_{11,9}^{11,9},e_{11,10}^{11,10}\end{array}\right\},`$ $`\mathrm{\Delta }q^{2\alpha +1}\left\{\begin{array}{c}e_{3,2}^{3,2},e_{4,2}^{4,2},e_{4,3}^{4,3},e_{5,2}^{5,2},e_{5,3}^{5,3},e_{5,4}^{5,4}\end{array}\right\},`$ $`\mathrm{\Delta }q^{2\alpha +7}\left\{\begin{array}{c}e_{13,12}^{13,12},e_{14,12}^{14,12},e_{14,13}^{14,13},e_{15,12}^{15,12},e_{15,13}^{15,13},e_{15,14}^{15,14}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +1}A_1\left\{\begin{array}{c}e_{6,2}^{6,2},e_{6,3}^{6,3},e_{7,2}^{7,2},e_{7,4}^{7,4},e_{8,2}^{8,2},e_{8,5}^{8,5},e_{9,3}^{9,3},e_{9,4}^{9,4},e_{10,3}^{10,3},e_{10,5}^{10,5},e_{11,4}^{11,4},e_{11,5}^{11,5}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +3}(\overline{q}^2q^2)A_1\left\{\begin{array}{c}e_{9,2}^{9,2},e_{10,2}^{10,2},e_{11,2}^{11,2},e_{11,3}^{11,3}\end{array}\right\},`$ $`\mathrm{\Delta }^2q^{2\alpha +1}A_0A_1\left\{\begin{array}{c}e_{6,1}^{6,1},e_{7,1}^{7,1},e_{8,1}^{8,1},e_{9,1}^{9,1},e_{10,1}^{10,1},e_{11,1}^{11,1}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +4}A_2\left\{\begin{array}{c}e_{12,6}^{12,6},e_{12,7}^{12,7},e_{12,9}^{12,9},e_{13,6}^{13,6},e_{13,8}^{13,8},e_{13,10}^{13,10}\\ e_{14,7}^{14,7},e_{14,8}^{14,8},e_{14,11}^{14,11},e_{15,9}^{15,9},e_{15,10}^{15,10},e_{15,11}^{15,11}\end{array}\right\},`$ $`\mathrm{\Delta }^2q^{\alpha +2}A_1\left\{\begin{array}{c}e_{7,3}^{7,3},e_{8,3}^{8,3},e_{8,4}^{8,4},e_{10,4}^{10,4}\end{array}\right\},\mathrm{\Delta }^2q^{\alpha +5}A_2\left\{\begin{array}{c}e_{13,7}^{13,7},e_{13,9}^{13,9},e_{14,9}^{14,9},e_{14,10}^{14,10}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +6}(\overline{q}^2q^2)A_2\left\{\begin{array}{c}e_{14,6}^{14,6},e_{15,6}^{15,6},e_{15,7}^{15,7},e_{15,8}^{15,8}\end{array}\right\},`$ $`\mathrm{\Delta }^2q^3A_1A_2\left\{\begin{array}{c}e_{12,2}^{12,2},e_{12,3}^{12,3},e_{12,4}^{12,4},e_{13,2}^{13,2},e_{13,3}^{13,3},e_{13,5}^{13,5}\\ e_{14,2}^{14,2},e_{14,4}^{14,4},e_{14,5}^{14,5},e_{15,3}^{15,3},e_{15,4}^{15,4},e_{15,5}^{15,5}\end{array}\right\},`$ $`\mathrm{\Delta }^3q^{\alpha +3}A_0A_1A_2\left\{\begin{array}{c}e_{12,1}^{12,1},e_{13,1}^{13,1},e_{14,1}^{14,1},e_{15,1}^{15,1}\end{array}\right\},`$ $`\mathrm{\Delta }^2q^{2\alpha +7}A_2A_3\left\{\begin{array}{c}e_{16,6}^{16,6},e_{16,7}^{16,7},e_{16,8}^{16,8},e_{16,9}^{16,9},e_{16,10}^{16,10},e_{16,11}^{16,11}\end{array}\right\},`$ $`\mathrm{\Delta }^3q^{\alpha +6}A_1A_2A_3\left\{\begin{array}{c}e_{16,2}^{16,2},e_{16,3}^{16,3},e_{16,4}^{16,4},e_{16,5}^{16,5}\end{array}\right\},\mathrm{\Delta }^3q^6(\overline{q}^2+1+q^2)A_1A_2\left\{\begin{array}{c}e_{15,2}^{15,2}\end{array}\right\},`$ $`\mathrm{\Delta }^2q^4\left\{\begin{array}{c}e_{10,7}^{10,7}\end{array}\right\},\mathrm{\Delta }^2q^5(q+\overline{q})\left\{\begin{array}{c}e_{11,6}^{11,6}\end{array}\right\},\mathrm{\Delta }^3q^4A_1A_2\left\{\begin{array}{c}e_{13,4}^{13,4}\end{array}\right\},`$ $`\mathrm{\Delta }^2q^5(\overline{q}^2q^2)A_1A_2\left\{\begin{array}{c}e_{14,3}^{14,3}\end{array}\right\},\mathrm{\Delta }^4q^6A_0A_1A_2A_3\left\{\begin{array}{c}e_{16,1}^{16,1}\end{array}\right\},`$ $`q^{3\alpha }\left\{\begin{array}{c}e_{2,1}^{1,2},e_{3,1}^{1,3},e_{4,1}^{1,4},e_{5,1}^{1,5}\\ e_{1,2}^{2,1},e_{1,3}^{3,1},e_{1,4}^{4,1},e_{1,5}^{5,1}\end{array}\right\},q^{2\alpha +6}\left\{\begin{array}{c}e_{16,11}^{11,16},e_{16,10}^{10,16},e_{16,9}^{9,16},e_{16,8}^{8,16},e_{16,7}^{7,16},e_{16,6}^{6,16}\\ e_{11,16}^{16,11},e_{10,16}^{16,10},e_{9,16}^{16,9},e_{8,16}^{16,8},e_{7,16}^{16,7},e_{6,16}^{16,6}\end{array}\right\},`$ $`q^{3\alpha +9}\left\{\begin{array}{c}e_{15,16}^{16,15},e_{14,16}^{16,14},e_{13,16}^{16,13},e_{12,16}^{16,12}\\ e_{16,15}^{15,16},e_{16,14}^{14,16},e_{16,13}^{13,16},e_{16,12}^{12,16}\end{array}\right\},q^{2\alpha }\left\{\begin{array}{c}e_{6,1}^{1,6},e_{7,1}^{1,7},e_{8,1}^{1,8},e_{9,1}^{1,9},e_{10,1}^{1,10},e_{11,1}^{1,11}\\ e_{1,6}^{6,1},e_{1,7}^{7,1},e_{1,8}^{8,1},e_{1,9}^{9,1},e_{1,10}^{10,1},e_{1,11}^{11,1}\end{array}\right\},`$ $`q^\alpha \left\{\begin{array}{c}e_{12,1}^{1,12},e_{13,1}^{1,13},e_{14,1}^{1,14},e_{15,1}^{1,15}\\ e_{1,12}^{12,1},e_{1,13}^{13,1},e_{1,14}^{14,1},e_{1,15}^{15,1}\end{array}\right\},q^{2\alpha +1}\left\{\begin{array}{c}e_{3,2}^{2,3},e_{4,2}^{2,4},e_{5,2}^{2,5},e_{4,3}^{3,4},e_{5,3}^{3,5},e_{5,4}^{4,5}\\ e_{2,3}^{3,2},e_{2,4}^{4,2},e_{2,5}^{5,2},e_{3,4}^{4,3},e_{3,5}^{5,3},e_{4,5}^{5,4}\end{array}\right\},`$ $`q^{\alpha +3}\left\{\begin{array}{c}e_{16,5}^{5,16},e_{16,4}^{4,16},e_{16,3}^{3,16},e_{16,2}^{2,16}\\ e_{5,16}^{16,5},e_{4,16}^{16,4},e_{3,16}^{16,3},e_{2,16}^{16,2}\end{array}\right\},q^{2\alpha +7}\left\{\begin{array}{c}e_{14,15}^{15,14},e_{13,15}^{15,13},e_{12,15}^{15,12},e_{13,14}^{14,13},e_{12,14}^{14,12},e_{12,13}^{13,12}\\ e_{15,14}^{14,15},e_{15,13}^{13,15},e_{15,12}^{12,15},e_{14,13}^{13,14},e_{14,12}^{12,14},e_{13,12}^{12,13}\end{array}\right\},`$ $`q^{\alpha +1}\left\{\begin{array}{c}e_{6,2}^{2,6},e_{7,2}^{2,7},e_{8,2}^{2,8},e_{6,3}^{3,6},e_{9,3}^{3,9},e_{10,3}^{3,10},e_{7,4}^{4,7},e_{9,4}^{4,9},e_{11,4}^{4,11},e_{8,5}^{5,8},e_{10,5}^{5,10},e_{11,5}^{5,11}\\ e_{2,6}^{6,2},e_{2,7}^{7,2},e_{2,8}^{8,2},e_{3,6}^{6,3},e_{3,9}^{9,3},e_{3,10}^{10,3},e_{4,7}^{7,4},e_{4,9}^{9,4},e_{4,11}^{11,4},e_{5,8}^{8,5},e_{5,10}^{10,5},e_{5,11}^{11,5}\end{array}\right\},`$ $`q^{\alpha +4}\left\{\begin{array}{c}e_{10,15}^{15,10},e_{11,15}^{15,11},e_{9,15}^{15,9},e_{11,14}^{14,11},e_{8,14}^{14,8},e_{7,14}^{14,7},e_{10,13}^{13,10},e_{8,13}^{13,8},e_{6,13}^{13,6},e_{9,12}^{12,9},e_{7,12}^{12,7},e_{6,12}^{12,6}\\ e_{15,11}^{11,15},e_{15,10}^{10,15},e_{15,9}^{9,15},e_{14,11}^{11,14},e_{14,8}^{8,14},e_{14,7}^{7,14},e_{13,10}^{10,13},e_{13,8}^{8,13},e_{13,6}^{6,13},e_{12,9}^{9,12},e_{12,7}^{7,12},e_{12,6}^{6,12}\end{array}\right\},`$ $`q^{\alpha +2}\left\{\begin{array}{c}e_{9,2}^{2,9},e_{10,2}^{2,10},e_{11,2}^{2,11},e_{7,3}^{3,7},e_{8,3}^{3,8},e_{11,3}^{3,11},e_{6,4}^{4,6},e_{8,4}^{4,8},e_{10,4}^{4,10},e_{6,5}^{5,6},e_{7,5}^{5,7},e_{9,5}^{5,9}\\ e_{2,9}^{9,2},e_{2,10}^{10,2},e_{2,11}^{11,2},e_{3,7}^{7,3},e_{3,8}^{8,3},e_{3,11}^{11,3},e_{4,6}^{6,4},e_{4,8}^{8,4},e_{4,10}^{10,4},e_{5,6}^{6,5},e_{5,7}^{7,5},e_{5,9}^{9,5}\end{array}\right\},`$ $`q^3\left\{\begin{array}{c}e_{7,6}^{6,7},e_{8,6}^{6,8},e_{9,6}^{6,9},e_{10,6}^{6,10},e_{8,7}^{7,8},e_{9,7}^{7,9},e_{11,7}^{7,11},e_{10,8}^{8,10},e_{11,8}^{8,11},e_{10,9}^{9,10},e_{11,9}^{9,11},e_{11,10}^{10,11}\\ e_{6,7}^{7,6},e_{6,8}^{8,6},e_{6,9}^{9,6},e_{6,10}^{10,6},e_{7,8}^{8,7},e_{7,9}^{9,7},e_{8,10}^{10,8},e_{7,11}^{11,7},e_{8,11}^{11,8},e_{9,10}^{10,9},e_{9,11}^{11,9},e_{10,11}^{11,10}\end{array}\right\},`$ $`1\left\{\begin{array}{c}e_{16,1}^{1,16}\\ e_{1,16}^{16,1}\end{array}\right\},q^3\left\{\begin{array}{c}e_{15,2}^{2,15},e_{14,3}^{3,14},e_{13,4}^{4,13},e_{12,5}^{5,12}\\ e_{5,12}^{12,5},e_{4,13}^{13,4},e_{3,14}^{14,3},e_{2,15}^{15,2}\end{array}\right\},q^4\left\{\begin{array}{c}e_{11,6}^{6,11},e_{10,7}^{7,10},e_{9,8}^{8,9}\\ e_{6,11}^{11,6},e_{7,10}^{10,7},e_{8,9}^{9,8}\end{array}\right\},`$ $`q^{\alpha +5}\left\{\begin{array}{c}e_{14,6}^{6,14},e_{15,6}^{6,15},e_{13,7}^{7,13},e_{15,7}^{7,15},e_{12,8}^{8,12},e_{15,8}^{8,15},e_{13,9}^{9,13},e_{14,9}^{9,14},e_{12,10}^{10,12},e_{14,10}^{10,14},e_{12,11}^{11,12},e_{13,11}^{11,13}\\ e_{6,14}^{14,6},e_{6,15}^{15,6},e_{7,13}^{13,7},e_{7,15}^{15,7},e_{8,12}^{12,8},e_{8,15}^{15,8},e_{9,13}^{13,9},e_{9,14}^{14,9},e_{10,12}^{12,10},e_{10,14}^{14,10},e_{11,12}^{12,11},e_{11,13}^{13,11}\end{array}\right\},`$ $`q^2\left\{\begin{array}{c}e_{12,2}^{2,12},e_{13,2}^{2,13},e_{14,2}^{2,14},e_{12,3}^{3,12},e_{13,3}^{3,13},e_{15,3}^{3,15},e_{12,4}^{4,12},e_{14,4}^{4,14},e_{15,4}^{4,15},e_{13,5}^{5,13},e_{14,5}^{5,14},e_{15,5}^{5,15}\\ e_{2,12}^{12,2},e_{3,12}^{12,3},e_{4,12}^{12,4},e_{2,13}^{13,2},e_{3,13}^{13,3},e_{5,13}^{13,5},e_{2,14}^{14,2},e_{4,14}^{14,4},e_{5,14}^{14,5},e_{3,15}^{15,3},e_{4,15}^{15,4},e_{5,15}^{15,5}\end{array}\right\},`$ $`\mathrm{\Delta }q^4\{+\overline{q}\left\{\begin{array}{c}e_{15,2}^{3,14},e_{14,3}^{4,13},e_{13,4}^{5,12}\\ e_{3,14}^{15,2},e_{4,13}^{14,3},e_{5,12}^{13,4}\end{array}\right\},1\left\{\begin{array}{c}e_{15,2}^{4,13},e_{14,3}^{5,12}\\ e_{4,13}^{15,2},e_{5,12}^{14,3}\end{array}\right\},+q\left\{\begin{array}{c}e_{15,2}^{5,12}\\ e_{5,12}^{15,2}\end{array}\right\}\},`$ $`\mathrm{\Delta }q^5\{\overline{q}\left\{\begin{array}{c}e_{11,6}^{7,10},e_{10,7}^{8,9},e_{10,7}^{9,8}\\ e_{7,10}^{11,6},e_{8,9}^{10,7},e_{9,8}^{10,7}\end{array}\right\},+1\left\{\begin{array}{c}e_{11,6}^{8,9},e_{11,6}^{9,8}\\ e_{8,9}^{11,6},e_{9,8}^{11,6}\end{array}\right\},q\left\{\begin{array}{c}e_{11,6}^{10,7}\\ e_{10,7}^{11,6}\end{array}\right\}\},`$ $`\mathrm{\Delta }q^{\alpha +2}\left\{\begin{array}{c}e_{9,2}^{3,7},e_{10,2}^{3,8},e_{7,3}^{4,6},e_{11,2}^{4,8},e_{11,3}^{4,10},e_{8,3}^{5,6},e_{8,4}^{5,7},e_{10,4}^{5,9}\\ e_{3,7}^{9,2},e_{3,8}^{10,2},e_{4,6}^{7,3},e_{4,8}^{11,2},e_{4,10}^{11,3},e_{5,6}^{8,3},e_{5,7}^{8,4},e_{5,9}^{10,4}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +5}\left\{\begin{array}{c}e_{14,6}^{7,13},e_{13,7}^{8,12},e_{15,6}^{9,13},e_{15,7}^{9,14},e_{13,9}^{10,12},e_{15,8}^{10,14},e_{14,9}^{11,12},e_{14,10}^{11,13}\\ e_{7,13}^{14,6},e_{8,12}^{13,7},e_{9,13}^{15,6},e_{9,14}^{15,7},e_{10,12}^{13,9},e_{10,14}^{15,8},e_{11,12}^{14,9},e_{11,13}^{14,10}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +3}\left\{\begin{array}{c}e_{9,2}^{4,6},e_{10,2}^{5,6},e_{11,2}^{5,7},e_{11,3}^{5,9}\\ e_{4,6}^{9,2},e_{5,6}^{10,2},e_{5,7}^{11,2},e_{5,9}^{11,3}\end{array}\right\},\mathrm{\Delta }q^{\alpha +6}\left\{\begin{array}{c}e_{14,6}^{8,12},e_{15,6}^{10,12},e_{15,7}^{11,12},e_{15,8}^{11,13}\\ e_{8,12}^{14,6},e_{10,12}^{15,6},e_{11,12}^{15,7},e_{11,13}^{15,8}\end{array}\right\},`$ $`\mathrm{\Delta }q^{2\alpha +\frac{1}{2}}A_0^{\frac{1}{2}}A_1^{\frac{1}{2}}\left\{\begin{array}{c}e_{6,1}^{2,3},e_{7,1}^{2,4},e_{8,1}^{2,5},e_{9,1}^{3,4},e_{10,1}^{3,5},e_{11,1}^{4,5}\\ e_{2,3}^{6,1},e_{2,4}^{7,1},e_{2,5}^{8,1},e_{3,4}^{9,1},e_{3,5}^{10,1},e_{4,5}^{11,1}\end{array}\right\},`$ $`\mathrm{\Delta }q^{2\alpha +\frac{3}{2}}A_0^{\frac{1}{2}}A_1^{\frac{1}{2}}\left\{\begin{array}{c}e_{6,1}^{3,2},e_{7,1}^{4,2},e_{8,1}^{5,2},e_{9,1}^{4,3},e_{10,1}^{5,3},e_{11,1}^{5,4}\\ e_{3,2}^{6,1},e_{4,2}^{7,1},e_{5,2}^{8,1},e_{4,3}^{9,1},e_{5,3}^{10,1},e_{5,4}^{11,1}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +3}A_1\left\{\begin{array}{c}+\overline{q}\left\{\begin{array}{c}e_{7,3}^{6,4},e_{8,3}^{6,5},e_{8,4}^{7,5},e_{10,4}^{9,5}\\ e_{6,4}^{7,3},e_{6,5}^{8,3},e_{7,5}^{8,4},e_{9,5}^{10,4}\end{array}\right\}\\ 1\left\{\begin{array}{c}e_{9,2}^{6,4},e_{10,2}^{6,5},e_{11,2}^{7,5},e_{11,3}^{9,5}\\ e_{6,4}^{9,2},e_{6,5}^{10,2},e_{7,5}^{11,2},e_{9,5}^{11,3}\end{array}\right\}\\ +q\left\{\begin{array}{c}e_{9,2}^{7,3},e_{10,2}^{8,3},e_{11,2}^{8,4},e_{11,3}^{10,4}\\ e_{7,3}^{9,2},e_{8,3}^{10,2},e_{8,4}^{11,2},e_{10,4}^{11,3}\end{array}\right\}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +6}A_2\left\{\begin{array}{c}\overline{q}\left\{\begin{array}{c}e_{13,7}^{12,8},e_{13,9}^{12,10},e_{14,9}^{12,11},e_{14,10}^{13,11}\\ e_{12,8}^{13,7},e_{12,10}^{13,9},e_{12,11}^{14,9},e_{13,11}^{14,10}\end{array}\right\}\\ +1\left\{\begin{array}{c}e_{14,6}^{12,8},e_{15,6}^{12,10},e_{15,7}^{12,11},e_{15,8}^{13,11}\\ e_{12,8}^{14,6},e_{12,10}^{15,6},e_{12,11}^{15,7},e_{13,11}^{15,8}\end{array}\right\}\\ q\left\{\begin{array}{c}e_{14,6}^{13,7},e_{15,6}^{13,9},e_{15,7}^{14,9},e_{15,8}^{14,10}\\ e_{13,7}^{14,6},e_{13,9}^{15,6},e_{14,9}^{15,7},e_{14,10}^{15,8}\end{array}\right\}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +2}A_0^{\frac{1}{2}}A_2^{\frac{1}{2}}\left\{\begin{array}{c}\overline{q}\left\{\begin{array}{c}e_{12,1}^{2,9},e_{13,1}^{2,10},e_{14,1}^{2,11},e_{15,1}^{3,11}\\ e_{2,9}^{12,1},e_{2,10}^{13,1},e_{2,11}^{14,1},e_{3,11}^{15,1}\end{array}\right\}\\ +1\left\{\begin{array}{c}e_{12,1}^{3,7},e_{13,1}^{3,8},e_{14,1}^{4,8},e_{15,1}^{4,10}\\ e_{3,7}^{12,1},e_{3,8}^{13,1},e_{4,8}^{14,1},e_{4,10}^{15,1}\end{array}\right\}\\ q\left\{\begin{array}{c}e_{12,1}^{4,6},e_{13,1}^{5,6},e_{14,1}^{5,7},e_{15,1}^{5,9}\\ e_{4,6}^{12,1},e_{5,6}^{13,1},e_{5,7}^{14,1},e_{5,9}^{15,1}\end{array}\right\}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\alpha +5}A_1^{\frac{1}{2}}A_3^{\frac{1}{2}}\left\{\begin{array}{c}+\overline{q}\left\{\begin{array}{c}e_{16,2}^{6,14},e_{16,3}^{6,15},e_{16,4}^{7,15},e_{16,5}^{8,15}\\ e_{6,14}^{16,2},e_{6,15}^{16,3},e_{7,15}^{16,4},e_{8,15}^{16,5}\end{array}\right\}\\ 1\left\{\begin{array}{c}e_{16,2}^{7,13},e_{16,3}^{9,13},e_{16,4}^{9,14},e_{16,5}^{10,14}\\ e_{7,13}^{16,2},e_{9,13}^{16,3},e_{9,14}^{16,4},e_{10,14}^{16,5}\end{array}\right\}\\ +q\left\{\begin{array}{c}e_{16,2}^{8,12},e_{16,3}^{10,12},e_{16,4}^{11,12},e_{16,5}^{11,13}\\ e_{8,12}^{16,2},e_{10,12}^{16,3},e_{11,12}^{16,4},e_{11,13}^{16,5}\end{array}\right\}\end{array}\right\},`$ $`\mathrm{\Delta }^2q^{\alpha +3}A_0^{\frac{1}{2}}A_1A_2^{\frac{1}{2}}\left\{\begin{array}{c}+\overline{q}\left\{\begin{array}{c}e_{12,1}^{6,4},e_{13,1}^{6,5},e_{14,1}^{7,5},e_{15,1}^{9,5}\\ e_{6,4}^{12,1},e_{6,5}^{13,1},e_{7,5}^{14,1},e_{9,5}^{15,1}\end{array}\right\}\\ 1\left\{\begin{array}{c}e_{12,1}^{7,3},e_{13,1}^{8,3},e_{14,1}^{8,4},e_{15,1}^{10,4}\\ e_{7,3}^{12,1},e_{8,3}^{13,1},e_{8,4}^{14,1},e_{10,4}^{15,1}\end{array}\right\}\\ +q\left\{\begin{array}{c}e_{12,1}^{9,2},e_{13,1}^{10,2},e_{14,1}^{11,2},e_{15,1}^{11,3}\\ e_{9,2}^{12,1},e_{10,2}^{13,1},e_{11,2}^{14,1},e_{11,3}^{15,1}\end{array}\right\}\end{array}\right\},`$ $`\mathrm{\Delta }^2q^{\alpha +6}A_1^{\frac{1}{2}}A_2A_3^{\frac{1}{2}}\left\{\begin{array}{c}+\overline{q}\left\{\begin{array}{c}e_{16,2}^{12,8},e_{16,3}^{12,10},e_{16,4}^{12,11},e_{16,5}^{13,11}\\ e_{12,8}^{16,2},e_{12,10}^{16,3},e_{12,11}^{16,4},e_{13,11}^{16,5}\end{array}\right\}\\ 1\left\{\begin{array}{c}e_{16,2}^{13,7},e_{16,3}^{13,9},e_{16,4}^{14,9},e_{16,5}^{14,10}\\ e_{13,7}^{16,2},e_{13,9}^{16,3},e_{14,9}^{16,4},e_{14,10}^{16,5}\end{array}\right\}\\ +q\left\{\begin{array}{c}e_{16,2}^{14,6},e_{16,3}^{15,6},e_{16,4}^{15,7},e_{16,5}^{15,8}\\ e_{14,6}^{16,2},e_{15,6}^{16,3},e_{15,7}^{16,4},e_{15,8}^{16,5}\end{array}\right\}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\frac{5}{2}}A_1^{\frac{1}{2}}A_2^{\frac{1}{2}}\left\{\begin{array}{c}e_{12,2}^{6,7},e_{13,2}^{6,8},e_{12,3}^{6,9},e_{13,3}^{6,10},e_{14,2}^{7,8},e_{12,4}^{7,9},e_{14,4}^{7,11},e_{13,5}^{8,10},e_{14,5}^{8,11},e_{15,3}^{9,10},e_{15,4}^{9,11},e_{15,5}^{10,11}\\ e_{6,7}^{12,2},e_{6,8}^{13,2},e_{6,9}^{12,3},e_{6,10}^{13,3},e_{7,8}^{14,2},e_{7,9}^{12,4},e_{7,11}^{14,4},e_{8,10}^{13,5},e_{8,11}^{14,5},e_{9,10}^{15,3},e_{9,11}^{15,4},e_{10,11}^{15,5}\end{array}\right\},`$ $`\mathrm{\Delta }q^{\frac{7}{2}}A_1^{\frac{1}{2}}A_2^{\frac{1}{2}}\left\{\begin{array}{c}e_{12,2}^{7,6},e_{13,2}^{8,6},e_{14,2}^{8,7},e_{12,3}^{9,6},e_{12,4}^{9,7},e_{13,3}^{10,6},e_{13,5}^{10,8},e_{15,3}^{10,9},e_{14,4}^{11,7},e_{14,5}^{11,8},e_{15,4}^{11,9},e_{15,5}^{11,10}\\ e_{7,6}^{12,2},e_{8,6}^{13,2},e_{8,7}^{14,2},e_{9,6}^{12,3},e_{9,7}^{12,4},e_{10,6}^{13,3},e_{10,8}^{13,5},e_{10,9}^{15,3},e_{11,7}^{14,4},e_{11,8}^{14,5},e_{11,9}^{15,4},e_{11,10}^{15,5}\end{array}\right\},`$ $`\mathrm{\Delta }q^4A_1^{\frac{1}{2}}A_2^{\frac{1}{2}}\{\overline{q}^{\frac{1}{2}}\left\{\begin{array}{c}e_{14,3}^{6,11},e_{13,4}^{7,10},e_{12,5}^{8,9}\\ e_{6,11}^{14,3},e_{7,10}^{13,4},e_{8,9}^{12,5}\end{array}\right\},+q^{\frac{1}{2}}\left\{\begin{array}{c}e_{13,4}^{9,8},e_{12,5}^{10,7},e_{15,2}^{6,11}\\ e_{9,8}^{13,4},e_{10,7}^{12,5},e_{6,11}^{15,2}\end{array}\right\}\},`$ $`\mathrm{\Delta }q^6A_1^{\frac{1}{2}}A_2^{\frac{1}{2}}\{\overline{q}^{\frac{1}{2}}\left\{\begin{array}{c}e_{14,3}^{9,8},e_{15,2}^{7,10},e_{12,5}^{11,6}\\ e_{9,8}^{14,3},e_{7,10}^{15,2},e_{11,6}^{12,5}\end{array}\right\},+q^{\frac{1}{2}}\left\{\begin{array}{c}e_{13,4}^{11,6},e_{14,3}^{10,7},e_{15,2}^{8,9}\\ e_{11,6}^{13,4},e_{10,7}^{14,3},e_{8,9}^{15,2}\end{array}\right\}\},`$ $`\mathrm{\Delta }^2q^4A_1^{\frac{1}{2}}A_2^{\frac{1}{2}}\{+\overline{q}^{\frac{1}{2}}\left\{\begin{array}{c}e_{14,3}^{7,10},e_{13,4}^{8,9}\\ e_{7,10}^{14,3},e_{8,9}^{13,4}\end{array}\right\},q^{\frac{1}{2}}\left\{\begin{array}{c}e_{14,3}^{8,9},e_{13,4}^{10,7}\\ e_{10,7}^{13,4},e_{8,9}^{14,3}\end{array}\right\}\},`$ $`\mathrm{\Delta }^2q^{\frac{11}{2}}(\overline{q}^2q^2)A_1^{\frac{1}{2}}A_2^{\frac{1}{2}}\{\overline{q}\left\{\begin{array}{c}e_{15,2}^{9,8}\\ e_{9,8}^{15,2}\end{array}\right\},+1\left\{\begin{array}{c}e_{15,2}^{10,7},e_{14,3}^{11,6}\\ e_{10,7}^{15,2},e_{11,6}^{14,3}\end{array}\right\},q\left\{\begin{array}{c}e_{15,2}^{11,6}\\ e_{11,6}^{15,2}\end{array}\right\}\},`$ $`\mathrm{\Delta }^2q^6A_1A_2\{\overline{q}^2\left\{\begin{array}{c}e_{13,4}^{12,5}\\ e_{12,5}^{13,4}\end{array}\right\},\overline{q}\left\{\begin{array}{c}e_{14,3}^{12,5}\\ e_{12,5}^{14,3}\end{array}\right\},1\left\{\begin{array}{c}e_{15,2}^{12,5},e_{14,3}^{13,4}\\ e_{12,5}^{15,2},e_{13,4}^{14,3}\end{array}\right\},q\left\{\begin{array}{c}e_{15,2}^{13,4}\\ e_{13,4}^{15,2}\end{array}\right\},q^2\left\{\begin{array}{c}e_{15,2}^{14,3}\\ e_{14,3}^{15,2}\end{array}\right\}\},`$ $`\mathrm{\Delta }^2q^5A_0^{\frac{1}{2}}A_1^{\frac{1}{2}}A_2^{\frac{1}{2}}A_3^{\frac{1}{2}}\times `$ $`\{\overline{q}^2\left\{\begin{array}{c}e_{16,1}^{6,11}\\ e_{6,11}^{16,1}\end{array}\right\},\overline{q}\left\{\begin{array}{c}e_{16,1}^{7,10}\\ e_{7,10}^{16,1}\end{array}\right\},1\left\{\begin{array}{c}e_{16,1}^{8,9},e_{16,1}^{9,8}\\ e_{8,9}^{16,1},e_{9,8}^{16,1}\end{array}\right\},q\left\{\begin{array}{c}e_{16,1}^{10,7}\\ e_{10,7}^{16,1}\end{array}\right\},q^2\left\{\begin{array}{c}e_{16,1}^{11,6}\\ e_{11,6}^{16,1}\end{array}\right\}\},`$ $`\mathrm{\Delta }q^3A_0^{\frac{1}{2}}A_3^{\frac{1}{2}}\{+\overline{q}^{\frac{3}{2}}\left\{\begin{array}{c}e_{16,1}^{2,15}\\ e_{2,15}^{16,1}\end{array}\right\},\overline{q}^{\frac{1}{2}}\left\{\begin{array}{c}e_{16,1}^{3,14}\\ e_{3,14}^{16,1}\end{array}\right\},+q^{\frac{1}{2}}\left\{\begin{array}{c}e_{16,1}^{4,13}\\ e_{4,13}^{16,1}\end{array}\right\},q^{\frac{3}{2}}\left\{\begin{array}{c}e_{16,1}^{5,12}\\ e_{5,12}^{16,1}\end{array}\right\}\},`$ $`\mathrm{\Delta }^3q^6A_0^{\frac{1}{2}}A_1A_2A_3^{\frac{1}{2}}\{+\overline{q}^{\frac{3}{2}}\left\{\begin{array}{c}e_{16,1}^{12,5}\\ e_{12,5}^{16,1}\end{array}\right\},\overline{q}^{\frac{1}{2}}\left\{\begin{array}{c}e_{16,1}^{13,4}\\ e_{13,4}^{16,1}\end{array}\right\},+q^{\frac{1}{2}}\left\{\begin{array}{c}e_{16,1}^{14,3}\\ e_{14,3}^{16,1}\end{array}\right\},q^{\frac{3}{2}}\left\{\begin{array}{c}e_{16,1}^{15,2}\\ e_{15,2}^{16,1}\end{array}\right\}\},`$ $`\mathrm{\Delta }q^{2\alpha +\frac{13}{2}}A_2^{\frac{1}{2}}A_3^{\frac{1}{2}}\left\{\begin{array}{c}e_{16,6}^{12,13},e_{16,7}^{12,14},e_{16,8}^{13,14},e_{16,9}^{12,15},e_{16,10}^{13,15},e_{16,11}^{14,15}\\ e_{12,13}^{16,6},e_{12,14}^{16,7},e_{13,14}^{16,8},e_{12,15}^{16,9},e_{13,15}^{16,10},e_{14,15}^{16,11}\end{array}\right\},`$ $`\mathrm{\Delta }q^{2\alpha +\frac{15}{2}}A_2^{\frac{1}{2}}A_3^{\frac{1}{2}}\left\{\begin{array}{c}e_{16,6}^{13,12},e_{16,7}^{14,12},e_{16,8}^{14,13},e_{16,9}^{15,12},e_{16,10}^{15,13},e_{16,11}^{15,14}\\ e_{13,12}^{16,6},e_{14,12}^{16,7},e_{14,13}^{16,8},e_{15,12}^{16,9},e_{15,13}^{16,10},e_{15,14}^{16,11}\end{array}\right\},`$ and the left handle $`C`$ has $`16`$ diagonal components: $`C=q^{4\alpha 6}\left\{\begin{array}{c}+q^6\left\{e_1^1\right\}\hfill \\ q^3\{q^3e_2^2,qe_3^3,\overline{q}e_4^4,\overline{q}^3e_5^5\}\hfill \\ +\{q^4e_6^6,q^2e_7^7,e_8^8,e_9^9,\overline{q}^2e_{10}^{10},\overline{q}^4e_{11}^{11}\}\hfill \\ \overline{q}^3\{\overline{q}^3e_{15}^{15},\overline{q}e_{14}^{14},qe_{13}^{13},q^3e_{12}^{12}\}\hfill \\ +\overline{q}^6\left\{e_{16}^{16}\right\}\hfill \end{array}\right\},`$ using the scaling factors: $`\kappa _\sigma =q^{4\alpha },\kappa _C=q^{4\alpha }.`$ ## Appendix B Evaluations of $`LG^3`$ and $`LG^4`$ Below, we present evaluations for $`LG^3`$ and $`LG^4`$ for a few $`2`$ and $`3`$-braid links. We use the same naming conventions for links as those of , except that we denote by $`2_{1a}^2`$ and $`2_{1b}^2`$ the $`2`$ component links determined respectively by the braids $`\sigma _1^{\pm 1}`$. To present evaluations of $`LG^m`$, we use a similar convention to that of . The expression $`s_0(A_0(q)),s_1(A_1(q)),\mathrm{},s_r(A_r(q))`$, where the $`s_i`$ are signs $`\pm `$ and the $`A_i(q)`$ are integer-coefficient Laurent polynomials in $`q`$, is intended to be read: $`s_0(A_0(q))+s_1(\overline{P}^1+P^1)(A_1(q))+\mathrm{}+s_r(\overline{P}^r+P^r)(A_r(q)).`$ In these expressions $`(A_i(q))`$ is only a list of terms of $`A_i(q)`$ rather than an explicit sum, viz we have written $`(x_1,x_2,\mathrm{},x_s)`$ for $`(x_1+x_2+\mathrm{}+x_s)`$. Recall that, for $`LG^{m,n}`$ (fixing $`m,n`$), we are using the variable $`P=p^2=q^{2\alpha +mn}`$; viz for $`LG^3LG^{3,1}`$ we use $`P=q^{2\alpha +2}`$, and for $`LG^4LG^{4,1}`$ we use $`P=q^{2\alpha +3}`$. For multicomponent links, if the polynomial is not invariant under $`P\overline{P}`$, we write it out in full. This situation only occurs here for $`LG^3`$ for links of $`2`$ components. Within the $`q`$-polynomials, the same general behaviours of the coefficients as reported for $`LG^2`$ in are seen. These calculations were performed on SUN Ultra 60 UNIX based workstations, with a main memory of 256Mb, and the larger calculations sometimes used all of this memory. ### Evaluations of $`LG^3`$ $`\begin{array}{ccc}\hfill LG_{3_1}^3& =& (q^2,2q^4,3q^6,q^8),+(q^2,2q^4,2q^6,q^8),(q^4,q^6,q^8),+\left(q^6\right)\hfill \\ & & \\ \hfill LG_{4_1}^3& =& +(5\overline{q}^4,9\overline{q}^2,17,9q^2,5q^4),(2\overline{q}^4,8\overline{q}^2,10,8q^2,2q^4),+(3\overline{q}^2,3,3q^2),\left(1\right)\hfill \\ & & \\ \hfill LG_{5_1}^3& =& +(q^4,4q^6,5q^8,5q^{10},3q^{12},q^{14}),(q^4,3q^6,5q^8,5q^{10},3q^{12},q^{14}),\hfill \\ & & +(q^4,2q^6,4q^8,4q^{10},3q^{12},q^{14}),(q^6,2q^8,3q^{10},3q^{12},q^{14}),\hfill \\ & & +(q^8,2q^{10},2q^{12},q^{14}),(q^{10},q^{12},q^{14}),+\left(q^{12}\right)\hfill \\ & & \\ \hfill LG_{5_2}^3& =& (7q^2,17q^4,32q^6,25q^8,15q^{10},3q^{12}),+(4q^2,15q^4,23q^6,22q^8,11q^{10},3q^{12}),\hfill \\ & & (7q^4,10q^6,12q^8,5q^{10},2q^{12}),+(4q^6,2q^8,2q^{10})\hfill \end{array}`$ $`\begin{array}{ccc}\hfill LG_{6_2}^3& =& (7\overline{q}^2,29,60q^2,74q^4,60q^6,26q^8,5q^{10}),\hfill \\ & & +(6\overline{q}^2,24,52q^2,67q^4,54q^6,26q^8,5q^{10}),\hfill \\ & & (2\overline{q}^2,14,32q^2,48q^4,41q^6,23q^8,5q^{10}),+(4,15q^2,24q^4,28q^6,14q^8,5q^{10}),\hfill \\ & & (4q^2,10q^4,12q^6,8q^8,2q^{10}),+(3q^4,3q^6,3q^8),\left(q^6\right)\hfill \\ & & \\ \hfill LG_{6_3}^3& =& +(9\overline{q}^6,52\overline{q}^4,106\overline{q}^2,145,106q^2,52q^4,9q^6),\hfill \\ & & (9\overline{q}^6,42\overline{q}^4,96\overline{q}^2,120,96q^2,42q^4,9q^6),\hfill \\ & & +(6\overline{q}^6,26\overline{q}^4,63\overline{q}^2,77,63q^2,26q^4,6q^6),\hfill \\ & & (2\overline{q}^6,13\overline{q}^4,28\overline{q}^2,40,28q^2,13q^4,2q^6),+(4\overline{q}^4,10\overline{q}^2,14,10q^2,4q^4),\hfill \\ & & (3\overline{q}^2,3,3q^2),+\left(1\right)\hfill \end{array}`$ $`\begin{array}{ccc}\hfill LG_{2_{1a}^2}^3& =& \left(\overline{p}^1p^1\right)(q,q^3,q^5),+\left(\overline{p}^3p^3\right)\left(q^3\right)\hfill \\ & & \\ \hfill LG_{2_{1b}^2}^3& =& +\left(\overline{p}^1p^1\right)(\overline{q}^5,\overline{q}^3,\overline{q}^1),\left(\overline{p}^3p^3\right)\left(\overline{q}^3\right)\hfill \\ & & \\ \hfill LG_{4_{1a}^2}^3& =& +\left(\overline{p}^1p^1\right)(q^3,3q^5,4q^7,3q^9,q^{11}),\left(\overline{p}^3p^3\right)(q^3,2q^5,3q^7,3q^9,q^{11})\hfill \\ & & +\left(\overline{p}^5p^5\right)(q^5,2q^7,2q^9,q^{11}),\left(\overline{p}^7p^7\right)(q^7,q^9,q^{11}),+\left(\overline{p}^9p^9\right)\left(q^9\right)\hfill \\ & & \\ \hfill LG_{4_{1b}^2}^3& =& \left(\overline{p}^1p^1\right)(4q^1,6q^3,8q^5,4q^7,2q^9),+\left(\overline{p}^3p^3\right)(4q^3,2q^5,2q^7)\hfill \\ & & \\ \hfill LG_{5_1^2}^3& =& +\left(\overline{p}^1p^1\right)(5\overline{q}^7,23\overline{q}^5,36\overline{q}^3,37\overline{q}^1,19q^1,6q^3),\hfill \\ & & \left(\overline{p}^3p^3\right)(5\overline{q}^7,14\overline{q}^5,28\overline{q}^3,22\overline{q}^1,13q^1,2q^3)\hfill \\ & & +\left(\overline{p}^5p^5\right)(2\overline{q}^7,8\overline{q}^5,12\overline{q}^3,10\overline{q}^1,4q^1)\hfill \\ & & \left(\overline{p}^7p^7\right)(3\overline{q}^5,3\overline{q}^3,3\overline{q}^1),+\left(\overline{p}^9p^9\right)\left(\overline{q}^3\right)\hfill \\ & & \\ \hfill LG_{6_1^2}^3& =& \left(\overline{p}^1p^1\right)(q^5,4q^7,6q^9,7q^{11},5q^{13},3q^{15},q^{17})\hfill \\ & & +\left(\overline{p}^3p^3\right)(q^5,3q^7,6q^9,6q^{11},5q^{13},3q^{15},q^{17})\hfill \\ & & \left(\overline{p}^5p^5\right)(q^5,2q^7,4q^9,5q^{11},5q^{13},3q^{15},q^{17})\hfill \\ & & +\left(\overline{p}^7p^7\right)(q^7,2q^9,4q^{11},4q^{13},3q^{15},q^{17})\hfill \\ & & \left(\overline{p}^9p^9\right)(q^9,2q^{11},3q^{13},3q^{15},q^{17}),+\left(\overline{p}^{11}p^{11}\right)(q^{11},2q^{13},2q^{15},q^{17})\hfill \\ & & \left(\overline{p}^{13}p^{13}\right)(q^{13},q^{15},q^{17}),+\left(\overline{p}^{15}p^{15}\right)\left(q^{15}\right)\hfill \\ & & \\ \hfill LG_{6_2^2}^3& =& +\left(\overline{p}^1p^1\right)(7q^3,26q^5,52q^7,59q^9,43q^{11},17q^{13},3q^{15})\hfill \\ & & \left(\overline{p}^3p^3\right)(4q^3,18q^5,36q^7,48q^9,34q^{11},16q^{13},3q^{15})\hfill \\ & & +\left(\overline{p}^5p^5\right)(7q^5,19q^7,27q^9,23q^{11},11q^{13},3q^{15})\hfill \\ & & \left(\overline{p}^7p^7\right)(7q^7,10q^9,12q^{11},5q^{13},2q^{15}),+\left(\overline{p}^9p^9\right)(4q^9,2q^{11},2q^{13})\hfill \\ & & \\ & & \\ \hfill LG_{6_1^3}^3& =& (10\overline{q}^2,38,82q^2,100q^4,84q^6,36q^8,8q^{10}),\hfill \\ & & +(7\overline{q}^2,32,69q^2,91q^4,73q^6,36q^8,7q^{10}),\hfill \\ & & (2\overline{q}^2,18,42q^2,63q^4,53q^6,29q^8,6q^{10}),+(5,19q^2,30q^4,33q^6,16q^8,5q^{10}),\hfill \\ & & (5q^2,11q^4,13q^6,8q^8,2q^{10}),+(3q^4,3q^6,3q^8),\left(q^6\right)\hfill \\ & & \\ \hfill LG_{6_2^3}^3& =& +(18\overline{q}^6,100\overline{q}^4,206\overline{q}^2,276,206q^2,100q^4,18q^6),\hfill \\ & & (17\overline{q}^6,81\overline{q}^4,183\overline{q}^2,230,183q^2,81q^4,17q^6),\hfill \\ & & +(11\overline{q}^6,48\overline{q}^4,117\overline{q}^2,143,117q^2,48q^4,11q^6),\hfill \\ & & (3\overline{q}^6,23\overline{q}^4,49\overline{q}^2,70,49q^2,23q^4,3q^6),+(6\overline{q}^4,16\overline{q}^2,22,16q^2,6q^4),\hfill \\ & & (4\overline{q}^2,4,4q^2),+\left(1\right)\hfill \end{array}`$ $`\begin{array}{ccc}\hfill LG_{6_3^3}^3& =& +(2q^6,2q^8,2q^{10}),(q^4,2q^6,3q^8,2q^{10},q^{12}),+(q^4,2q^6,3q^8,2q^{10},q^{12}),\hfill \\ & & (q^6,q^8,q^{10},q^{12}),+(q^{10},q^{12},q^{14}),(q^{10},q^{12},q^{14}),+\left(q^{12}\right)\hfill \end{array}`$ ### Evaluations of $`LG^4`$ $`\begin{array}{ccc}\hfill LG_{2_{1a}^2}^4& =& +(q^2,q^4,2q^6,q^8,q^{10}),(q^3,q^5,q^7,q^9),+\left(q^6\right)\hfill \\ & & \\ \hfill LG_{2_{1b}^2}^4& =& +(\overline{q}^{10},\overline{q}^8,2\overline{q}^6,\overline{q}^4,\overline{q}^2),(\overline{q}^9,\overline{q}^7,\overline{q}^5,\overline{q}^3),+\left(\overline{q}^6\right)\hfill \\ & & \\ \hfill LG_{3_1}^4& =& +(q^4,2q^6,4q^8,4q^{10},5q^{12},2q^{14},q^{16}),(q^5,2q^7,4q^9,4q^{11},3q^{13},2q^{15}),\hfill \\ & & +(q^6,2q^8,2q^{10},3q^{12},q^{14},q^{16}),(q^9,q^{11},q^{13},q^{15}),+\left(q^{12}\right)\hfill \\ & & \\ \hfill LG_{4_{1a}^2}^4& =& +(q^6,2q^8,5q^{10},8q^{12},10q^{14},8q^{16},7q^{18},2q^{20},q^{22}),\hfill \\ & & (q^7,3q^9,6q^{11},8q^{13},9q^{15},7q^{17},4q^{19},2q^{21}),\hfill \\ & & +(q^8,3q^{10},5q^{12},7q^{14},6q^{16},6q^{18},2q^{20},q^{22}),\hfill \\ & & (q^9,2q^{11},3q^{13},5q^{15},4q^{17},3q^{19},2q^{21}),+(q^{12},2q^{14},2q^{16},3q^{18},q^{20},q^{22}),\hfill \\ & & (q^{15},q^{17},q^{19},q^{21}),+\left(q^{18}\right)\hfill \\ & & \\ \hfill LG_{5_1}^4& =& +(q^8,2q^{10},6q^{12},10q^{14},15q^{16},16q^{18},15q^{20},10q^{22},7q^{24},2q^{26},q^{28}),\hfill \\ & & (q^9,3q^{11},7q^{13},12q^{15},15q^{17},15q^{19},13q^{21},8q^{23},4q^{25},2q^{27}),\hfill \\ & & +(q^{10},4q^{12},7q^{14},11q^{16},13q^{18},13q^{20},9q^{22},7q^{24},2q^{26},q^{28}),\hfill \\ & & (q^{11},3q^{13},6q^{15},9q^{17},10q^{19},10q^{21},7q^{23},4q^{25},2q^{27}),\hfill \\ & & +(q^{12},2q^{14},4q^{16},6q^{18},7q^{20},6q^{22},6q^{24},2q^{26},q^{28}),\hfill \\ & & (q^{15},2q^{17},3q^{19},5q^{21},4q^{23},3q^{25},2q^{27}),\hfill \\ & & +(q^{18},2q^{20},2q^{22},3q^{24},q^{26},q^{28}),(q^{21},q^{23},q^{25},q^{27}),+\left(q^{24}\right)\hfill \end{array}`$
warning/0004/hep-ph0004046.html
ar5iv
text
# Towards softer scales in hot QCD ## I Introduction The framework of this paper is pure gauge theory in thermal equilibrium. with $`g<<1`$ It has been realized that the near-equilibrium dynamics of high temperature QCD involves a hierarchy of length scales scale $`(T)^1`$ inverse of the typical momentum of the plasma particles scale $`(gT)^1`$ electric screening length and first scale of collective excitations scale $`(g^2T\mathrm{ln}1/g)^1`$ damping length of colour excitations scale ($`g^2T)^1`$ non perturbative magnetic fluctuations. Effective theories can be constructed at a given scale that integrate out the smaller scales. The prototype is the hard thermal loop ( HTL) effective theory that integrate out the scale $`(T)^1`$ . To integrate out the hard modes $`pT`$ means to calculate loop diagrams with external momenta $`<<T`$ and internal momenta of order $`T`$. It generates effective propagators and effective vertices for the field modes with momenta $`<<T`$. To leading order, only one-loop diagrams survive and the result is the hard thermal loop $`n`$-point functions. Remarkably, the generating functional of the hard thermal loops is gauge invariant. Bödeker’s break-through was to realize that one may integrate out the scale $`(gT)^1`$. The summation now involves an infinite set of multiloop diagrams with loop momenta $`gT`$ and effective HTL propagators and effective vertices. One way to construct effective theories for the soft field modes of the plasma is to use classical transport equations. The soft field modes behave classically, this is due to the fact that the momentum scales of interest are small compared to the temperature, so that the modes have large occupation numbers. The resulting set of equations that characterize the effective theory has now been obtained through different approaches . This work will follow the approach of Blaizot and Iancu to the construction of the effective theories, we summarize it here. At high temperature, the non-abelian pure gauge theory describes a weakly coupled plasma whose constituents, gluons have a typical momenta $`kT`$ and a typical distance $`(T)^1`$. These plasma particles may take part in long-wavelength collective excitations, which are fluctuations in the average density of the plasma particles. These excitations may be induced by a weak external disturbance. The collective excitations at scale $`x`$ are described by a field $`W(x,𝐯)`$, a colour matrix in the adjoint representation $`W(x,𝐯)=W^a(x,𝐯)T^a`$, where $`𝐯`$ is the direction of propagation of the excitation. Transport equations describe the space-time evolution of those long-wavelenth excitations. The soft fields that one is interested in, are represented by mean fields. In the transport equation, they couple to the collective excitations via a mean-field term $`𝐯.𝐄^a(x)`$ and via the covariant derivative. The soft fields obey Maxwell equations, where the collective excitations act as an induced source $$D^\nu F_{\mu \nu }(x)=j_\mu ^{ext}(x)+j_\mu ^{ind}(x)$$ (1) where the induced current is the response of the plasma to the initial perturbation $`j_\mu ^{ext}(x)`$ $$j_\mu ^{aind}(x)=m_D^2\frac{d\mathrm{\Omega }_𝐯}{4\pi }v_\mu W^a(x,𝐯)$$ (2) with $$m_D^2=\frac{g^2NT^2}{3}$$ (3) On the space-time scale $`(gT)^1`$, the transport equation is $$v.D_x^{ab}W^b(x,𝐯)=𝐯.𝐄^a(x)$$ (4) with $`v_\mu (v_0=1,𝐯)`$. The effective theory for the soft field modes is obtained as follows. By solving the transport equation via a Green function, one can express the collective excitation $`W(x,𝐯)`$ as a functional of the soft fields. Then the induced current, Eq.(2), acts as a generating functional for the $`n`$-point amplitudes of the soft fields. The resulting amplitudes are the HTL amplitudes. When one is interested in collective excitations involving colour fluctuations on larger wavelength, one integrates out both scales $`(T)^1`$ and $`(gT)^1`$. It turns out that the resulting set of equations is similar. The new element is the inclusion of the effect of collisions and they are dominated by small angle scattering. Indeed a scattering process which hardly changes the momentum of a hard particle, can change its colour charge. The transport equation at larger wavelength $`x>>(gT)^1`$ is $$[v.D_x^{ab}+\widehat{C}]W^b(x,𝐯)=𝐯.𝐄^a(x)$$ (5) The collision term $`\widehat{C}=C(𝐯,𝐯^{})`$ is non local in $`𝐯`$, but local in $`x`$ and blind to colour. The gluons $`pT`$ take part in the collective motion, the gluons $`pgT`$ are exchanged in the collision process. The soft fields obey the Maxwell equations (1) and the effective theory is obtained from the induced current (2). The resulting new feature is the effect of dissipation as a result of collisions. A new scale appears, the colour relaxation time $`1/\gamma `$ with $`\gamma g^2T\mathrm{ln}1/g`$. This purely dissipative description is appropriate to the study of the relaxation of a weak initial disturbance at scale $`x`$. However colour excitations may also be generated by thermal fluctuations in the plasma. The fluctuation-dissipation theorem relates the two phenomena. In Bödeker’s alternative construction of the effective theory (without an external source) , the collision term in Eq.(5) is accompanied by a Gaussian white noise term, which arises from the thermal fluctuations of initial conditions of the field with momentum $`gT`$. The noise term keeps the soft modes in thermal equilibrium, it injects energy which compensates for the loss of energy at scale $`x`$ from the damping term. One quantity that turns out to be particularly sensitive to the change of scale is the polarization tensor in the magnetic sector. Precisely that quantity enters the collision term of the transport equation in a strategic way. Indeed the interaction rate between two hard particles is dominated by soft momentum transfer $`qgT`$ where the squared modulus of the resummed gluon propagator enters $$Im_D^2_{q_0<<q}\frac{dq_0d_3q}{q^4+(\mathrm{Im}\mathrm{\Pi }^t)^2}\frac{1}{q^2}$$ where $`1/q^2`$ comes from energy-momentum conservation at each vertex When the scale $`T`$ has been integrated out, $`\mathrm{\Pi }^t`$ has the well known HTL expression $`\mathrm{Im}\mathrm{\Pi }^t=\pi m_D^2q_0/q`$ for $`q_0<<q`$ so that $`I𝑑q/q`$ leading to the $`\mathrm{ln}1/g`$ factor in the scale of the collision term. That expression for $`\mathrm{\Pi }^t`$ turns out to be valid at the scale $`gT`$ only. Indeed, when the scale $`gT`$ is integrated out the resulting $`\mathrm{\Pi }^t`$ is $`\mathrm{Im}\mathrm{\Pi }^t=m_D^2q_0/3\gamma `$ for $`q_0<<q<<\gamma `$ so that the integral $`I`$ now behaves as $`I𝑑q/q^2`$. Again this $`\mathrm{\Pi }^t`$ is valid at the scale $`g^2T\mathrm{ln}1/g`$. If one wants to reach the scale $`g^2T`$, one has to integrate out the scale $`g^2T\mathrm{ln}1/g`$ in order to see how it affects the physics at the scale $`g^2T`$. Arnold and Yaffe have recently attacked the problem of taking the scale $`g^2T\mathrm{ln}1/g`$ into account. As they are interested in the colour conductivity, they concentrate on static quantities. They make use of effective theories and have computed the one-loop contribution to $`\mathrm{\Pi }^l(q_0=0,q)`$ in an effective theory. In this work, the one-loop contribution to the polarization tensor $`\mathrm{\Pi }^{\mu \nu }(q_0,q)`$ will be computed for loop momenta $`g^2T\mathrm{ln}1/g`$ and external momenta $`g^2T`$ assuming $`\mathrm{ln}1/g>>1`$. It is only the first term of an infinite series of multiloop diagrams. However, following a remark made by Blaizot and Iancu for the case of momenta $`gT`$, it will be possible to identify in the one-loop term, the new collision term that should be added in the transport equation in order to sum the infinite series. This is only a first step towards integrating out the scale $`g^2T\mathrm{ln}1/g`$. We will not address the more ambitious question whether integrating out this scale amounts to take into account collisions with exchange gluon $`g^2T\mathrm{ln}1/g`$. The one-soft-loop contribution to the polarization tensor $`\mathrm{\Pi }^{\mu \nu }(q_0,q)`$ comes from a well known pair of diagrams whose contribution adds up to give a transverse $`\mathrm{\Pi }^{\mu \nu }`$. One diagram possesses two effective three-gluon vertices and two effective propagators, the other one has an effective four-gluon vertex and an effective propagator. For the case of a loop momentum $`g^2T\mathrm{ln}1/g`$, up to now explicit expressions have only been written for the polarization tensor and for the linearized form of the transport equation (5) . One part of this work will be devoted to write down the explicit form of the needed vertices. In the following, HTL amplitudes refer to the effective amplitudes when the scale $`(T)^1`$ is integrated out, soft amplitudes refer to the case when the scales $`(T)^1`$ and $`(gT)^1`$ are integrated out. The remarkable similarities between the HTL amplitudes and the soft amplitudes will be a recurrent theme in this work: The $`n`$-point amplitudes are obtained from the induced current through the same steps, they obey the same Ward identities. In the one-soft-loop diagram, this similarity will show up again and it will lead to the existence of the new collision operator $`\widehat{C^{}}`$ which exhibits features similar to the operator $`\widehat{C}`$. There are also interesting differences, the soft amplitudes contain damping in a way that does not conflict with gauge symmetry. The transport equation (4) has been shown to be gauge invariant, the transport equation (5) has been established in the Coulomb gauge, however the collision operator is expected to be gauge-independent . Section 2 is devoted to the polarization tensor $`\mathrm{\Pi }^{\mu \nu }`$. An explicit Lorentz-covariant form is obtained for the case when the scale $`(gT)^1`$ is integrated out. $`\mathrm{\Pi }^{\mu \nu }`$ appears under a form that allows to interpolate between the scale $`gT`$ and the scale $`g^2T\mathrm{ln}1/g`$. Then it is detailed how the one-loop contribution to the gluon self-energy with loop momenta $`g^2T\mathrm{ln}1/g`$ allows to identify the new collision operator, under which assumptions. In Section 3, explicit forms of the effective three-gluon and four-gluon vertices are obtained. The response approach makes use of a Retarded Green function, it leads to the use of the Retarded/Advanced formalism. The $`n`$-point Retarded soft amplitudes are obtained from the induced current and are shown to obey tree-like Ward identities in any leg. The reason is, they are a sum of tree diagrams, the propagator along the tree is the (linearized) one of $`W^b(x,𝐯)`$, i.e. the soft fields induce the long-wavelength collective excitations. A needed generalization is then made for the four-gluon vertex. In Section 4 the one-soft-loop self-energy diagrams are evaluated in the Retarded/Advanced formalism. The diagram with 3-gluon effective vertices is evaluated first. Loop momenta $`gT`$ and $`g^2T\mathrm{ln}1/g`$ are successively considered. For the case $`g^2T\mathrm{ln}1/g`$ one part of the new collision term is identified, its physical interpretation, its scale are examined. Then, the one-loop self-energy diagram with a 4-gluon vertex is treated in a similar way. The total one-loop $`\mathrm{\Pi }^{\mu \nu }`$ is transverse, the total collision operator has a zero mode. The similarities between the two collision operators appear. In Section 5, some properties and consequences of the new collision operator $`\widehat{C^{}}`$ are studied, in particular its infrared behaviour. Conclusions are in Section 6. In an Appendix, the explicit expression of $`\mathrm{\Pi }(q_0,q)`$ at the scale $`g^2T\mathrm{ln}1/g`$ is written down and its analytical properties in the complex $`q_0`$ plane are studied. Another Appendix contains explicit expressions for the eigenvalues of the operator $`\widehat{C^{}}`$ for the pure magnetic sector’s case. ## II The Polarization tensor Our notations are $`K(k_0,𝐤),K^2=k_0^2k^2`$ ### A The collision-resummed $`W`$ propagator The transport equation for the collective $`W`$ field at a scale $`x>>(gT)^1`$ is $$(v.D_x+\widehat{C})W(x,𝐯)=𝐯.𝐄(x)$$ (6) where $`D_x`$ is the covariant derivative and $`\widehat{C}`$ the collision operator. The linearized equation is in Fourier space $$(iv.K+\widehat{C})W(K,𝐯)=𝐯.𝐄(K)$$ (7) where $`v.K=v_0k_0𝐯.𝐤`$ and $`v(v_0=1,𝐯)`$ with $`𝐯^2=1`$ , and $`\widehat{C}`$ is an operator in $`𝐯`$ space $$\widehat{C}W=\frac{d\mathrm{\Omega }_𝐯^{}}{4\pi }C(𝐯,𝐯^{})W(K,𝐯^{})$$ (8) $`C(𝐯,𝐯^{})`$ is a real function, symmetric in $`𝐯`$ and $`𝐯^{}`$ $$C(𝐯,𝐯^{})=\gamma \delta _{S_2}(𝐯𝐯^{})m_D^2\frac{g^2NT}{2}\mathrm{\Phi }(𝐯.𝐯^{})$$ (9) $$\frac{d\mathrm{\Omega }_𝐯^{}}{4\pi }C(𝐯,𝐯^{})=0$$ (10) and the scale of $`\widehat{C}`$ is set by $$\gamma g^2T\mathrm{ln}1/g$$ (11) The explicit form of $`\mathrm{\Phi }(𝐯.𝐯^{})`$ is in Sec.IV B. The eigenvalues of the $`\widehat{C}`$ operator are all positive, as the first term in (9) dominates over the second one, except for the eigenvalue zero of $`\widehat{C}`$, i.e. Eq.(10). The linearized $`W`$ propagator is the Green function associated with Eq.(7) i.e. the inverse operator $$\widehat{G}_R(K,v)=(v.K+i\widehat{C})^1$$ (12) When $`\widehat{C}`$ is replaced by $`ϵ>0`$, $`G_R`$ is the retarded Green function associated with the drift equation. Note that this operator in $`𝐯`$ space should properly be written $$\widehat{G}_R(K,v)=(v.K+iv_0\widehat{C})^1$$ (13) where $``$ is the identity operator in $`𝐯`$ space, so that $$\widehat{G}_R(K,v_0,𝐯)=\widehat{G}_R(K,v_0,𝐯)$$ (14) The advanced propagator is $$\widehat{G}_A(K,v)=(v.Ki\widehat{C})^1$$ (15) and one has $$\widehat{G}_A(K,v)=\widehat{G}_R(K,v)^{}\widehat{G}_A(K,v)=\widehat{G}_R(K,v)$$ (16) Operations in $`𝐯`$ space We adopt the notations of Arnold and Yaffe The measure is $`𝑑\mathrm{\Omega }_𝐯^{}/(4\pi )`$ $`<>_v`$ denotes averaging over the direction $`𝐯`$ $`<\delta _{S_2}(𝐯𝐯^{})>_v^{}=1=<>_v^{}`$ (17) $`\widehat{C}W=<C(𝐯,𝐯^{})W(𝐯^{})>_v^{}`$ (18) $`<v_i\widehat{C}v_j>=<v_iC(𝐯,𝐯^{})v_{}^{}{}_{j}{}^{}>_{v,v^{}}`$ (19) so that $`>`$ represent any function independent of $`𝐯`$ . As stressed in an important property of $`\widehat{C}`$ is $$\widehat{C}>=0\mathrm{or}<\widehat{C}=0$$ (20) from (10), with the useful consequence $$(v.K+i\widehat{C})^1v.K>=(v.K+i\widehat{C})^1(v.K+i\widehat{C})>=>=>$$ (21) and a similar one for $`(v.Ki\widehat{C})^1`$ ### B The collision-resummed polarisation tensor Blaizot and Iancu have extracted from the linearized induced current, the form of the polarization tensor that takes into account the full effect of collisions . Defining $$W(K,𝐯)=iW^i(K,𝐯)E^i(K)$$ (22) the new functions $`W^i(K,𝐯)`$ satisfy, from (7) $$(v.K+i\widehat{C})W^i(K,𝐯)=v^i$$ (23) and they obtained (see Eqs.(4.6-4.7) in ) $`\mathrm{\Pi }^{\mu i}`$ $`=`$ $`q_0m_D^2{\displaystyle \frac{d\mathrm{\Omega }_𝐯}{4\pi }v^\mu W^i(Q,𝐯)}`$ (24) $`\mathrm{\Pi }^{\mu 0}`$ $`=`$ $`q^im_D^2{\displaystyle \frac{d\mathrm{\Omega }_𝐯}{4\pi }v^\mu W^i(Q,𝐯)}`$ (25) with $`m_D^2=g^2NT^2/3`$. The solution of (23) may be written in terms of the retarded inverse operator defined in Eq. (12) $$W_R^i(K,𝐯)=(v.K+i\widehat{C})^1v^i$$ (26) Eq.(25) become $$\mathrm{\Pi }_R^{\mu i}(Q)=m_D^2q_0<v^\mu (v.Q+i\widehat{C})^1v^i>_{v,v^{}}$$ (27) $$\mathrm{\Pi }_R^{\mu o}(Q)=m_D^2<v^\mu (v.Q+i\widehat{C})^1𝐯.𝐪>_{v,v^{}}$$ (28) With $`𝐯.𝐪=q_0v.Q`$ and Eq.(21), (28) becomes $$\mathrm{\Pi }^{\mu 0}=m_D^2[q_0<v^\mu (v.Q+i\widehat{C})^1>_{v,v^{}}<v^\mu >_{v,v^{}}]$$ (29) so that $`\mathrm{\Pi }^{\mu \nu }`$ may be written in the compact form $$\mathrm{\Pi }_R^{\mu \nu }(Q)=m_D^2[q_0<v^\mu (v.Q+i\widehat{C})^1v^\nu >_{v,v^{}}g^{\mu 0}g^{\nu 0}]$$ (30) (30) explicits many properties of $`\mathrm{\Pi }^{\mu \nu }`$. Indeed $`\mathrm{\Pi }^{\mu \nu }(Q)=\mathrm{\Pi }^{\nu \mu }(Q)`$ since $`\widehat{C}`$ is symetric in $`𝐯`$ space, and one readily verifies that $`Q_\mu \mathrm{\Pi }^{\mu \nu }=Q_\nu \mathrm{\Pi }^{\mu \nu }=0`$ since, again with Eq.(21) $$Q_\mu \mathrm{\Pi }^{\mu \nu }=m_D^2[q_0<v^\nu >_{v.v^{}}q_0g^{\nu 0}]=0$$ (31) Moreover, for momenta $`q_0,q>>g^2\mathrm{ln}1/g`$ , one may replace the operator $`\widehat{C}`$ by $`ϵ>0`$, the $`W`$ propagator is now diagonal in $`𝐯`$ space and one recovers the well-known HTL form for $`\mathrm{\Pi }^{\mu \nu }`$ $$\mathrm{\Pi }_R^{\mu \nu }(Q)=m_D^2[q_0<v^\mu (v.Q+iϵ)^1v^\nu >_{v,v^{}}g^{\mu 0}g^{\nu 0}]$$ (32) (30) interpolates between the collisionless scale $`(T)^1`$ where the mean field $`W`$ propagates on a straight line and the scale $`(gT)^1`$ where collisions cause fluctuations of the direction $`𝐯`$ of the W field. Another aspect of (30) follows. An interpretation in terms of tree diagrams may be attached to the first term, i.e. to the Landau damping term. The propagator along the tree is the collision-resummed $`W`$ propagator. It carries the momentum of the incoming gluon, $`v^\mu `$ is the vertex for a gluon (polarisation $`\mu `$) attached to the $`W`$ line. The W propagator is blind to colour. The retarded prescription of the incoming gluon determines the retarded prescription for the $`W`$ propagator. In Sec.III B the expression (30) will be recovered in the framework of the functional derivatives of the induced current. In Appendix A , an explicit form of $`\mathrm{\Pi }^{\mu \nu }`$ is given, as a continuous fraction, and its analytic properties in the complex $`q_0`$ plane are examined. ### C Collision operator and one-loop soft exchange in $`\mathrm{\Pi }^{ji}`$ In his quest for an effective theory for the soft field modes $`Qg^2T`$ , Bödeker’s first step was to estimate the contribution from the one-soft-loop diagrams to the polarization tensor $`\mathrm{\Pi }^{\mu \nu }(Q)`$ for loop mpmentum $`KgT`$ and $`Qg^2T`$. Those one-loop diagrams involve hard thermal loop effective vertices and propagators. Then higher loop diagrams with momentum $`KgT`$ were resummed via a transport equation, so that both scales $`T`$ and $`gT`$ were integrated out . This transport equation is in terms of the field $`W(K,𝐯)`$ that describes the collective excitations of the hard particles. Fluctuations in $`𝐯`$ are caused by collisions and taken into account via the operator $`\widehat{C}(𝐯,𝐯^{})`$. The linearized transport equation may be written (see Eq.(23)) $$(v.K+i\widehat{C})W^i(K,𝐯)=v^i$$ (33) while the integration over the scale $`T`$ only, lead to the linearized transport equation $$v.KW^i=v^i$$ (34) The complete solution to the transport equation (33) involves the inverse operator $`(v.K+i\widehat{C})^1`$ $$W_R^i(K,v)=(v.K+i\widehat{C})^1v^i$$ (35) As recently stressed by Blaizot and Iancu , alternatively one may solve the transport equation (33) by iteration. One obtains a series expansion in powers of $`\widehat{C}`$ whose first terms reproduce the expansion in the number of loops with momentum $`KgT`$. Indeed, writing Eq.(33) $$v.KW^i=v^ii\widehat{C}W^i$$ (36) one obtains $$W_R^{i(0)}=\frac{v^i}{v.K+iϵ},ϵ>0$$ (37) $$W_R^{i(1)}=\frac{1}{v.K+iϵ}(i)\widehat{C}\frac{v^i}{v.K+iϵ}=\frac{1}{v.K+iϵ}<(i)C(𝐯,𝐯^{})\frac{v^i}{v^{}.K+iϵ}>_v^{}$$ (38) Since the polarization tensor is simply related to the $`W^i`$ field (see Eq. (24)) $$\mathrm{\Pi }^{ji}(Q)=q_0m_D^2<v^jW^i(Q,𝐯)>_v$$ (39) one obtains $$\mathrm{\Pi }^{ji(0)}(Q)=q_0m_D^2<v^j\frac{1}{v.Q+iϵ}v^i>_v$$ (40) $`\mathrm{\Pi }^{ji(1)}(Q)`$ $`=`$ $`q_0m_D^2<{\displaystyle \frac{v^j}{v.Q+iϵ}}(i)\widehat{C}{\displaystyle \frac{v^j}{v.Q+iϵ}}>_{v,v^{}}`$ (41) $`=`$ $`q_0m_D^2<{\displaystyle \frac{v^j}{v.Q+iϵ}}(i)C(𝐯,𝐯^{}){\displaystyle \frac{v^j}{v^{}.Q+iϵ}}>_{v,v^{}}`$ (42) $`\mathrm{\Pi }^{ji(0)}`$ is the hard thermal loop contribution to the polarization tensor, $`\mathrm{\Pi }^{ji(1)}`$ is the one-loop soft momentum contribution for $`KgT`$ with $`C(𝐯,𝐯^{})`$ as in (9) . The conclusion is, one can identify the collision operator $`\widehat{C}`$ if $`\mathrm{\Pi }^{ji(1)}`$ is known and appears under the form (42). We shall follow this strategy. We want to integrate out the momenta $`Kg^2T\mathrm{ln}1/g`$ to see how they affect the dynamics of the soft fields $`Qg^2T`$. We shall assume $`\mathrm{ln}1/g>>1`$ so that the scales are well separated. If one assumes that the summation over the scale $`g^2T\mathrm{ln}1/g`$ can be made via a transport equation which involves a collision operator $`\widehat{C}^{}`$, i.e. $$(v.K+i\widehat{C}+i\widehat{C}^{})W^i(K,𝐯)=v^i$$ (43) the full solution will be in terms of an inverse operator $$W^i=(v.K+i\widehat{C}+i\widehat{C}^{})^1v^i$$ (44) and the iterative solution will be $$W^{i(0)}=(v.K+i\widehat{C})^1v^i$$ (45) $$W^{i(1)}=(v.K+i\widehat{C})^1(i)\widehat{C}^{}(v.K+i\widehat{C})^1v^i$$ (46) so that $$\mathrm{\Pi }_{}^{}{}_{}{}^{ji(0)}(Q)=q_0m_D^2<v^j(v.Q+i\widehat{C})^1v^i>_{v,v^{}}$$ (47) $$\mathrm{\Pi }_{}^{}{}_{}{}^{ji(1)}(Q)=q_0m_D^2<v^j(v.Q+i\widehat{C})^1(i)\widehat{C}^{}(v.Q+i\widehat{C})^1v^i>_{allv}$$ (48) $`\mathrm{\Pi }_{}^{}{}_{}{}^{ji(0)}`$ is the form found in Eq. (30) for the polarization tensor, when the scales $`T`$ and $`gT`$ have been integrated out. $`\mathrm{\Pi }_{}^{}{}_{}{}^{ji(1)}`$ will correspond to the one-loop soft momentum exchange $`Kg^2T\mathrm{ln}1/g`$, where the one-loop diagrams involve effective vertices and resummed propagators with the scales $`T`$ and $`gT`$ integrated out. The effective vertices will be constructed in Sec.III. The one-loop soft momentum exchange $`Kg^2T\mathrm{ln}1/g`$ contribution to the polarization tensor $`\mathrm{\Pi }_{}^{}{}_{}{}^{ji(1)}`$ will be computed in Sec.IV, it will indeed show up as in (48). This fact will allow to identify the operator $`\widehat{C}^{}`$. The expression for $`C(𝐯,𝐯^{})`$ involves by construction an integration over the momentum $`KgT`$. So does $`C^{}(𝐯,𝐯^{})`$ with momentum $`Kg^2T\mathrm{ln}1/g`$. There wont be any over-counting in $`\widehat{C}+\widehat{C}^{}`$ in Eqs.(43,44) if one limits the integration range over the space momentum $`k`$ $`\mu _2<k<\mu _1\mathrm{for}\widehat{C}\mu _3<k<\mu _2\mathrm{for}\widehat{C}^{}`$ (50) $`gT<\mu _1<T,g^2T\mathrm{ln}1/g<\mu _2<gT,g^2T<\mu _3<g^2T\mathrm{ln}1/g`$ ## III The effectives vertices In the response approach, a weak external perturbation is applied to the plasma at some scale $`x`$ with an adiabatic switching-on at time $`\mathrm{}`$ . The response of the plasma is an induced current. The $`n`$-point soft amplitudes are obtained as functional derivatives of this current with respect to the mean field. The response approach involves the Retarded 2-point Green function (the appropriate one to study the response of the plasma to a perturbation which vanishes at time $`\mathrm{}`$) and it gives out the $`n`$-point Retarded amplitudes $`V(X_1;X_2\mathrm{}.X_n)`$ where $`X_1^0`$ is the largest time (see Sec. III B). In momentum space, the 2-point Retarded function is the analytical continuation into the upper complex $`p_0`$ plane of the Imaginary Time (IT) Green function, i.e. one approaches the real energy axis from above $`G(p_0+iϵ)=G_{ret}`$ . The $`n`$-point Retarded amplitude is the analytical continuation $`p_i^0+iϵ_i`$ of the IT amplitude such that $`ϵ_1>0`$ and all other $`ϵ_i<0`$. (The constraint $`_ip_i^0=0`$ is imposed on any analytical continuation into the complex energy variables $`p_i^0`$, in particular $`_{i=1}^nϵ_i=0`$ ) . Since the known amplitudes will be the Retarded ones, it is natural to use the Retarded/Advanced formalism in computing the two one-soft-loop diagrams that contribute to $`\mathrm{\Pi }_{\mu \nu }`$. Are needed the propagator, the 3-point and 4-point effective vertices. We summarize first the features of this formalism as they will be needed later on, in particular for the 4-point vertex. ### A A summary of the R/A formalism It is a Real Time formalism where the propagators are the Retarded and Advanced propagators of the zero temperature Field Theory. General properties have been established . An external leg $`l`$ may be of type $`R`$ or of type $`A`$, it incoming energy is $`p_l^0+iϵ_l,ϵ_l>0`$ type $`R`$, $`ϵ_l<0`$ type $`A`$. The $`n`$-point amplitudes are defined with all momenta incoming $$\underset{l=1}{\overset{n}{}}p_l^0=0\mathrm{and}\underset{l=1}{\overset{n}{}}ϵ_l=0$$ Then the only non-zero three-point functions have two legs of type $`R`$ and one leg of type $`A`$, or two legs of type $`A`$ and one leg of type $`R`$. The general properties are, for bosons $$V(P_{1R},P_{2A},P_{3A})=V^{}(P_{1A},P_{2R},P_{3R})$$ (51) and for massless bosonic fields in the QED/QCD case $$V((P_1)_A,(P_2)_R,(P_3)_R)=V(P_{1R},P_{2A},P_{3A})$$ (52) With the same notation, the two-point function, should be written $$\mathrm{\Pi }_R(P)\mathrm{\Pi }(P_R)=\mathrm{\Pi }(P_R,(P)_A)$$ (53) and the relations analogous to (51,52) are $`\mathrm{\Pi }(P_A)`$ $`=`$ $`\mathrm{\Pi }^{}(P_R)`$ (54) $`\mathrm{\Pi }((P)_R)`$ $`=`$ $`\mathrm{\Pi }(P_A)`$ (55) Our results will obey those general properties. See (30) for $`\mathrm{\Pi }^{\mu \nu }`$. The $`n`$-point Retarded functions that appear in the response approach are $`V(P_{1R},P_{2A},\mathrm{}P_{nA})`$ , they are related by complex conjugation to the Advanced ones $$V(P_{1R},P_{2A},\mathrm{}P_{nA})=V^{}(P_{1A},P_{2R},\mathrm{}P_{nR})$$ (56) When constructing a diagram, because any $`n`$-point vertex is defined with all incoming momenta, a propagator joins an $`R`$ leg to an $`A`$ leg since a $`P_R(p_0+iϵ,𝐩)`$ incoming a vertex comes from another vertex with incoming momentum $`(p_0iϵ,𝐩)`$ i.e.$`(P)_A`$. One can define an $`ϵ`$-flow along any tree diagram. This flow is incoming the $`R`$ legs and outgoing the $`A`$ legs, it obeys the rules of an electric current since there is no source or sink at a vertex (where $`_iϵ_i=0`$). In this R/A formalism, the thermal Bose weight $`n(p_0)`$ that are associated with the loop momenta in the Imaginary Time formalism, are carried by the vertex in a very specific way, they are attached to the vertices that possess two (or more) legs of type $`A`$ $$\mathrm{\Gamma }(P_{1R},P_{2R},P_{3A})=V(P_{1R},P_{2R},P_{3A})$$ (57) $$\mathrm{\Gamma }(P_{1R},P_{2A},P_{3A})=V(P_{1R},P_{2A},P_{3A})N(P_2,P_3)$$ (58) with $$N(P_2,P_3)=n(p_2^0)+n(p_3^0)+1=\frac{n(p_2^0)n(p_3^0)}{n(p_2^0+p_3^0)}$$ (59) $$N(P,P)=n(p^0)+n(p^0)+1=0$$ (60) For three legs of type $`A`$ $$\mathrm{\Gamma }(P_{1R},P_{2A},P_{3A},P_{4A})=V(P_{1R},P_{2A},P_{3A},P_{4A})𝒩(P_2,P_3,P_4)$$ (61) with $$𝒩(P_2,P_3,P_4)=\frac{n(p_2^0)n(p_3^0)n(p_4^0)}{n(p_2^0+p_3^0+p_4^0)}=N(P_2,P_3)N(P_2+P_3,P_4)$$ (62) In order to build one of the pair of one-loop diagrams contributing to the self-energy, a 4-point amplitude with two legs of type R and two legs of type A is needed. Indeed a loop has to be made by joining an R leg to an A leg after the insertion of a propagator. That amplitude possesses a new feature, the prescription $`p_i^0+iϵ_i`$ on the external legs does not constrain all the energy variables of the amplitude . For the case $`ϵ_1>0,ϵ_2>0,ϵ_3<0,ϵ_4<0`$, $`(ϵ_1+ϵ_2+ϵ_3+ϵ_4=0)`$, the energy variable $`p_1^0+p_3^0=(p_2^0+p_4^0)`$ may have $`ϵ_1+ϵ_3>0`$ or $`<0`$. The same for the variable $`p_1^0+p_4^0`$. As a result there are, for this case, four distinct analytical continuations of the Imaginary time amplitude, known as Generalized-Retarded amplitudes and they are linked by one relation (see references in ) $`V^{(4)}(ϵ_1+ϵ_3>0,ϵ_1+ϵ_4>0)+V^{(4)}(ϵ_1+ϵ_3<0,ϵ_1+ϵ_4<0)`$ (64) $`=V^{(4)}(ϵ_1+ϵ_3>0,ϵ_1+ϵ_4<0)+V^{(4)}(ϵ_1+ϵ_3<0,ϵ_1+ϵ_4>0)`$ The 4-point amplitude RRAA that occurs in the R/A formalism is a specific combination of those analytic continuations . The rules are easily stated when the amplitude is a sum of tree diagrams, the only case to be encountered in this work. One splits the tree diagrams into three groups, which depend respectively on $`P_1+P_2,P_1+P_3,P_1+P_4`$ and $`\mathrm{\Gamma }(P_{1R},P_{2R},P_{3A},P_{4A})=F_1((P_1+P_2)_R)N(P_3,P_4)`$ (67) $`+F_2((P_1+P_3)_R)N(P_3,P_2+P_4)+F_2((P_1+P_3)_A)N(P_4,P_1+P_3)`$ $`+F_3((P_1+P_4)_R)N(P_4,P_2+P_3)+F_3((P_1+P_4)_A)N(P_3,P_1+P_4)`$ From Eqs. (59, 60) $$N(P_3,P_2+P_4)+N(P_4,P_1+P_3)=N(P_3,P_4)=N(P_4,P_2+P_3)+N(P_3,P_1+P_4)$$ (68) Those weights are compatible with the complex conjugate relation $$\frac{\mathrm{\Gamma }(P_{1R},P_{2R},P_{3A},P_{4A})}{\mathrm{\Gamma }^{}(P_{1A},P_{2A},P_{3R},P_{4R})}=\frac{N(P_3,P_4)}{N(P_1,P_2)}$$ (69) indeed $$\frac{N(P_3,P_4)}{N(P_1,P_2)}=\frac{n(p_3^0)n(p_4^0)}{n(p_1^0)n(p_2^0)}=\frac{N(P_3,P_2+P_4)}{N(P_2,P_1+P_3)}=\frac{N(P_4,P_1+P_3)}{N(P_1,P_2+P_4)}=\mathrm{}$$ (70) as may be shown by writing $`N(P_i,P_j)=N(P_i,P_j)`$ in each denominator. ### B Soft Amplitudes from the induced current The soft amplitudes will be obtained as functional derivative of the induced current with respect to the mean field in a way that closely parallels the case of the HTL amplitudes (see Sec.5 of ). One set of Ward Identities satisfied by the amplitudes will follow from the conservation law obeyed by the induced current. It is convenient to introduce the Retarded Green function associated with the transport equation (5) for the $`W(X,𝐯)`$ field. $`i(v.D_x+\widehat{C})G_{ret}(X,Y;𝐯,𝐯^{})`$ $`=`$ $`\delta ^{(4)}(XY)\delta _{S_2}(𝐯𝐯^{})`$ (71) $`G_{ret}(X,Y;𝐯,𝐯^{})`$ $`=`$ $`0\mathrm{f}orX_0<Y_0`$ (72) with $$\widehat{C}G_{ret}(X,Y;𝐯,𝐯^{})=\frac{d\mathrm{\Omega }_{𝐯\mathrm{"}}}{4\pi }C(𝐯,𝐯\mathrm{"})G_{ret}(X,Y;𝐯\mathrm{"},𝐯^{})$$ (73) For the case $`\widehat{C}=0`$, the solution is known in terms of the parallel transporter along a straight line $`G_{C=0}(X,Y;𝐯)`$ $`=`$ $`i{\displaystyle _0^{\mathrm{}}}𝑑u\delta ^{(4)}(XYvu)U(X,Y)`$ (74) $`U(X,Xvu)`$ $`=`$ $`P\{\mathrm{exp}igu{\displaystyle _0^1}dsv.A(Xvu(1s))\}`$ (75) where $`P`$ is the path-ordering operator. An iterative solution to Eq.(71) may be written as a power series in $`\widehat{C}`$ in terms of $`G_{C=0}`$. As a consequence, the following identity, true for $`G_{C=0}\delta _{S_2}(𝐯𝐯^{})`$, also holds for $`G_{ret}(X,Y;𝐯,𝐯^{})`$ $$\frac{G_{ret}(X,Y;𝐯,𝐯^{})}{A_c^\mu (X_1)}=\frac{d\mathrm{\Omega }_{𝐯\mathrm{"}}}{4\pi }G_{ret}(X,X_1;𝐯,𝐯\mathrm{"})gT^cv\mathrm{"}_\mu G_{ret}(X_1,Y;𝐯\mathrm{"},𝐯^{})$$ (76) where the time arguments satisfy $`X^0X_1^0Y^0`$ and $`T^c`$ is in the adjoint representation. This identity is of central importance to the structure of the amplitudes. The solution to the transport equation (5) may be written $$W^a(X,𝐯)=id^4Y\frac{d\mathrm{\Omega }_𝐯^{}}{4\pi }G_{ret}^{ab}(X,Y;𝐯,𝐯^{})𝐯^{}.𝐄^b(Y)$$ (77) and from Eq.(2) the induced current become $$j_\mu ^{aind}(X)=im_D^2\frac{d\mathrm{\Omega }_𝐯}{4\pi }\frac{d\mathrm{\Omega }_𝐯^{}}{4\pi }d^4Yv_\mu G_{ret}^{ab}(X,Y;𝐯,𝐯^{})𝐯^{}.𝐄^b(Y)$$ (78) Writing $$𝐯^{}.𝐄(Y)=v^iF_{0i}(Y)=_{Y_O}(v^\mu A_\mu )v^{}.D_YA_0(Y)$$ (79) Eq.(78) may be transformed into a form similar to the case of the HTL amplitudes $$j_\mu ^{aind}(X)=m_D^2g_{0\mu }A_0^a(X)+im_D^2\frac{d\mathrm{\Omega }_𝐯}{4\pi }\frac{d\mathrm{\Omega }_𝐯^{}}{4\pi }d^4Yv_\mu G_{ret}^{ab}(X,Y;𝐯,𝐯^{})_{Y_0}(v^{}.A^b(Y))$$ (80) Indeed, for the second term on the right handside of (79) $`v^{}.D_Y`$ is allowed to act on the left on $`G_{ret}(X,Y;𝐯,𝐯^{})`$; with the use of (71), it brings the first term on the right handside of (80), the term $`G_{ret}(X,Y;𝐯,𝐯\mathrm{"})C(𝐯\mathrm{"},𝐯^{})`$ does not contribute as the integration over $`𝐯^{}`$ makes use of the property (10) of $`\widehat{C}`$. The soft one-particle-irreducible amplitudes are then obtained $$\mathrm{\Pi }_{\mu \nu }^{ab}(X,Y)=\frac{j_\mu ^{aind}(X)}{A_b^\nu (Y)}|_{A=0}$$ (81) and for the $`n+1`$-gluon amplitude $$g^nV_{\mu \mu _1\mathrm{}\mu _n}^{aa_1\mathrm{}a_n}(X;X_1\mathrm{}X_n)=\frac{^n}{A_{a_n}^{\mu _n}(X_n)\mathrm{}A_{a_1}^{\mu _1}(X_1)}j_\mu ^a(X)|_{A=0}$$ (82) For all these amplitudes, $`X^0`$ is the largest time. After differentiation, the Green function that will enter the amplitudes correspond to the case $`A=0`$, i. e. to $$i(v._X+\widehat{C})G_{ret,A=0}(X,Y;𝐯,𝐯^{})=\delta ^{(4)}(XY)\delta _{S_2}(𝐯𝐯^{})$$ (83) or in momentum space $$(v.P+i\widehat{C})G_{ret}(P;𝐯,𝐯^{})=\mathrm{i}.e.G_{ret}(P;𝐯,𝐯^{})=(v.P+i\widehat{C})^1$$ (84) This is the Green function introduced in Sec.II A (see Eq.(12)) and called the $`W`$ propagator. Moreover, from Eq.(1), the induced current obeys the conservation law $$D^\mu j_\mu ^{aind}(X)=0$$ (85) Writing $`D^\mu =^\mu \delta ^{ab}+gf^{abc}A_c^\mu `$ and formally $$j_\mu ^{aind}=\mathrm{\Pi }_{\mu \nu }^{ab}A_b^\nu +\frac{g}{2!}V_{\mu \nu \rho }^{aa_2a_3}A_{a_2}^\nu A_{a_3}^\rho +\frac{g^2}{3!}V_{\mu \nu \rho \sigma }^{aa_2a_3a_4}A_{a_2}^\nu A_{a_3}^\rho A_{a_4}^\sigma +\mathrm{}$$ (86) The term in $`g^{n2}`$ of (85) gives a Ward identity for the $`X`$ variable, relating the $`n`$-point amplitude to $`n1`$-point amplitudes. For $`\mathrm{\Pi }_{\mu \nu }^{ab}(X)`$, the functional derivative (81) acting on the first term in (80) gives $$m_D^2\delta ^{ab}g_{0\mu }g_{0\nu }\delta ^{(4)}(XY)$$ In the second term, the derivative has to act on $`_{Y_0}v^{}.A^b(Y)`$. In momentum space, one gets back the expression (30) obtained in Sec.II B. From the properties of the functional derivatives and from the identity (76) for the Green function, follow the consequences: \- the $`n+1`$-point amplitude is symmetric in the exchange of any pair of the legs $`X_1,X_2,\mathrm{}X_n`$ \- a functional derivative inserts a vertex along a propagator at an intermediate time \- the order in colour space and in $`𝐯`$ space follows the time order \- a diagram will correspond to a specific time ordering and there will be a sum over all chronological orderings of $`X_1^0,X_2^0,\mathrm{}X_n^0`$. $`X^0`$ is the largest time, a characteristic feature of the $`n+1`$-Retarded amplitude \- the form has the minimum number of colour $`T`$ matrices in the adjoint representation, none for $`\mathrm{\Pi }_{\mu \nu }`$, one for $`V_{\mu \nu \rho }`$, two for $`V_{\mu \nu \rho \sigma }`$ \- one goes back to the corresponding HTL amplitudes by replacing $`\widehat{C}`$ by $`ϵ`$ where $``$ is the identity in $`𝐯`$ space. The Fourier transform of an $`n`$-point amplitude is defined $`(2\pi )^4\delta (P_1+P_2+\mathrm{}+P_n)V(P_1,P_2,\mathrm{}P_n)=`$ (88) $`{\displaystyle d_4X_1\mathrm{}d_4X_n\mathrm{exp}i(P_1X_1+\mathrm{}+P_nX_n)V(X_1,X_2\mathrm{}X_n)}`$ all the momenta are incoming the amplitude and the colour indices will be written $`a_1=1,a_2=2,\mathrm{}`$. All the properties are consequences of the transport equation for the $`W`$ field. The transport equation (5) has been established in the Coulomb gauge, it is expected to be gauge independent . At the HTL level, there is no induced current for the ghost field, i.e. no effective vertices. It is likely to be the same for the soft amplitudes. ### C The 3-gluon vertex The result of the functional differentiation (82) on Eq.(80) is in momentum space $`V_{\mu \nu \rho }^{123}(P_{1R},P_{2A},P_{3A})=im_D^2\{`$ (91) $`f^{123}<v_\mu (v.P_1+i\widehat{C})^1v_{}^{}{}_{\nu }{}^{}(v.(P_1+P_2)+i\widehat{C})^1v\mathrm{"}_\rho (p_3^0)>_{vv^{}v\mathrm{"}}`$ $`+f^{132}<v_\mu (v.P_1+i\widehat{C})^1v_{}^{}{}_{\rho }{}^{}(v.(P_1+P_3)+i\widehat{C})^1v\mathrm{"}_\nu (p_2^0)>_{vv^{}v\mathrm{"}}\}`$ $`V_{\mu \nu \rho }^{123}(P_{1R},P_{2A},P_{3A})=im_D^2f^{123}<v_\mu (v.P_1+i\widehat{C})^1`$ (93) $`[v_{}^{}{}_{\nu }{}^{}(v.P_3+i\widehat{C})^1v\mathrm{"}_\rho (p_3^0)v_{}^{}{}_{\rho }{}^{}(v.P_2+i\widehat{C})^1v\mathrm{"}_\nu (p_2^0)]>_{vv^{}v\mathrm{"}}`$ The first term in (91) corresponds to the time ordering $`X_1^0>X_2^0>X_3^0`$, the second term to the time ordering $`X_1^0>X_3^0>X_2^0`$. The amplitude is symmetric in the exchange $`23`$ of all indices (momentum, Lorentz, colour), it has no other symmetry. The 3-point amplitude is a Retarded one, on the left handside of (91) each momentum has been given an index R or A according to the rule stated in Sec III A. It is worth noticing that the energy $`p_i^0`$ that enters the numerator in each term of (93) is the one associated with the earliest time, a consequence of the expression (80) for the induced current. One may reverse the whole string of operators in $`𝐯`$ space. When $`\widehat{C}`$ is replaced by $`ϵ`$, one recovers Eq.(5.22) of . The Ward identity stemming from the conserved current (85) writes $$p_1^\mu V_{\mu \nu \rho }(P_{1R},P_{2A},P_{3A})=if^{123}[\mathrm{\Pi }_{\nu \rho }((P_1+P_2)_R,P_{3A})\mathrm{\Pi }_{\nu \rho }((P_1+P_3)_R,P_{2A})]$$ (94) It is obeyed by the expression (93) in a straightforward way since from Eq.(21) $$<v.P_1(v.P_1+i\widehat{C})^1=<$$ Moreover there is a Ward identity with respect to $`P_2`$ (or $`P_3`$). Indeed, in (91) when $`v_\nu `$ sits in the middle of a string of operators in $`𝐯`$ space, one writes $$v.P_2=(v.(P_1+P_2)+i\widehat{C})(v.P_1+i\widehat{C})$$ so that $$(v.P_1+i\widehat{C})^1v.P_2(v.(P_1+P_2)+i\widehat{C})^1=(v.P_1+i\widehat{C})^1(v.(P_1+P_2)+i\widehat{C})^1$$ (95) When $`v_\nu `$ sits at the end of a string one uses $$(v.P_2+i\widehat{C})^1v.P_2>=>$$ and one gets from (91) $`p_2^\nu V_{\mu \nu \rho }^{123}=if^{123}m_D^2\{`$ (97) $`<v_\mu (v.(P_1+P_2)+i\widehat{C})^1v_\rho >(p_3^0)+<v_\mu (v.P_1+i\widehat{C})^1v_\rho >(p_3^0p_2^0)\}`$ $$p_2^\nu V_{\mu \nu \rho }^{123}=if^{123}[\mathrm{\Pi }_{\mu \rho }(P_{1R},(P_2+P_3)_A)\mathrm{\Pi }_{\mu \rho }((P_1+P_2)_R,P_{3A})]$$ (98) To summarize, $`V_{\mu \nu \rho }^{123}`$ obeys tree-like Ward identities with respect to all legs, and this property will generalize to the amplitudes $`V(X;X_1,X_2,\mathrm{}X_n)`$ although the conservation law for the current enforces it only for the $`X`$ leg (see the 4-point vertex in the next Sec.). The reason is, the expressions are associated with forward-scattering tree diagrams of a collective excitation, i.e. of the $`W`$ field. If one thinks about the complex conjugate amplitude, where $`X^0`$ is the earliest time, one may interpret Eq.(91) as follows. The incoming soft gluon $`P_1`$ excites a collective excitation of momentum $`P_1`$ which propagates, emits successively the soft gluons $`(P_2)`$ and $`(P_3)`$ and disappears at the last emission. This interpretation is valid at the HTL level as an alternative to the hard-loop interpretation. When the scale $`gT`$ is integrated out, the new feature is that it is the only surviving one. In contrast, in the one-loop interpretation, there is a symmetry with respect to all legs, so that the leg $`P_{1R}`$ could sit in the middle (rather than sitting first), there would be a fork in the $`ϵ`$-flow (an allowed fact) and one can easily check that the Ward identity (94) stemming from the conserved current would not be satisfied when $`ϵ`$ is replaced by $`\widehat{C}`$. It is also apparent on the one-loop 4-point function. ### D Four-gluon Vertices #### 1 The vertex RAAA For the 4-point amplitude, the result of the functional differentiation is $`V_{\mu \nu \rho \sigma }^{1234}(P_{1R},P_{2A},P_{3A},P_{4A})=m_D^2<v_\mu (v.P_1+i\widehat{C})^1`$ (100) $`[f^{12m}f^{34m}v_\nu (v.(P_1+P_2)+i\widehat{C})^1v_\rho (v.P_4+i\widehat{C})^1v_\sigma (p_4^0)+5\mathrm{t}erms]>_{allv}`$ where $`+5`$ terms mean the addition of the permutations $`ij`$ (in all indices) in any pair among the legs $`2,3,4`$. The written term corresponds to the time order $`X_1^0>X_2^0>X_3^0>X_4^0`$, the order in colour indices and Lorentz indices follows the order in time. An interpretation may be given in terms of tree diagrams, the gluons $`(P_2),(P_3),(P_4)`$ are emitted by the collective excitation induced by the gluon $`P_1`$. This 4-gluon amplitude’s form was written in for the case $`\widehat{C}=0`$, i.e. for the HTL case in the Imaginary Time formalism. The following tree-like Ward identities are satisfied $`ip_1^\mu V_{\mu \nu \rho \sigma }^{1234}(P_{1R},P_{2A},P_{3A},P_{4A})=`$ (103) $`f^{12m}V_{\nu \rho \sigma }^{m34}((P_1+P_2)_R,P_{3A},P_{4A})+f^{13m}V_{\rho \nu \sigma }^{m24}((P_1+P_3)_R,P_{2A},P_{4A})`$ $`+f^{14m}V_{\sigma \nu \rho }^{m23}((P_1+P_4)_R,P_{2A},P_{3A})`$ a consequence of the conserved current, and for the leg $`P_4`$ (or $`P_3`$, or $`P_2`$) $`ip_4^\sigma V_{\mu \nu \rho \sigma }^{1234}(P_{1R},P_{2A},P_{3A},P_{4A})=`$ (106) $`f^{41m}V_{\mu \nu \rho }^{m23}((P_1+P_4)_R,P_{2A},P_{3A})+f^{42m}V_{\mu \nu \rho }^{1m3}(P_{1R},(P_2+P_4)_A,P_{3A})`$ $`+f^{43m}V_{\mu \nu \rho }^{12m}(P_{1R},P_{2A},(P_3+P_4)_A)`$ in a way similar to the case of the 3-point vertex and with the use of the Jacobi identity $$f^{12m}f^{34m}+f^{42m}f^{13m}+f^{32m}f^{41m}=0$$ (107) It has to be remembered that the 3-point vertex $`V_{\mu \nu \rho }^{123}(P_{1R},P_{2A},P_{3A})`$ as given by (93) is symmetric with respect to the two legs A, and that the R leg sits first in $`𝐯`$ space. It is worth emphasizing the origin of the tree-like structure of the $`n`$-point Retarded amplitudes. The existence of the covariant derivative is the only source of non-linearity in $`A(X)`$ because the mean gauge field enters linearly in the induced current in Eq. (80), just as in the HTL case. #### 2 The vertex RRAA What enters a one-loop self energy diagram is a 4-point function of the type RRAA; indeed a loop has to be maid by joining an R leg to an A leg. The 4-point function just obtained is of the type RAAA (and its complex conjugate ARRR). For the amplitudes of the type RRAA, as explained in Sec.III A, a new feature appears. The prescription $`ϵ_1>0,ϵ_2>0,ϵ_3<0,ϵ_4<0`$ allows for the energy variable $`p_1^0+p_3^0=(p_2^0+p_4^0)`$, $`ϵ_1+ϵ_3`$ to be either $`>0`$ or $`<0`$, and one has to sum over the two possible analytical continuations of the Imaginary Time amplitude in that variable with appropriate weights (see Eq.(67)). The same for the variable $`p_1^0+p_4^0=(p_2^0+p_3^0)`$. A consequence is that the Ward identities take forms somewhat different from the ones just written. For example, instead of Eq.(103), one has $`ip_1^\mu V_{\mu \nu \rho \sigma }^{1234}(P_{1R},P_{2R},P_{3A},P_{4A})=f^{12m}V_{\nu \rho \sigma }^{m34}((P_1+P_2)_R,P_{3A},P_{4A})`$ (111) $`+f^{13m}[{\displaystyle \frac{N(P_3,P_2+P_4)}{N(P_3,P_4)}}V_{\nu \rho \sigma }^{2m4}(P_{2R},(P_1+P_3)_R,P_{4A})`$ $`+{\displaystyle \frac{N(P_4,P_1+P_3)}{N(P_3,P_4)}}V_{\nu \rho \sigma }^{2m4}(P_{2R},(P_1+P_3)_A,P_{4A})]`$ $`+f^{14m}[34]`$ where $`[34]`$ means a term obtained from the factor that multiplies $`f^{13m}`$ by the exchange of all indices of $`3`$ and $`4`$. $`N(P_i,P_j)`$ is defined in Eq.(59), and from (68) $$N(P_3,P_2+P_4)+N(P_4,P_1+P_3)=N(P_3,P_4)$$ In the response approach, the amplitudes RRAA have to be extracted from correlation functions higher than the 2-point ones, i.e. they are beyond the truncation of the Schwinger-Dyson hierarchy made by Blaizot and Iancu . However the tree-like structure found for the RAA$`\mathrm{}`$A soft amplitudes very likely extends to all soft amplitudes. We have been able to write down a tree-like amplitude which satisfies all the needed requirements, i.e. \- direct causality flow along each tree \- symmetry with respect to the two legs of type R, and with respect to the two legs of type A \- Ward identity similar to Eq.(111) in any leg. Our procedure follows. At the HTL level, all 4-point amplitudes arise from different analytical continuations of the IT amplitude. At the one-loop level, it has been shown explicitely how the R/A formalism, which introduces a weight on the tree-like vertices RAA, do lead for the RRAA amplitudes to a specific combination of analytical continuations of the IT amplitude . We have started from the one-loop HTL 4-point amplitude, written a la Frenkel-Taylor , i.e. with a colour factor Tr$`(T_1T_2T_3T_4)`$, with appropriate $`ϵ`$-flow and weight $`N(P_i,P_j)`$ for the case RRAA. Summing over 6 permutations, the colour structure has been reduced to the simplest one, i.e. to factors such as $`f^{12m}f^{34m}`$. Then, using partial fraction expansion, it has been transformed into a direct $`ϵ`$-flow along each tree. By construction, this amplitude satisfies all the required properties listed above. Then a form has emerged that generalizes to the case when each $`ϵ`$ is replaced by the operator $`\widehat{C}`$ in $`𝐯`$ space, and satisfies all requirements. It is $`V_{\mu \nu \rho \sigma }^{1234}(P_{1R},P_{2R},P_{3A},P_{4A})N(P_3,P_4)m_D^2={\displaystyle \frac{1}{2}}f^{12m}f^{34m}N(P_3,P_4)`$ (118) $`<\{v_\mu (v.P_1+i\widehat{C})^1v_\nu v_\nu (v.P_2+i\widehat{C})^1v_\mu \}(v.(P_1+P_2)+i\widehat{C})^1`$ $`\{v_\rho (v.P_4+i\widehat{C})^1v_\sigma (p_4^0)v_\sigma (v.P_3+i\widehat{C})^1v_\rho (p_3^0)\}>_{allv}`$ $`+f^{13m}f^{24m}(p_1^0+p_3^0)`$ $`\{N(P_3,P_2+P_4)<v_\mu (v.P_1+i\widehat{C})^1v_\rho (v.(P_1+P_3)+i\widehat{C})^1v_\nu (v.P_4+i\widehat{C})^1v_\sigma >_{allv}`$ $`N(P_4,P_1+P_3)<v_\nu (v.P_2+i\widehat{C})^1v_\sigma (v.(P_2+P_4)+i\widehat{C})^1v_\mu (v.P_3+i\widehat{C})_1v_\rho >_{allv}\}`$ $`+f^{14m}f^{23m}(p_1^0+p_4^0)\{34\}+`$ where $`\{34\}`$ means a term obtained from the factor multiplying $`f^{13m}f^{24m}`$ by the exchange of all indices of $`3`$ and $`4`$ , $`N(P_i,P_j)`$ is defined in (59) and $``$ is another term that depends on the same propagators that the ones in the factor $`f^{12m}f^{34m}`$. $``$ is written explicitely in Appendix B, together with some details on the Ward identities. The terms written explicitely in (118) may be seen as a naive extension of the RAAA case; in the factor of $`f^{12m}f^{34m}`$, $`X_1^0`$ and $`X_2^0`$ are the later times, $`X_3^0`$ and $`X_4^0`$ the earlier ones; in the factor of $`f^{13m}f^{24m}`$, the first term corresponds to the time ordering $`X_1^0>X_3^0>X_2^0>X_4^0`$, the second one to $`X_2^0>X_4^0>X_1^0>X_3^0`$. What is needed for a self-energy diagram is the forward amplitude with two identical colours, for example colour $`2`$ = colour $`3`$ and $`P_2+P_3=0=P_1+P_4`$. The terms $`f^{14m}f^{23m}`$ disappear. Moreover it will be shown later on that only the term whose factor is $`N(P_3,P_2+P_4)`$ contributes to a soft loop. ## IV The one-loop self-energy diagrams with effective vertices ### A A general expression for the diagram with 3-gluon vertices At the HTL level, the Retarded/Advanced formalism is an alternative to the Imaginary Time formalism. It avoids the analytic continuations that are necessary in the Imaginary Time formalism. It has been used for explicit calculations of the photon self-energy within the HTL framework at the two-loop level by Aurenche and collaborators . We adopt their conventions. The propagators are the retarded and advanced effective propagators, i.e. with $`ϵ>0`$ $$\mathrm{\Delta }(P_R)=\mathrm{\Delta }(p_0+iϵ,p)=\frac{i}{(p_0+iϵ)^2p^2\mathrm{\Pi }(p_0+iϵ,p)}$$ (1) $$\mathrm{\Delta }(P_A)=\mathrm{\Delta }(p_0iϵ,p)=\mathrm{\Delta }(P_R)^{}\mathrm{\Delta }(P_R)=\mathrm{\Delta }((P)_A)$$ (2) from Eqs.(54,55) for $`\mathrm{\Pi }(P)`$. In covariant gauges, the effective gluon propagator is $$D^{\mu \nu }(P)=𝒫_t^{\mu \nu }\mathrm{\Delta }^t(P)+𝒫_l^{\mu \nu }\mathrm{\Delta }^l(P)+i\xi \frac{p^\mu p^\nu }{P^4}$$ (3) $`𝒫_t^{ij}`$ $`=`$ $`\delta ^{ij}\widehat{p}^i\widehat{p}^j,𝒫_t^{\mu 0}=0`$ (4) $`𝒫_l^{\mu \nu }`$ $`=`$ $`g^{\mu \nu }{\displaystyle \frac{p^\mu p^\nu }{P^2}}𝒫_t^{\mu \nu }`$ (5) $$\mathrm{\Pi }^{\mu \nu }(P)=𝒫_t^{\mu \nu }(P)\mathrm{\Pi }^t+𝒫_l^{\mu \nu }(P)\mathrm{\Pi }^l(P)$$ (6) In a Coulomb-like gauge $$D^{\mu \nu }(P)=𝒫_t^{\mu \nu }\mathrm{\Delta }^t(P)+g_{\mu 0}g_{\nu 0}\frac{P^2}{p^2}\mathrm{\Delta }^l(P)+i\xi \frac{p^\mu p^\nu }{p^4}$$ (7) The three-gluon vertex has been obtained in Sec.III C, i.e. Eq.(93). In this section it will be written $$gV_{\mu \rho \sigma }^{123}(Q_R,P_A,S_A)=if^{123}gV_{\mu \rho \sigma }(Q_R,P_A,S_A)$$ (8) and the $`i`$ is dropped to be consistent with the convention for the propagator (1). As detailed in Sec.III A, the 3-point vertices to be used in the R/A formalism are $`\mathrm{\Gamma }^{\mu \rho \sigma }(Q_A,P_R,R_R)`$ $`=`$ $`V^{\mu \rho \sigma }(Q_A,P_R,R_R)`$ (9) $`\mathrm{\Gamma }^{\mu \rho \sigma }(Q_R,P_A,R_A)`$ $`=`$ $`V^{\mu \rho \sigma }(Q_R,P_A,R_A)(n(p_0)+n(r_0)+1)`$ (10) $`V^{\mu \rho \sigma }(Q_R,P_A,R_A)`$ is antisymmetric in the exchange of the two legs of type A, it satisfies the general properties Eqs.(51,52). Two relations will be useful $`V^{\mu \rho \sigma }((Q)_A,(P)_R,(R)_R)`$ $`=`$ $`V^{\mu \rho \sigma }(Q_R,P_A,R_A)`$ (11) $`n(p_0)+1/2`$ $`=`$ $`(n(p_0)+1/2)`$ (12) When one considers the self-energy $`\mathrm{\Pi }(Q_R)`$, the one-loop diagram with 3-gluon vertices is, in the R/A formalism, the sum of the three diagrams drawn on Fig.1(a)(b)(c) (a propagator joins an $`R`$ leg to an $`A`$ leg, see Sec.III A). With the definition of momenta of Fig.1(d), leaving out a colour factor $`\delta _{ab}`$ $`\mathrm{\Pi }_{(3g)}^{\mu \nu }(Q_R)={\displaystyle \frac{Ng^2}{2}}{\displaystyle \frac{d_4p}{i(2\pi )^4}}`$ (16) $`(n(p_0)+1/2)[D_{\rho \rho ^{}}(P_R)V^{\rho \mu \sigma }(P_R,Q_R,(S)_A)D_{\sigma \sigma ^{}}(S_R)V^{\rho ^{}\nu \sigma ^{}}((P)_A,(Q)_A,S_R)]`$ $`(n(s_0)+1/2)[D_{\rho \rho ^{}}(P_A)V^{\rho \mu \sigma }(P_A,Q_R,(S)_R)D_{\sigma \sigma ^{}}(S_A)V^{\rho ^{}\nu \sigma ^{}}((P)_R,(Q)_A,S_A)]`$ $`(n(p_0)n(s_0))[D_{\rho \rho ^{}}(P_A)V^{\mu \rho \sigma }(Q_R,P_A,(S)_A)D_{\sigma \sigma ^{}}(S_R)V^{\nu \rho ^{}\sigma ^{}}((Q)_A,(P)_R,S_R)]`$ Several comments have to be made \- From (11) the two vertices entering each term are identical, up to a sign \- The prescription $`R`$ or $`A`$ on each leg fixes the way one should approach the discontinuity in that variable. Then one writes $`S=P+Q`$ and one may integrate over $`p_0`$ either on the real axis, or in the complex $`p_0`$ plane. Although $`s_0=p_0+q_0`$ , it make sense to speak of discontinuities in the $`p_0`$ or $`q_0`$ or $`s_0`$ variable, just as at $`T=0`$ one considers singularities in the $`s`$ or $`t`$ or $`u`$ channel of the 4-point function. \- The properties in the complex $`p_0`$ plane to be used in the subsequent argument are identical whether $`(v.P_R)^1`$ enters the HTL vertices, or $`(v.P+i\widehat{C})^1`$ enters the soft vertices. Indeed, when $`p_0+iϵ`$ enters, the meaning is that one should approach the singularities along the real $`p_0`$ axis from above. The operator $`\widehat{C}`$ has an infinite tower of positive eigenvalues $`c_l`$ and a zero eigenvalue $`c_0`$, i.e. $`(v.P+i\widehat{C})`$ vanishes for $`p_0=𝐯.pic_l`$, i.e. those points are in the lower complex $`p_0`$ plane or along the real axis. \- The third term of the right handside of Eq.(16) corresponds to diagram(c) of Fig.1. In the first and second term, a term $`(n(q_0)+1/2)`$ has been dropped in the thermal weight. The reason is: all factors in the bracket of the first (or second) term have singularities in the $`p_0`$ complex plane only on one side of the real axis; by closing the integration contour on the other half plane, they are seen to give a vanishing contribution when they are multiplied by $`(n(q_0)+1/2)`$. By the same token $`1/2`$ coul be dropped, it has been kept because of the parity property (12) . Conversely, $`n(p_0)`$ has poles at $`p_0=\pm in2\pi T`$ all along the imaginary axis. Thermal weights depending on the external momenta do appear in the R/A formalism. They give a vanishing contribution at any loop order because they are associated with terms whose singularities all lie one one side of the real axis. It has to be so in order to agree with the Imaginary Time formalism, where thermal weights only depend on loop momenta (see the coth method ). This feature will turn out to be an useful one when one considers soft momentum exchange $`p`$. \- Eq. (16) is a completely general formula and appears in many cases . ### B The case of a soft exchange around the loop We want to use Eq.(16) to compute a soft momentum exchange around the loop, i.e. $`p_0<<T`$. Now comes Bödeker’s argument (translated from Imaginary Time to R/A formalism). For loop momenta $`p_0<<T`$ one may drop the first and second term in Eq.(16). Indeed, as said above, in the first term, the term in brackets has singularities only on one side of the real $`p_0`$ axis, and if one closes the integration contour on the other side, the first contribution comes from the pole $`p_0=i2\pi T`$ of $`n(p_0)`$ i.e. a hard energy, and it will lead to terms of order $`q_0/T`$. The residue of the pole $`p_0=0`$ of $`n(p_0)`$ is zero as the two terms in (16) whose factor is $`n(p_0)`$ cancel each other at that point ($`\mathrm{\Pi }(p_0=0,p)=0`$, the vertices are identical up to a sign). In what follows, only one of these two terms is kept, the contribution to be obtained will come from a region away from $`p_0=0`$. The same argument applies to the variable $`s_0`$ and the second term. As a consequence, one is left with the third term, and for $`p_0<<T`$ and $`q_0<<T`$, one may approximate the thermal weights i.e. keep only the pole close to the origin and neglect the other ones’ contribution. Changing the loop variable from $`P`$ to $`K`$ $$P=KQ/2,S=K+Q/2$$ (17) $$n(p_0)n(s_0)T(\frac{1}{p_0}\frac{1}{s_0})=\frac{Tq_0}{(k_0q_0/2)(k_0+q_0/2)}$$ (18) so that a factor $`q_0`$ is extracted from the thermal weight. Note that in terms of the set of diagrams drawn on Fig.1, this argument amounts to keep only the diagram where the vertex with two legs of type $`A`$ is not the one with a $`(Q)_A`$ leg. It will be a useful feature when one turns to the other case, i.e. one soft loop diagram with a 4-point vertex. First, momenta $`KgT`$ around the loop wil be considered and the result obtained in will be reproduced, then momenta $`Kg^2T\mathrm{ln}1/g`$ will be studied. The vertex $`V`$ is the total vertex, bare + effective. However diagrams containing only bare vertices (gluon + ghost), or containing one bare and one effective vertex have been shown in to give a subleading contribution, mostly because a bare vertex contains no factor $`1/v.Q`$ (or $`(v.Q+i\widehat{C})^1`$). At the HTL level, there is no effective ghost vertex, and it is likely to be the same for the case $`Kg^2T\mathrm{ln}1/g`$. We shall not consider the case $`\mu =\nu =0`$ in $`\mathrm{\Pi }^{\mu \nu }`$ as our aim is the collision operator. For momenta $`KgT`$ the effective vertex is the HTL one, Eqs.(8),(93) with $`P_1=Q,P_2=S=(K+Q/2),P_3=P=KQ/2`$ and $`\widehat{C}`$ replaced by $`ϵ(ϵ>0)`$ where $`=\delta _{S_2}(𝐯𝐯^{})`$ is the identity in $`𝐯`$ space. $$V^{\mu \sigma \rho }(Q_R,(S)_A,P_A)=m_D^2<\frac{v_1^\mu v_1^\sigma v_1^\rho }{v_1.Q_R}(\frac{(k_0+q_0/2)}{v_1.(K+Q/2)_R}\frac{(k_0q_0/2)}{v_1.(KQ/2)_A})>_{v_1}$$ (19) $$V^{\mu \sigma \rho }(Q_R,(S)_A,P_A)=m_D^2𝒱^{\mu \rho \sigma }(K,Q)$$ (20) and the resummed propagators (3,6) are those with the HTL $`\mathrm{\Pi }^{\mu \nu }`$, as in Eq.(32), so that keeping only the third term in Eq.(16). and with (17,18) $`\mathrm{\Pi }_{(3g)}^{\mu \nu }(Q_R)=q_0m_D^4{\displaystyle \frac{g^2NT}{2}}{\displaystyle \frac{d_4k}{i(2\pi )^4}\frac{1}{(k_0q_0/2)(k_0+q_0/2)}}`$ (22) $`D_{\rho \rho ^{}}((KQ/2)_A)D_{\sigma \sigma ^{}}((K+Q/2)_R)𝒱^{\mu \rho \sigma }(K,Q)𝒱^{\nu \rho ^{}\sigma ^{}}(K,Q)`$ $`𝒱^{\nu \rho ^{}\sigma ^{}}`$ is obtained from $`𝒱^{\mu \rho \sigma }`$ with the substitutions $`\mu \nu ,\rho \rho ^{},\sigma \sigma ^{},v_1v_2`$ If one now drops $`Qg^2T`$ in front of $`KgT`$ in (22,19) $$k_0(\frac{1}{v_1.K_R}\frac{1}{v_1.K_A})=k_0\mathrm{disc}\frac{1}{v_1.K}=k_02\pi i\delta (v_1.K)$$ (23) and $`\mathrm{\Pi }_{(3g)}^{\mu \nu }(Q_R)=q_0m_D^4{\displaystyle \frac{g^2NT}{2}}{\displaystyle \frac{d_4k}{i(2\pi )^4}D_{\rho \rho ^{}}(K_A)D_{\sigma \sigma ^{}}(K_R)}`$ (25) $`<v_1^\mu {\displaystyle \frac{1}{v_1.Q_R}}v_1^\rho v_1^\sigma \mathrm{disc}{\displaystyle \frac{1}{v_1.K}}>_{v_1}<v_2^\nu {\displaystyle \frac{1}{v_2.Q_R}}v_2^\rho ^{}v_2^\sigma ^{}\mathrm{disc}{\displaystyle \frac{1}{v_2.K}}>_{v_2}`$ The result does not depend on the gauge parameter $`\xi `$ owing to the $`\delta `$ functions. $`\mathrm{\Pi }_{(3g)}^{\mu \nu }`$ was found via this perturbative approach in in the Imaginary Time formalism. Comparing (25) with (42) for the case $`\mu =j,\nu =i`$, one identifies the contribution from this diagram to the collision operator $`\widehat{C}`$, or equivalently to $`\mathrm{\Phi }`$ defined in Eq.(9) $$\mathrm{\Phi }(𝐯_1.𝐯_2)=\frac{d_4k}{(2\pi )^4}|v_1.D(K).v_2|^2(2\pi )^2\delta (v_1.K)\delta (v_2.K)$$ (26) with $$v_1.D(K).v_2=(𝐯_1.𝒫_t.𝐯_2)\mathrm{\Delta }^t(K)+\frac{K^2}{k^2}\mathrm{\Delta }^l(K)$$ in both covariant and Coulomb gauges. $`\mathrm{\Phi }`$ was also found by other methods . $`\mathrm{\Phi }(𝐯_1.𝐯_2)`$ has been interpreted as the collision cross section of two fast particles with soft momentum exchange $`K`$; the $`\delta `$ functions enforce the conservation of energy-momentum at each vertex in the near-forward direction . The integration range for the space momentum $`k`$ is limited $`\mu _2<k<\mu _1`$ (see (50) ). Turning now to the case when the exchange momentum is $`Kg^2T\mathrm{ln}1/g`$ the gluon propagators and the 3-gluon vertices have to be modified. The polarization tensor is now as in (30) and its related vertex as in (8),(93) with $`P_1=Q,P_2=(K+Q/2),P_3=KQ/2`$. The only change in the expression (22) for the one-loop soft momentum exchange $`\mathrm{\Pi }_{(3g)}^{\mu \nu }`$ is in $`𝒱^{\mu \rho \sigma }`$ and $`𝒱^{\nu \rho ^{}\sigma ^{}}`$ $`𝒱_{}^{}{}_{}{}^{\mu \rho \sigma }(K,Q)=<v_1^\mu (v_1.Q+i\widehat{C})^1`$ (29) $`\{(k_0+q_0/2)v_1^\rho (v_1.(K+Q/2)+i\widehat{C})^1v_1^\sigma `$ $`(q_0/2k_0)v_1^\sigma (v_1.(Q/2K)+i\widehat{C})^1v_1^\rho \}>_{v_1,v_1^{},v_1^{\prime \prime }}`$ and $`𝒱_{}^{}{}_{}{}^{\nu \rho ^{}\sigma ^{}}`$ is obtained by the substitution $`\mu \nu ,\rho \rho ^{},\sigma \sigma ^{}`$ and $`v_1,v_1^{},v_1^{\prime \prime }v_2,v_2^{},v_2^{\prime \prime }`$ The Ward identities (94,98) imposed on this 3-gluon vertex read $$Q_\mu 𝒱_{}^{}{}_{}{}^{\mu \rho \sigma }=m_D^2(\mathrm{\Pi }^{\rho \sigma }((K+Q/2)_R)\mathrm{\Pi }^{\rho \sigma }((Q/2K)_R)$$ (30) $$(K+Q/2)_\sigma 𝒱_{}^{}{}_{}{}^{\mu \rho \sigma }=m_D^2(\mathrm{\Pi }^{\mu \rho }((Q/2K)_R)\mathrm{\Pi }^{\mu \rho }(Q_R))$$ (31) $`\mathrm{\Pi }_{(3g)}^{\mu \nu }`$ is in the desired form (48). The factors $`v_\mu (v.Q+i\widehat{C})^1\mathrm{}\mathrm{}(v.Q+i\widehat{C})^1v_\nu `$ come from the tree-like structure of the Retarded effective vertices where $`v_\mu (v.Q+i\widehat{C})^1`$ sits first in $`𝐯`$ space. If one now drops $`Qg^2T`$ in front of $`Kg^2T\mathrm{ln}1/g`$, which is only valid for $`\mathrm{ln}1/g>>1`$ $`𝒱_{}^{}{}_{}{}^{\mu \rho \sigma }(K,Q)=k_0<v_1^\mu (v_1.Q+i\widehat{C})^1`$ (33) $`\{v_1^\rho (v_1.K+i\widehat{C})^1v_1^\sigma v_1^\sigma (v_1.Ki\widehat{C})^1v_1^\rho \}>_{v_1,v_1^{},v_1^{\prime \prime }}`$ In particular, for $`\rho `$ and $`\sigma `$ spacelike $$𝒱_{}^{}{}_{}{}^{\mu ij}=k_0<v_1^\mu (v_1.Q+i\widehat{C})^1\{v_1^iW_R^j(K,𝐯_1)v_1^jW_A^i(K,𝐯_1)\}>_{v_1}$$ (34) where the collective field $`W_R^i(K,𝐯)`$ has been defined in (26) and $`W_A^i(K,𝐯)`$ in a similar way (see (15)). The resulting contribution to the polarization tensor is $$\mathrm{\Pi }_{}^{}{}_{(3g)}{}^{\mu \nu }(Q_R)=q_0m_D^4\frac{g^2NT}{2}\frac{d_4k}{i(2\pi )^4}D_{\rho \rho ^{}}(K_A)D_{\sigma \sigma ^{}}(K_R)\frac{1}{k_0^2}𝒱_{}^{}{}_{}{}^{\mu \rho \sigma }(K,Q)𝒱_{}^{}{}_{}{}^{\nu \rho ^{}\sigma ^{}}(K,Q)$$ (35) with $`𝒱^{}`$ as in (33) or (34). A consequence of (35) is, from (30) $`Q_\nu \mathrm{\Pi }_{}^{}{}_{(3g)}{}^{\mu \nu }=q_0m_D^2{\displaystyle \frac{g^2NT}{2}}{\displaystyle \frac{d_4k}{i(2\pi )^4}D_{\rho \rho ^{}}(K_A)D_{\sigma \sigma ^{}}(K_R)}`$ (37) $`{\displaystyle \frac{1}{k_0^2}}𝒱_{}^{}{}_{}{}^{\nu \rho ^{}\sigma ^{}}(K,Q)(\mathrm{\Pi }^{\rho ^{}\sigma ^{}}(K_R)\mathrm{\Pi }^{\rho ^{}\sigma ^{}}(K_A))`$ In Sec.IV D the other one-loop diagram with a 4-gluon vertex will be computed and one will obtain $$Q_\nu (\mathrm{\Pi }_{}^{}{}_{(3g)}{}^{\mu \nu }+\mathrm{\Pi }_{}^{}{}_{(4g)}{}^{\mu \nu })=0$$ (38) Comparing (35)( 33) with the expression (48) for the case $`\mu =j,\nu =i`$, one obtains the collision term arising from the diagram of Fig.1(d). Defining $$C^{}(𝐯_1,𝐯_2)=m_D^2\frac{g^2NT}{2}([\mathrm{\Phi }_{(3g)}^{}(𝐯_1,𝐯_2)+\mathrm{\Phi }_{(4g)}^{}(𝐯_1,𝐯_2)]$$ (39) $`\mathrm{\Phi }_{(3g)}^{}(𝐯_1,𝐯_2)={\displaystyle \frac{d_4k}{(2\pi )^4}D_{\rho \rho ^{}}(K_A)D_{\sigma \sigma ^{}}(K_R)}`$ (42) $`<v_1^\rho (v_1.K+i\widehat{C})^1v_{}^{}{}_{1}{}^{\sigma }v_1^\sigma (v_1.Ki\widehat{C})^1v_{}^{}{}_{1}{}^{\rho }>_{v_1^{}}`$ $`<v_2^\rho ^{}(v_2.K+i\widehat{C})^1v_{}^{}{}_{2}{}^{\sigma ^{}}v_2^\sigma ^{}(v_2.Ki\widehat{C})^1v_{}^{}{}_{2}{}^{\rho ^{}}>_{v_2^{}}`$ where the range of integration over the space momentum is $`\mu _3<k<\mu _2`$ (see(50)). The reality of $`\mathrm{\Phi }_{(3g)}^{}`$ is a consequence of the Bose symmetry of the two soft gluons in the effective vertices. Note that one may go back to the expression describing the momentum $`KgT`$ by replacing the operator $`\widehat{C}`$ by $`ϵ,ϵ>0`$, then the inverse operator becomes diagonal in $`𝐯`$ space, and one gets back $`\mathrm{\Phi }(𝐯_1,𝐯_2)`$ as in (26). The constraint $`v_1.K=0=k_0𝐯_1.𝐤`$ in Eq.(26) is essential for the gauge independence of $`\mathrm{\Phi }_{(3g)}`$ in two ways: i) the contraction $`k_\rho v_1^\rho `$ vanishes ii) $`(𝐯.𝐤)^2=k_0^2`$ is used to get the same result in covariant and Coulomb gauges. This is no more true for $`\mathrm{\Phi }_{(3g)}^{}`$. One has $`k_\rho <v_1^\rho (v_1.K+i\widehat{C})^1v_{}^{}{}_{1}{}^{\sigma }v_1^\sigma (v_1.Ki\widehat{C})^1v_{}^{}{}_{1}{}^{\rho }>_{v_1^{}}`$ (44) $`=<i\widehat{C}(v_1.K+i\widehat{C})^1v_{}^{}{}_{1}{}^{\sigma }>_{v_1^{}}`$ which is zero when integrated over $`v_1`$ or contracted with $`k_\sigma `$. For $`k_0,k<<\gamma `$, it reduces to $`v_1^\sigma `$. For a longitudinal gluon exchange, the result depends on the gauge parameter $`\xi `$ and on the gauge (covariant or Coulomb). As the transport equation (5) has been established in the strict Coulomb gauge (and expected to be gauge independent), one may wish to stay in this gauge. An intuitive picture for the collision term $`\mathrm{\Phi }_{(3g)}^{}(𝐯_1,𝐯_2)`$ is obtained if one limits oneself to the dominant exchange of transverse gluons. From (3) $$D^{\rho \rho ^{}}(K)\mathrm{\Delta }^t(K)𝒫_t^{\rho \rho ^{}}=\mathrm{\Delta }^t(K)𝒫_t^{ii^{}}$$ (45) then from (34), the collective field $`W^i(K,𝐯)`$ appears in $`\mathrm{\Phi }_{(3g)}^{}`$, its transverse part is parallel to $`v_t^i`$ $`W^a(K,𝐯)`$ $`=`$ $`iW^i(K,𝐯)E^{ia}(K)`$ (46) $`W^i(K,𝐯)`$ $`=`$ $`\widehat{k}_iW^l+v_t^iW^t`$ (47) $`𝒫_t^{ii^{}}(k)W^i^{}(K,𝐯)`$ $`=`$ $`𝒫_t^{ii^{}}v^i^{}W^t(K,𝐯)`$ (48) so that $`\mathrm{\Phi }_{(3g)}^{}(𝐯_1,𝐯_2)={\displaystyle }{\displaystyle \frac{d_4k}{(2\pi )^4}}|\mathrm{\Delta }^t(K_R)(𝐯_1.𝒫_t.𝐯_2)|^2`$ (50) $`()[W_R^t(K,𝐯_1)W_A^t(K,𝐯_1)][W_R^t(K,𝐯_2)W_A^t(K,𝐯_2)]`$ $`\mathrm{\Phi }_{(3g)}^{}(𝐯_1,𝐯_2)`$ may be interpreted as the near-forward collision cross section of two collective excitations. The constraint at each vertex is no more a strict particle-like conservation $`2\pi \delta (v_1.K)`$ as in (26), but a smeared one involving the “width” of the $`W^t`$ field. A consequence of (42) is $`{\displaystyle \frac{d\mathrm{\Omega }_{𝐯_2}}{4\pi }\mathrm{\Phi }_{(3g)}^{}(𝐯_1,𝐯_2)}={\displaystyle \frac{d_4k}{(2\pi )^4}D_{\rho \rho ^{}}(K_A)D_{\sigma \sigma ^{}}(K_R)\frac{1}{k_0m_D^2}}`$ (52) $`(\mathrm{\Pi }^{\rho ^{}\sigma ^{}}(K_R)\mathrm{\Pi }^{\rho ^{}\sigma ^{}}(K_A))<v_1^\rho (v_1.K+i\widehat{C})^1v_{}^{}{}_{1}{}^{\sigma }v_1^\sigma (v_1.Ki\widehat{C})^1v_{}^{}{}_{1}{}^{\rho }>_{v_1^{}}`$ One notices that the contribution to the damping rate of a hard gluon arising from the range $`Kg^2T\mathrm{ln}1/g`$ is obtained from (52) if one substitutes to the $`𝐯_1^{}`$ average the quantity $`v_1^\rho v_1^\sigma (2\pi i)\delta (v_1.K)`$ . In Sec. IV D , one will find that $$\frac{d\mathrm{\Omega }_{𝐯_2}}{4\pi }[\mathrm{\Phi }_{(3g)}^{}(𝐯_1,𝐯_2)+\mathrm{\Phi }_{(4g)}^{}(𝐯_1,𝐯_2)]=0$$ (53) i.e. a relation similar to (10) for the case of the $`\widehat{C}`$ operator. ### C A sum rule for the collision operators $`\widehat{C}`$ and $`\widehat{C}^{}`$ In order to know the scale of $`\mathrm{\Phi }_{(3g)}^{}`$, and of $`\widehat{C}^{}`$ defined in (39), one may average over $`𝐯_1`$ and $`𝐯_2`$ $`\overline{\mathrm{\Phi }}={\displaystyle \frac{d\mathrm{\Omega }_{𝐯_1}}{4\pi }\frac{d\mathrm{\Omega }_{𝐯_2}}{4\pi }\mathrm{\Phi }_{(3g)}^{}(𝐯_1,𝐯_2)}`$ (56) $`={\displaystyle \frac{d_4k}{(2\pi )^4}D_{\rho \rho ^{}}^{}(K_R)D_{\sigma \sigma ^{}}(K_R)}`$ $`[\mathrm{\Pi }^{\rho ^{}\sigma ^{}}(K_R)\mathrm{\Pi }^{\rho ^{}\sigma ^{}}(K_A)][(\mathrm{\Pi }^{\rho \sigma }(K_R)\mathrm{\Pi }^{\rho \sigma }(K_A)]({\displaystyle \frac{1}{k_0^2m_D^4}})`$ This relation, found in the case $`Kg^2T\mathrm{ln}1/g`$ , also holds in the case $`KgT`$, i.e. when $`\mathrm{\Phi }_{(3g)}^{}(𝐯_1,𝐯_2)`$ is replaced by $`\mathrm{\Phi }(𝐯_1,𝐯_2)`$. Indeed, for $`KgT`$ , $`\mathrm{\Pi }^{\rho \sigma }(K)`$ is as in (32) i.e. $$\frac{i}{k_0m_D^2}[\mathrm{\Pi }^{\rho \sigma }(K_R)\mathrm{\Pi }^{\rho \sigma }(K_A)]=\frac{d\mathrm{\Omega }_{𝐯_1}}{4\pi }v_1^\rho v_1^\sigma 2\pi \delta (v_1.K)$$ (57) and (57) inserted into (56) leads to $`\mathrm{\Phi }(𝐯_1,𝐯_2)`$ as in (26). $`\overline{\mathrm{\Phi }}`$ is a gauge independent quantity $$\overline{\mathrm{\Phi }}=\frac{d_4k}{(2\pi )^4}[2|\mathrm{\Delta }^t|^2[\frac{2\mathrm{I}m\mathrm{\Pi }^t}{k_0m_D^2}]^2+|\mathrm{\Delta }^l|^2[\frac{2\mathrm{I}m\mathrm{\Pi }^l}{k_0m_D^2}]^2]$$ (58) Relation (58) allows an easy comparison between the infrared behaviour of the cases $`KgT`$ and $`Kg^2T\mathrm{ln}1/g`$. One restricts oneself to the dominant transverse gluon exchange. For $`k<m_D`$, $`|\mathrm{\Delta }^t(K)|^2`$put a strong weight upon the region $`k_0<<k`$ and one may write $$\mathrm{I}m\mathrm{\Pi }^t(k_0,k)k_0\mathrm{I}m\stackrel{~}{\mathrm{\Pi }}^t(k_0=0,k)$$ $$|\mathrm{\Delta }^t(K_R)|^2=\frac{1}{(k_0^2k^2\mathrm{Re}\mathrm{\Pi }^t)^2+(\mathrm{Im}\mathrm{\Pi }^t)^2}\frac{1}{k^4+k_0^2(\mathrm{Im}\stackrel{~}{\mathrm{\Pi }}^t)^2}$$ (59) $$\frac{dk_0}{k^4+k_0^2(\mathrm{Im}\stackrel{~}{\mathrm{\Pi }}^t)^2}=\frac{\pi }{k^2|\mathrm{Im}\stackrel{~}{\mathrm{\Pi }}^t|}$$ (60) so that from (58) $$\overline{\mathrm{\Phi }}\frac{dkk^2}{4\pi ^3}\frac{\pi }{k^2|\mathrm{Im}\stackrel{~}{\mathrm{\Pi }}^t|}2(\frac{2\mathrm{I}\mathrm{m}\stackrel{~}{\mathrm{\Pi }}^t}{m_D^2})^2$$ (61) $$\overline{\mathrm{\Phi }}\frac{1}{m_D^2}\frac{2}{\pi ^2}𝑑k\frac{|\mathrm{Im}\mathrm{\Pi }^t|}{k_0m_D^2}|_{k_0=0}(k)$$ (62) \- For the region $`kgT`$, from (57) $$\mathrm{Im}\mathrm{\Pi }^t=k_0m_D^2\frac{1}{k}\frac{\pi }{4}(1\frac{k_0^2}{k_2})\theta (k^2k_0^2)m_D^2\frac{k_0}{k}\frac{\pi }{4}=k_0\mathrm{Im}\stackrel{~}{\mathrm{\Pi }}^t$$ (63) so that $$\overline{\mathrm{\Phi }}_1\frac{1}{m_D^2}\frac{1}{2\pi }_{\mu _2}\frac{dk}{k}$$ (64) the integral has an infrared logarithmic divergence ($`\mu _2,\mu _3`$ have been defined in (50)) and $$\gamma =m_D^2\frac{g^2NT}{2}\overline{\mathrm{\Phi }}\frac{g^2NT}{4\pi }\mathrm{ln}\frac{m_D}{\mu _2}$$ (65) \- For the region $`kg^2T\mathrm{ln}1/g`$ , from Eq.(A7) in Appendix A $$\mathrm{\Pi }^t(k_0,k)m=\frac{m_D^2k_0}{3}\mathrm{\Sigma }_1(k_0,k)$$ (66) where $`\mathrm{\Sigma }_1`$ is one matrix element of $`(v.K+i\widehat{C})^1`$ $$\overline{\mathrm{\Phi }}_2\frac{1}{m_D^2}\frac{2}{3\pi ^2}_{\mu _3}^{\mu _2}𝑑k|\mathrm{I}m\mathrm{\Sigma }_1(k_0=0,k)|$$ (67) From (A5,A1) one sees that the scale of $`\mathrm{\Sigma }_1`$ is $`\gamma `$ and for $`k<<\gamma `$ $$\mathrm{\Sigma }_1(k_0=0,k<<\gamma )=\frac{i}{\gamma \delta _1}\mathrm{a}nd\mathrm{Im}\stackrel{~}{\mathrm{\Pi }}^t(0,k)=\frac{m_D^2}{3}\frac{1}{\gamma \delta _1}$$ (68) so that $`\overline{\mathrm{\Phi }}_2`$ has no infrared divergence. In fact it is unlikely that the main contribution to $`\overline{\mathrm{\Phi }}_2`$ comes from the momenta $`k<\gamma /3`$. Indeed, as discussed in Appendix A, there is an imaginary part of $`\mathrm{\Pi }^t`$ in the range $`|k_0|<k`$ analogous to (63) i.e. to the Landau-damping term. However it only exists for $`k>\gamma /3`$. Eq.(67) sets the scale of the operator $`\widehat{C}^{}`$ as $`g^2NT`$ (see Eq.(39)) with a weak dependence on both cutoff. One may look at the contribution from the momenta $`kg^2T\mathrm{ln}1/g`$ (i.e. in Eq.(61), $`\mathrm{Im}\stackrel{~}{\mathrm{\Pi }}^t`$ in the denominator is as in (68)) to two other quantities \- the damping rate of a hard point-like gluon. Then in $`\overline{\mathrm{\Phi }}_2`$ as in (61), one $`\mathrm{Im}\stackrel{~}{\mathrm{\Pi }}^t`$ in the numerator is as in (68) and the other one as in (63), and $`\overline{\mathrm{\Phi }}_2`$ has an infrared log divergence. \- the interaction rate of two point-like hard gluons. Then in $`\overline{\mathrm{\Phi }}_2`$ both $`\mathrm{Im}\stackrel{~}{\mathrm{\Pi }}^t`$ in the numerator are as in (63), and $`\overline{\mathrm{\Phi }}_2`$ has a linear divergence, as stated in the introduction. ### D The diagram with a 4-gluon effective vertex The contribution to $`\mathrm{\Pi }_{\mu \nu }(Q)`$ of the diagram drawn on Fig.2(a) is considered. One treats simultaneously both cases i.e. when the momentum $`k`$ running around the loop is i) $`kgT`$ (where $`\widehat{C}`$ is replaced by $`ϵ`$) ii) $`kg^2T\mathrm{ln}1/g`$. As explained in Sec.III A, one has to join the leg $`P_{3A}`$ to the leg $`P_{2R}`$ in the 4-gluon vertex $$ig^2()N(P_3,P_4)V_{\mu \sigma \rho \nu }^{1234}(P_{1R},P_{2R},P_{3A},P_{4A})$$ given in Eq.(118) with $$P_3=P_2=K,P_1=P_4=Q$$ (69) (Note that the Lorentz indices $`\sigma `$ and $`\nu `$ have been exchanged compared to (118)). With colour $`2`$ = colour $`3`$, the term $`f^{23m}f^{14m}`$ disappears in (118), all the other colour factors are identical and $`_{2,m}f^{12m}f^{24m}=\delta ^{14}N`$. Leaving out the colour factor $`\delta ^{14}`$ $`\mathrm{\Pi }_{(4g)}^{{}_{}{}^{}\mu \nu }(Q_R)=ig^2Nm_D^2{\displaystyle \frac{d_4k}{i(2\pi )^4}}D_{\rho \sigma }(P_{2R}=(K)_R)N(P_3=K,P_4=Q)`$ (71) $`V_{\mu \sigma \rho \nu }^{1234}(P_{1R}=Q_R,P_{2R}=(K)_R,P_{3A}=K_A,P_{4A}=(Q)_A)`$ with $$N(P_i,P_j)=n(p_i^0)+n(p_j^0)+1=n(p_i^0)n(p_j^0)$$ (72) One may symmetrize the integrant by adding the term obtained in the change of variables $`KK`$ : $`F^{}(K)=(F(K)+F(K))/2`$. We now proceed in a way similar to the 3-gluon vertices’ case (see Sec.IV B) in order to kill most of the terms appearing in Eqs.(118),(B7) : (i) All the terms whose thermal weight depend on the external momentum $`Q`$ (i.e. $`n(p_4^0)=n(q_0))`$ give a vanishing contribution. In the complex $`k_0`$ plane, they have all their singularities on the same side of the real axis, and one may close the contour in the other half plane. Indeed the factor $`D_{\rho \sigma }((K)_R)=D_{\rho \sigma }(k_0+iϵ,𝐤)`$ is multiplied for example by $`<v_\sigma (v.P_2+i\widehat{C})^1v_\mu (v.(P_1+P_2)+i\widehat{C})^1v_\nu (v.P_3+i\widehat{C})^1v_\rho >=`$ (74) $`<v_\sigma (v.K+i\widehat{C})^1v_\mu (v.(QK)+i\widehat{C})^1v_\nu (v.K+i\widehat{C})^1v_\rho >`$ (ii) For a soft momentum running around the loop $`k_0<<T`$, one may drop all terms whose factor is $`N(P_4,P_i)=n(p_i^0)n(q_0)`$ , ($`P_i=P_3=K`$ or $`P_i=P_1+P_3=Q+K`$), because the first contribution will come from the pole $`p_i^0=2\pi T`$ and it will lead to terms of order $`q_0/T`$. There is no singularity when $`p_i^00`$ since \- the factor $`(p_1^0+p_3^0)N(P_4,P_1+P_3)`$ is regular as $`p_1^0+p_3^00`$ \- for $`p_3^0=k_00`$, most of the terms whose factor is $`n(p_3^0)`$ cancel, there remain two terms $$\mathrm{}v_\rho [(v.Q𝐯.𝐤+i\widehat{C})^1+(v.Q+𝐯.𝐤+i\widehat{C})^1]v_\sigma \mathrm{}q_0n(k_0)D_{\rho \sigma }(k_0=0,𝐤)$$ which cancel with the term obtained in the change of variable $`KK`$. (iii) The only surviving term in Eqs.(118),(B7) has singularities on both sides of the real $`k_0`$ axis, it is $`<v_\mu (v.P_1+i\widehat{C})^1v_\rho (v.(P_1+P_3)+i\widehat{C})^1v_\sigma (v.P_4+i\widehat{C})^1v_\nu >D_{\rho \sigma }((K)_R)=`$ (76) $`v_\mu (v.Q+i\widehat{C})^1v_\rho (v.(Q+K)+i\widehat{C})^1v_\sigma (v.Q+i\widehat{C})^1v_\nu >D_{\rho \sigma }(K+iϵ)`$ with a weight $`N(P_3,P_2+P_4)(p_1^0+p_3^0)=(n(p_3^0)n(p_1^0+p_3^0))(p_1^0+p_3^0)`$ (78) $`({\displaystyle \frac{T}{k_0}}{\displaystyle \frac{T}{k_0+q_0}})(k_0+q_0)={\displaystyle \frac{Tq_0}{k_0}}`$ This term is associated with the diagram drawn on Fig.2(b). One adds the term obtained in the substitution $`KK`$, i.e. $$\mathrm{}v_\rho (v.(QK)+i\widehat{C})^1v_\sigma \mathrm{}D_{\rho \sigma }((K)_R=(K)_A)(\frac{Tq_0}{k_0})$$ and one may complete with irrelevant terms (singularities in $`k_0`$ on same side) to obtain $`\mathrm{\Pi }_{}^{}{}_{(4g)}{}^{\mu \nu }(Q)=iq_0{\displaystyle \frac{g^2NT}{2}}m_D^2{\displaystyle \frac{d_4k}{i(2\pi )^4}\frac{1}{k_0}[D_{\rho \sigma }((K)_R)D_{\rho \sigma }(K_R)]}`$ (81) $`<v^\mu (v.Q+i\widehat{C})^1[v^\rho (v.(K+Q)+i\widehat{C})^1v^\sigma `$ $`+v^\sigma (v.(QK)+i\widehat{C})^1v^\rho ](v.Q+i\widehat{C})^1v^\nu >`$ Neglecting $`Q`$ in front of $`K`$, the resulting $`\mathrm{\Pi }_{}^{}{}_{(4g)}{}^{\mu \nu }`$ is $`\mathrm{\Pi }_{}^{}{}_{(4g)}{}^{\mu \nu }(Q)=iq_0{\displaystyle \frac{g^2NT}{2}}m_D^2{\displaystyle \frac{d_4k}{i(2\pi )^4}\frac{1}{k_0}[D_{\rho \sigma }(K_R)D_{\rho \sigma }(K_A)]}`$ (83) $`<v^\mu (v.Q+i\widehat{C})^1[v^\rho (v.K+i\widehat{C})^1v^\sigma v^\sigma (v.Ki\widehat{C})^1v^\rho ](v.Q+i\widehat{C})^1v^\nu >`$ A consequence is $`Q_\nu \mathrm{\Pi }_{}^{}{}_{(4g)}{}^{\mu \nu }(Q)=iq_0{\displaystyle \frac{g^2NT}{2}}m_D^2{\displaystyle \frac{d_4k}{i(2\pi )^4}\frac{1}{k_0}[D_{\rho \sigma }(K_R)D_{\rho \sigma }(K_A)]}`$ (85) $`<v^\mu (v.Q+i\widehat{C})^1[v^\rho (v.K+i\widehat{C})^1v^\sigma v^\sigma (v.Ki\widehat{C})^1v^\rho ]>`$ where Eq.(21) has been used. One sees that the factor that depends on $`v`$ in (85) is $`𝒱_{}^{}{}_{}{}^{\mu \nu \rho }(K,Q)/k_0`$ as in Eq.(33). From (3,6) $$D_{\rho \sigma }(K_R)D_{\rho \sigma }(K_A)=iD_{\rho \rho ^{}}(K_A)D_{\sigma \sigma ^{}}(K_R)(\mathrm{\Pi }^{\rho ^{}\sigma ^{}}(K_R)\mathrm{\Pi }^{\rho ^{}\sigma ^{}}(K_A))$$ (86) so that comparing with (37) one obtains $$Q_\mu \mathrm{\Pi }_{}^{}{}_{(4g)}{}^{\mu \nu }=Q_\mu \mathrm{\Pi }_{}^{}{}_{(3g)}{}^{\mu \nu }$$ (87) Moreover, comparing $`\mathrm{\Pi }_{}^{}{}_{(4g)}{}^{\mu \nu }(Q)`$ as in (83) with (48) for the case $`\mu =j,\nu =i`$ one extracts a collision term (see (39)) $`\mathrm{\Phi }_{}^{}{}_{(4g)}{}^{}(𝐯,𝐯^{})=i{\displaystyle \frac{d_4k}{(2\pi )^4}\frac{1}{k_0m_D^2}[D_{\rho \sigma }(K_R)D_{\rho \sigma }(K_A)]}`$ (89) $`[v^\rho (v.K+i\widehat{C})_{v,v^{}}^1v_{}^{}{}_{}{}^{\sigma }v^\sigma (v.Ki\widehat{C})_{v,v^{}}^1v_{}^{}{}_{}{}^{\rho }]`$ $`\mathrm{\Phi }_{}^{}{}_{(4g)}{}^{}`$ is an operator. When applied on a function $`>`$ that does not depend on $`𝐯^{}`$ $$\widehat{\mathrm{\Phi }}_{}^{}{}_{(4g)}{}^{}>=\frac{d\mathrm{\Omega }_{𝐯_1^{}}}{4\pi }\mathrm{\Phi }_{}^{}{}_{(4g)}{}^{}(𝐯_1,𝐯_1^{})=\frac{d\mathrm{\Omega }_{𝐯_2}}{4\pi }\mathrm{\Phi }_{}^{}{}_{(3g)}{}^{}(𝐯_1,𝐯_2)$$ (90) from (52) and (86), i.e. the collision operator has a zero mode $$\widehat{C}^{}>=\widehat{\mathrm{\Phi }}_{}^{}{}_{(4g)}{}^{}>+\widehat{\mathrm{\Phi }}_{}^{}{}_{(3g)}{}^{}>=0$$ (91) $`\mathrm{\Phi }_{}^{}{}_{(4g)}{}^{}`$ does not depend on the gauge parameter $`\xi `$, but the longitudinal gluon exchange does depend on the gauge (covariant or Coulomb) since $`k_0𝐯.𝐤`$. If one restrict oneself to transverse gluon exchange, (89) may be written $`\mathrm{\Phi }_{}^{}{}_{(4g)}{}^{}(𝐯,𝐯^{})=i{\displaystyle \frac{d_4k}{(2\pi )^4}\frac{1}{k_0m_D^2}[\mathrm{\Delta }^t(K_R)\mathrm{\Delta }^t(K_A)]}`$ (93) $`v_t^i[(v.K+i\widehat{C})^1(v.Ki\widehat{C})^1]_{vv^{}}v_{}^{}{}_{t}{}^{i}`$ To go back to the expression that corresponds to momentum exchange $`KgT`$, one replaces $`\widehat{C}`$ by $`ϵ`$ in (89) and one gets back the local term in $`\widehat{C}`$ (see (9)). Indeed, the identity operator in $`𝐯`$ space is $`=\delta _{S_2}(𝐯𝐯^{})`$, and one recognizes in (86,89) the hard gluon damping rate $`\gamma `$ if $`\widehat{C}`$ is replaced by $`ϵ`$. $`\mathrm{\Phi }_{}^{}{}_{(4g)}{}^{}`$ comes from the self-energy diagram drawn on Fig.2(b) involving a collective-excitation propagator and a soft-gluon propagator, it is interpreted as the damping of a collective excitation (at a scale larger than $`k^1(g^2T\mathrm{ln}1/g)^1`$), its scale is $`\overline{\mathrm{\Phi }}_2`$ discussed in Sec.IV C. ## V Properties of the Collision operator $`\widehat{C}^{}`$ As a result of Sec.IV, the collision operator $`\widehat{C}^{}`$ that takes into account the collisions with exchanged gluon $`kg^2T\mathrm{ln}1/g`$ is $$\widehat{C}^{}=C^{}(𝐯,𝐯^{})=m_D^2\frac{g^2NT}{2}(\mathrm{\Phi }_{}^{}{}_{(3g)}{}^{}(𝐯,𝐯^{})+\mathrm{\Phi }_{}^{}{}_{(4g)}{}^{}(𝐯,𝐯^{}))$$ (1) with $`\mathrm{\Phi }_{}^{}{}_{(3g)}{}^{}(𝐯,𝐯^{})`$ as in (42) , or (50) for transverse gluon exchange, $`\mathrm{\Phi }_{}^{}{}_{(4g)}{}^{}(𝐯,𝐯^{})`$ as in (89) , or (93). Both are functions of $`𝐯.𝐯^{}`$ as there is no other vector available. In this section one restricts oneself to the dominant transverse gluon exchange $`\mathrm{\Phi }_{}^{}{}_{(3g)}{}^{}(𝐯_1,𝐯_2)={\displaystyle \frac{d_4k}{(2\pi )^4}|\mathrm{\Delta }^t(K_R)|^2}`$ (4) $`<v_{1i}^t(v_1.K+i\widehat{C})^1v_{}^{}{}_{1j}{}^{t}v_{1j}^t(v_1.Ki\widehat{C})^1v_{}^{}{}_{1i}{}^{t}>_{v_1^{}}`$ $`<v_{2i}^t(v_2.K+i\widehat{C})^1v_{}^{}{}_{2j}{}^{t}v_{2j}^t(v_2.Ki\widehat{C})^1v_{}^{}{}_{2i}{}^{t}>_{v_2^{}}`$ $`\mathrm{\Phi }_{}^{}{}_{(4g)}{}^{}(𝐯,𝐯^{})=i{\displaystyle \frac{d_4k}{(2\pi )^4}\frac{1}{k_0m_D^2}[\mathrm{\Delta }^t(K_R)\mathrm{\Delta }^t(K_A)]}`$ (6) $`v_t^i[(v.K+i\widehat{C})^1(v.Ki\widehat{C})^1]v_{}^{}{}_{t}{}^{i}`$ where $`v_t^i=v^i\widehat{k}^i𝐯.\widehat{𝐤}`$ The operator $`C^{}(𝐯,𝐯^{})`$ shares many features of the operator $`C(𝐯,𝐯^{})`$: \- it commutes with the rotations in $`𝐯`$ space, its eigenvalues are $$c_l^{}=<P_l(𝐯.𝐯^{})C^{}(𝐯,𝐯^{})>_{v,v^{}}$$ (7) and its eigenvectors are $`Y_l^m(𝐯)`$, \- $`\widehat{C}^{}`$ is written as an integral over the soft momenta $`k`$ of the exchanged gluon in the near-forward collision of two collective excitations of direction $`𝐯`$ and $`𝐯^{}`$, ($`kgT`$ for $`\widehat{C}`$, $`kg^2T\mathrm{ln}1/g`$ for $`\widehat{C}^{}`$). All the dependance on $`𝐯`$ and $`𝐯^{}`$ is in the effective vertices, i.e. in the (smeared) energy-momentum conservation at the vertex \- the soft gluon exchange $`kg^2T\mathrm{ln}1/g`$ put a strong weight upon the region $`k_0(k^2/T)\mathrm{ln}1/g`$ and one can set $`k_0=0`$ in the effective vertices whose scale is $`k_0kg^2T\mathrm{ln}1/g`$ (with one exception). (For $`\widehat{C}`$ one also sets $`k_0=0`$ to leading order in the vertices). The resulting scale for $`c_l^{}`$ is $`g^2NT`$, with a weak dependance on both cutoff $`\mu _3<k<\mu _2`$. \- the collision operator is made of two terms. One term comes from a self-energy diagram, in fact it is the damping of the collective excitation at a scale larger than $`1/k`$. It dominates the spectrum of the operator, except for the $`l=0`$ eigenvalue where both terms cancel each other, giving rise to a zero eigenvalue. The contribution from the other term vanishes for $`l=1,3,5\mathrm{}`$ to leading order. The operators $`\widehat{C}`$ and $`\widehat{C^{}}`$ also show some differences \- the infrared behaviour of the eigenvalues differ. For a transverse gluon exchange, the $`c_l`$ have a logarithmic infrared divergence that shows up in both terms of the collision operator. The $`c_l^{}`$ are infrared finite with one exception: the $`l=1`$ eigenvalue has a linear infrared divergence (see underneath). \- the operator $`\widehat{C}`$ is expected to be gauge independent. As seen in Sec.IV B, $`\widehat{C}`$ takes an identical form in covariant and Coulomb gauges. For $`\widehat{C^{}}`$, the longitudinal soft gluon exchange is very likely gauge dependent, apparently a consequence of the finite lifetime of the collective excitations at a smaller scale. The collision operators $`\widehat{C}`$ and $`\widehat{C}^{}`$ are so similar that it is easy to take into account both types of collisions, i.e. when the exchanged gluon has $`kgT`$ and when $`kg^2T\mathrm{ln}1/g`$. One just has to substitute the operator $`\widehat{C}+\widehat{C}^{}`$ to the operator $`\widehat{C}`$ in all the relations written in Sec.II and Sec.III. The transport equation for the $`W`$ field is now $$(v.D+i\widehat{C}+i\widehat{C}^{})W=𝐯.𝐄$$ and from it, one deduces the induced current, then the softer amplitudes. These softer amplitudes are tree-like and they obey tree-like Ward identities. This approach, via the polarization tensor, proposes an interpretation of $`\widehat{C}`$ and of $`\widehat{C}^{}`$ somewhat different from the loss and gain interpretation that results from the transport equation via the Schwinger-Dyson approach . In that alternative approach, the Bose symmetry of the effective vertices with respect to two legs of the same type is the dominant constraint, as a consequence the colour factors differ from the ones that appear in the Schwinger-Dyson case. ### A The eigenvalues of the operator $`\widehat{C}^{}`$ In Appendix C, matrix elements such as $$M_{(4g)}^{(l)}(K)=<P_l(𝐯.𝐯^{})v_i^t[(v.K+i\widehat{C})^1(v.Ki\widehat{C})^1]v_{}^{}{}_{i}{}^{t}>_{v,v^{}}$$ (8) are expressed in terms of matrix elements of $`(v.K\pm i\widehat{C})^1`$, which are themselves continued fractions, fonctions of $`k_0,k`$ and of all the eigenvalues $`c_l`$ of the operator $`\widehat{C}`$. The case of the eigenvalue $`l=0`$ of $`\widehat{C}^{}`$ has been treated in Sec.IV C. Many encountered features are valid for all $`c_l^{}`$. For $`l=0`$, the contributions of both terms $`\mathrm{\Phi }_{(3g)}^{}`$ and $`\mathrm{\Phi }_{(4g)}^{}`$ are equal and opposite, it has been called $`\overline{\mathrm{\Phi }}_2`$. $`\overline{\mathrm{\Phi }}_2`$ has been expressed in terms of the imaginary part of the $`l=1,m=1`$ eigenvalue of $`(v.K+i\widehat{C})^1`$ which is finite as $`k0`$. One consequence is that the integral over $`k`$ in (4) and (6) is infrared finite (see Eq.(67)). One can see that there is no infrared divergence in the contribution of $`\mathrm{\Phi }_{(4g)}^{}`$ to $`c_l^{}`$ in the following way. The scale of the matrix elements of $`(v.K+i\widehat{C})`$ is the scale of $`\widehat{C}`$, i.e. $`\gamma g^2T\mathrm{ln}1/g`$. For $`k<<\gamma `$, the matrix elements of $`(v.K+i\widehat{C})^1(k_0+i\widehat{C})^1`$ do not depend on $`m=𝐥.\widehat{𝐤}`$, but only on $`l`$, they are the inverse of $`<P_l(𝐯.𝐯^{})(k_0+i\widehat{C})>_{v,v^{}}=k_0+ic_l`$. In Eq.(8), there is no dependance on $`m`$ for $`k<<\gamma `$ and one may write $`𝐯_t.𝐯_t^{}P_l(𝐯.𝐯^{})`$ $`=`$ $`{\displaystyle \frac{2}{3}}𝐯.𝐯^{}P_l(𝐯.𝐯^{})`$ (9) $`=`$ $`{\displaystyle \frac{2}{3}}{\displaystyle \frac{1}{2l+1}}[(l+1)P_{l+1}(𝐯.𝐯^{})+lP_{l1}(𝐯.𝐯^{})]`$ (10) i.e. for $`k<<\gamma `$, Eq. (8) becomes $$M_{(4g)}^{(l)}(k_0)=\frac{2}{3}\frac{1}{2l+1}[\frac{l+1}{k_0+ic_{l+1}}+\frac{l}{k_0+ic_{l1}}\mathrm{c}.c.]$$ (11) where c.c. means complex conjugate. One can set $`k_0=0`$ in $`M_{(4g)}^{(l)}(k_0)`$ and one concludes that, just as for the case $`l=0`$, the integral over $`k`$ has no infrared divergence as $`k0`$. And the region $`k<<\gamma `$ of the integrant does not depend on $`l`$ for large $`l`$ since $`c_l=\gamma \delta _l`$ with $`\delta _l\gamma /l`$. There is one exception: the case $`l=1`$ for $`c_l^{}`$, where one sees from (11) that the eigenvalue $`c_0=0`$ of the operator $`\widehat{C}`$ enters. As fully discussed in Appendix C 3, one finds a linear divergence as $`k0`$ in $`c_1^{}`$ (whatever the order of the limits $`k_00`$ and $`k0`$), whose origin is the zero eigenvalue of $`\widehat{C}`$ (see Eqs.(C30,C38)). The contribution from $`\mathrm{\Phi }_{(3g)}^{}`$ vanishes for $`l`$ odd, as the entering matrix elements vanish for $`k_0=0`$. For $`l`$ even, only the sector $`|m|=1`$ of $`(v.K+i\widehat{C})^1`$ enters, the contribution is infrared finite (see Appendix C 2). ### B The colour conductivity The properties of this quantity are first summarized for the case when $`gT`$ has been integrated out. In Sec.III B, the induced current $`j_\mu ^{ind}(X)`$ has been written in terms of the soft field $`E^b(Y)`$. One may introduce the conductivity tensor $$j_\mu ^a(X)=d_4Y\sigma _{\mu i}^{ab}(X,Y)E_i^b(Y)$$ (12) The comparison with Eq.(78) in Sec.III B gives $$\sigma _{\mu i}^{ab}(X,Y)=im_D^2<v_\mu G_{ret}^{ab}(X,Y;A;𝐯,𝐯^{})v_i^{}>_{v,v^{}}$$ (13) The linearized part is in momentum space $`\sigma _{\mu i}^{ab}(K)|_{A=0}`$ $`=`$ $`im_D^2<v_\mu (v.K+i\widehat{C})^1v_i^{}>_{v,v^{}}\delta ^{ab}`$ (14) $`\mathrm{\Pi }^{\mu i}(K)`$ $`=`$ $`ik_0\sigma ^{\mu i}(K)`$ (15) One introduces the transverse and the longitudinal part of the conductivity $`E^i(K)`$ $`=`$ $`\widehat{k}^iE_l+E_t^i`$ $`\sigma _t`$ $`=`$ $`{\displaystyle \frac{i}{k_0}}\mathrm{\Pi }^t={\displaystyle \frac{i}{2k_0}}𝒫_t^{ij}\mathrm{\Pi }^{ij}`$ $`\sigma _l`$ $`=`$ $`i{\displaystyle \frac{k_0}{k^2}}\mathrm{\Pi }^l=i{\displaystyle \frac{k^i}{k^2}}\mathrm{\Pi }^{0i}`$ $`\sigma _t`$ $`=`$ $`im_D^2{\displaystyle \frac{1}{2}}<v_t^i(v.K+i\widehat{C})^1v_t^i>`$ (16) $`\sigma _l`$ $`=`$ $`im_D^2{\displaystyle \frac{1}{k^2}}<k_0(v.K+i\widehat{C})^1𝐯.𝐤>`$ (17) With $`v.K=k_0𝐯.𝐤`$ and relation (21), the expression for $`\sigma _l`$ may be written in alternative forms $`\sigma _l`$ $`=`$ $`im_D^2{\displaystyle \frac{1}{k^2}}<𝐯.𝐤(v.K+i\widehat{C})^1𝐯.𝐤>`$ (18) $`\sigma _l`$ $`=`$ $`im_D^2{\displaystyle \frac{k_0}{k^2}}[k_0<(v.K+i\widehat{C})^1>1]`$ (19) i.e. in terms respectively of the matrix elements $`l=m=1`$ and $`l=m=0`$ of $`(v.K+i\widehat{C})^1`$ $`\sigma _t`$ $`=`$ $`{\displaystyle \frac{im_D^2}{3}}\mathrm{\Sigma }_1(k_0.k)`$ (20) $`\sigma _l`$ $`=`$ $`im_D^2{\displaystyle \frac{k_0}{k^2}}(k_0\mathrm{\Sigma }_0(k_0.k)1)`$ (21) In Appendix C, Eq.(C61) gives $`k_0(k_0\mathrm{\Sigma }_01)/k^2`$ as a compact continued fraction for an easy comparison with Eq.(A5) for $`\mathrm{\Sigma }_1`$. For $`k<\gamma /3`$, the expansion in $`k^2`$ of the continued fraction converges for all $`k_0`$ (see Appendix A) and one obtains $`\sigma _t(k_0,k<<\gamma )`$ $`=`$ $`{\displaystyle \frac{im_D^2}{3}}[{\displaystyle \frac{1}{k_0+ic_1}}+{\displaystyle \frac{1}{5}}{\displaystyle \frac{k^2}{(k_0+ic_1)(k_0+ic_2)}}+O(k^4)]`$ $`\sigma _l(k_0,k^2<<k_0\gamma )`$ $`=`$ $`{\displaystyle \frac{im_D^2}{3}}[{\displaystyle \frac{1}{k_0+ic_1}}+{\displaystyle \frac{4}{15}}{\displaystyle \frac{k^2}{(k_0+ic_1)(k_0+ic_2)}}+O(k^4)]`$ i.e. $`\sigma _t`$ and $`\sigma _l`$ differ by terms of order $`k^2/(k_0+ic_1)(k_0+ic_2)`$. For $`\sigma _l`$, a different limit is (see (C61)) $$\sigma _l(k_0\gamma <<k^2)=im_D^2\frac{k_0}{k^2}$$ We now examine how the collisions at the scale $`g^2T\mathrm{ln}1/g`$ affect the conductivities. As said at the beginning of Sec.V, the solution that includes both types of collisions is $$\mathrm{\Pi }^{ji}(K)=k_0m_D^2<v^j(v.K+i\widehat{C}+i\widehat{C}^{})^1v^i>$$ (22) As a consequence, in the expressions just written for $`\sigma _t,\sigma _l`$, i.e. Eqs.(16, 18, 19), one just have to substitute $`\widehat{C}+\widehat{C}^{}`$ to $`\widehat{C}`$, i.e. substitute $`c_l+c_l^{}`$ to $`c_l`$ in every continued fraction, in particular $$\sigma _t(k_0,k<<\gamma )=\sigma _l(k_0,k^2<<k_0\gamma )=\frac{m_D^2}{3}\frac{i}{k_0+i(c_1+c_1^{})}$$ (23) so that $$\sigma _t=\sigma _l=\sigma (k_0<<\gamma ,k<k_0)=\frac{m_D^2}{3}\frac{1}{c_1+c_1^{}}$$ (24) A Comparison with related work We now examine how these results are related to the pioneers’ work of Arnold and Yaffe . Their interest is in the next-to-leading-log contribution to the colour conductivity. Their approach makes use of effective theories. They are lead to compute the same pair of one-loop diagrams contributing to a self-energy in an effective static theory, in the Coulomb gauge. Their operator $`\widehat{O}(\mathrm{𝟎})`$ is identical to the collision operator $`i\widehat{C}^{}`$, if in $`\widehat{C}^{}`$ one makes the static approximation $`k_0=0`$ in the effective vertices, and one performs the integral over $`k_0`$ with only the weight of the soft transverse propagator; it is an approximation which is done at the very end in this work, they do it at the start. Their $`W`$ field propagator $$\widehat{G}_0(𝐤)=i(v.K+i\widehat{C})^1|_{k_0=0}$$ is a real quantity (see the matrix elements $`i\mathrm{\Sigma }_m(k_0=0,k)`$ in Eq.(A5)). They are lead to consider the same quantity $$<v_l\widehat{O}(\mathrm{𝟎})v_l>=ic_1^{}$$ and they find that the relevant matrix element is $$<v_lv_i\widehat{G}_0(𝐤)v_jv_l>𝒫_t^{ij}(k)$$ in agreement with $`M_{(4g)}^{(l=1)}(k_0=0,k)`$ in Eq.(8). The expression of this matrix element in terms of those of $`\widehat{G}_0(𝐤)`$ agrees with Eq.(C42) in Appendix C 3. They encounter the infrared linear divergence of $`c_1^{}`$ which disappears upon dimensional regularization. At the soft level ($`gT`$ integrated out), their conductivity tensor agrees with Eq.(18) for $`\sigma _l`$. However, in their approach, once the contribution of the momenta $`kg^2T\mathrm{ln}1/g`$ is included, the effective conductivity is not merely obtained by the substitution of $`c_1+c_1^{}=c_1(1+c_1^{}/c_1)`$ to $`c_1`$ in the expression for $`\sigma _l`$, another term is added arising from the small momentum expansion of their self-energy (see a detailed comparison after Eq.(C50) in App.C 3). ## VI Conclusion To the near-forward collision of two collective colour excitations of the thermal gluons $`pT`$ is associated a collision operator $`\widehat{C}`$ when the gluon exchanged during the collision has $`kgT`$. This operator has an infinite number of eigenvalues $`c_l`$ that may be interpreted as the multipole moments of a rate. The $`l=0`$ eigenvalue is zero, a cancellation between a damping term and another term, the other $`c_l`$ are dominated by the damping term, whose scale is $`g^2T\mathrm{ln}1/g`$. When inserted into the transport equation that describes the evolution at some space-time scale $`x`$ of the collective excitations, the collision term accounts for the effective damping of the excitations due to the integration over smaller scales. This work has presented two results. The first result is the explicit form of the effective $`n`$-gluon amplitudes when the scale $`T`$ and $`gT`$ are integrated out. These amplitudes exhibit remarkable properties. The picture that emerges follows. The central role is played by the collective colour excitations. The soft gluons of momentum $`k<<gT`$ are emitted by the collective excitations that occur at the scale $`1/k`$. These excitations are associated with the transport equation, not with the structures seen in the resummed gluon propagator (such as quasiparticle poles). The effective amplitudes are tree-like and they obey tree-like Ward identities. The propagator along the tree, the one of the collective excitation, propagates an infinite tower of damping rates and one undamped mode. All the rates show an infrared log divergence. The presence of the undamped mode (the eigenvalue $`c_0=0`$) turns out to be essential i) for the Ward identities to be satisfied, ii) for the strong similitude with the HTL amplitudes. The second result is, a new collision operator $`\widehat{C^{}}`$ that allows to take into account the collisions with soft gluon exchange $`kg^2T\mathrm{ln}1/g`$ has been found by means of a perturbative approach, i.e. by computing the one-soft-loop diagrams (with loop momentum $`kg^2T\mathrm{ln}1/g`$) which enter the polarization tensor. The operator $`\widehat{C^{}}`$ shares many features of the operator $`\widehat{C}`$ : a zero mode, an infinite number of eigenvalues $`c_l^{}`$ which are expressed in terms of the $`c_l`$. For a transverse gluon exchange, the landscape in the infrared has changed. The logarithmic divergence that was entering the damping rates $`c_l`$ has disappeared. The eigenvalues $`c_l^{}`$ are finite and of order $`g^2T`$ (with some dependence on $`\mathrm{ln}1/g`$). There is one exception, the $`l=1`$ eigenvalue exhibits a linear infrared divergence, which is linked to the zero mode of the operator $`\widehat{C}`$ (as a result of angular momentum combination). The properties of $`\widehat{C}`$ and $`\widehat{C^{}}`$ are so similar that it is easy to take into account the collisions with $`kg^2T\mathrm{ln}1/g`$. One just has to substitute $`\widehat{C}+\widehat{C^{}}`$ to $`\widehat{C}`$ in the transport equation, i.e. $`c_l+c_l^{}`$ to $`c_l`$ in its solution. All what has just been said for the $`n`$-gluon amplitudes where $`\widehat{C}`$ enters ($`T`$ and $`gT`$ integrated out) are valid for the softer amplitudes where $`\widehat{C}+\widehat{C^{}}`$ enters; they are tree-like and satisfy tree-like Ward identities. The quantities $`c_l+c_l^{}`$ are infrared finite except for the $`l=1`$ case. This $`l=1`$ eigenvalue dominates the polarization tensor at momentum $`q<<g^2T\mathrm{ln}1/g`$. One possible interpretation of this linear infrared divergence is that it is the signal of new physics occurring at the scale $`(g^2T)^1`$. The collision operator $`\widehat{C}`$ is gauge independent, in contrast the longitudinal gluon exchange in the operator $`\widehat{C^{}}`$ is very likely gauge dependent. It remains to be seen whether integrating out the scale $`g^2T\mathrm{ln}1/g`$ only amounts to take into account the collisions through $`\widehat{C^{}}`$. Also, this work sheads a new light on what kind of process is being summed by means of the operator $`\widehat{C}`$ or $`\widehat{C^{}}`$ in the polarization tensor . The soft gluon polarization tensor may be written as an infinite series in powers of $`\widehat{C}`$ (at scale $`q<<gT`$) or of $`\widehat{C^{}}`$ (at scale $`q<<g^2T\mathrm{ln}1/g`$). $`\widehat{C}`$ and $`\widehat{C^{}}`$ are obtained from a truncation of part of the effective vertices that enter the one-soft-loop diagrams. The truncation amounts to cut-off one initial propagator of the collective excitation. One term entering the collision operator has been shown to come from a self-energy diagram involving a collective excitation and a soft gluon (in fact a damping rate); the terms in the series in powers of that term have alternating factors, one collective excitation’s propagator, one self-energy factor, one propagator …, i.e. the series is just the resummation of the self-energy of the collective excitation. For the collision’s other term, the alternating factors are one collective excitation’s propagator, two soft gluons, one propagator …. Since the exchanged gluons have $`k_0<k`$, this is a $`t`$ channel picture; in the crossed channel, one collective excitation scatters and dies out, it is replaced by another collective excitation at a different space-time point, this transition rate involves a two-gluon exchange. This picture is scale invariant. When one goes from $`\widehat{C}`$ to $`\widehat{C^{}}`$ the change is in the collective excitation’s propagator, a straight-line propagation for $`\widehat{C}`$, fluctuations in the direction $`𝐯`$ for $`\widehat{C^{}}`$ as a result of the processes included in $`\widehat{C}`$. ### Acknowledgments The author wishes to thank E. Iancu for a useful discussion at INT and for his constant interest, and E. Petitgirard for an e-correspondence. Part of this work was done while the author stayed at the INT session ”Non-equilibrium Dynamics in Quantum Field Theory”(INT-99-3) and the author wishes to thank the INT and L. Yaffe for the extended hospitality and for generating many friendly, open discussions. The hospitality of the LAPTH-Annecy is acknowledged. ## A The analytic properties of $`\mathrm{\Pi }`$ in the complex $`k_0`$ plane ### 1 An explicit expression for $`\mathrm{\Pi }^t`$ and $`\mathrm{\Pi }^l`$ Arnold and Yaffe have written down the characteristics of the operator $`[k_0𝐯.𝐤+iC(𝐯,𝐯^{})]^1`$ for the case $`k_0=0`$. The extension to the case $`k_00`$ is straightforward. $`C(𝐯,𝐯^{})`$ commutes with the rotations in $`𝐯`$ space, its eigenvectors are the spherical harmonics $`\sqrt{4\pi }Y_l^m(𝐯)=|lm>`$ with the measure $`d\mathrm{\Omega }_𝐯/4\pi `$ $$\widehat{C}|lm>=c_l|lm>=(\gamma \delta _l)|lm>$$ (A1) $$c_l=<C(𝐯,𝐯^{})P_l(𝐯.𝐯^{})>_{vv^{}}c_0=0,c_l>0\mathrm{for}l>0$$ (A2) $`C(𝐯,𝐯^{})`$ is given in (9). If one chooses $`\widehat{𝐤}`$ as the $`z`$ axis $$𝐯.𝐤|lm>=k(b_l^{(m)}|l+1m>+b_{l1}^{(m)}|l1m>)$$ (A3) $$b_l^{(m)}=\sqrt{\frac{(l+1)^2m^2}{4(l+1)^21}}$$ (A4) For fixed $`m`$, the matrix $`[k_0𝐯.𝐤+iC(𝐯,𝐯^{})]`$ is a tri-diagonal matrix whose inverse is known. In particular, the element $$\mathrm{\Sigma }_m=<l=mm|(k_0𝐯.𝐤+iC(𝐯,𝐯^{}))^1|l=mm>$$ is the continued fraction $$\mathrm{\Sigma }_m(k_0,k)=\frac{1}{k_0+ic_m}\frac{1}{1b_m^{(m)2}{\displaystyle \frac{\rho _m^2}{1b_{m+1}^{(m)2}{\displaystyle \frac{\rho _{m+1}^2}{1b_{m+2}^{(m)2}{\displaystyle \frac{\rho _{m+2}^2}{1...}}}}}}}$$ (A5) where $$\rho _m^2=\frac{k^2}{(k_0+ic_m)(k_0+ic_{m+1})}$$ (A6) the transverse part of the self-energy is $$\mathrm{\Pi }_R^t(k_0,k)=\frac{m_D^2}{3}k_0\mathrm{\Sigma }_1(k_0,k)$$ (A7) since from (30),(6) $$\mathrm{\Pi }_R^t(k_0,k)=m_D^2k_0<v^i(v.K+i\widehat{C})^1v_{}^{}{}_{}{}^{j}>_{vv^{}}(\delta _{ij}\widehat{k}_i\widehat{k}_j)\frac{1}{2}$$ (A8) and $$\frac{1}{4\pi }𝐯_t.𝐯_t^{}=\frac{1}{3}[Y_1^1(𝐯)Y_1^1(𝐯^{})+Y_1^1(𝐯)Y_1^1(𝐯^{})]$$ (A9) Similarly, from (30) $$\mathrm{\Pi }_R^l=\mathrm{\Pi }_{00}=m_D^2[k_0\mathrm{\Sigma }_0(k_0,k)1]$$ (A10) If one restricts oneself to transverse gluon exchange, to leading order, the eigenvalues of the collision operator, defined in (A1), have the properties $$\delta _{2l+1}=0,0<\delta _{2l}\delta _2=\frac{5}{8}\gamma \delta _{2l}\frac{2\gamma }{l}\mathrm{as}l\mathrm{}$$ (A11) so that the upper bound $`\overline{\delta }`$ for $`\delta _l`$ is $`\delta _2`$. In the following, the discussion is restricted to $`\mathrm{\Pi }^t`$. The extension to any $`\mathrm{\Sigma }_m`$ is immediate. ### 2 The singularity-free region in the complex $`k_0`$ plane i)The Legendre Functions If one sets $`\gamma 0`$ and $`\delta _l=0`$ , the continued fraction (A5) is a function of $`\rho ^2=(k/(k_0+i\gamma ))^2`$, its value is well known from the case $`\gamma =0`$ $$\mathrm{\Pi }_R^t=\frac{m_D^2}{3}\frac{k_0}{k}[Q_0(\frac{k_0+i\gamma }{k})Q_2(\frac{k_0+i\gamma }{k})]$$ (A12) The analytic properties of $`Q_l(z)`$ in the complex $`z`$ plane are seen from $$Q_l(z)=\frac{1}{2}_1^1𝑑x\frac{P_l(x)}{zx}$$ $`Q_l(z)`$ has a logarithmic singularity at $`z+1`$ and $`z=1`$, and for $`z>1`$, $`Q_l(z)`$ has a series expansion $$Q_l(z)=\underset{nl}{}a_{ln}z^n\mathrm{with}a_{ln}0$$ For $`z>1`$ this series is absolutely convergent i.e. $`_na_{ln}z^n`$ converges. ii) The case $`\delta _l0`$ The continued fraction (A5) can be expanded in powers of $`k^2`$ and the domain of convergence of this expansion delimited. In this expansion there appear products of $$\frac{k^2}{(k_0+i(\gamma \delta _m))(k_0+i(\gamma \delta _{m+1}))}$$ Since all numerical coefficients of the expansion in $`k^2`$ are positive, the series has as an upperbound the series of the modulus of its terms. Then one just needs a lower bound for $$k_0+i(\gamma \delta _m)^2=(\mathrm{Re}k_0)^2+(\mathrm{Im}k_0+\gamma \delta _m)^2$$ If $`\overline{\delta }`$ is the upper bound of the $`\delta _m(\overline{\delta }<2\gamma /3)`$, the lower bounds are $`\mathrm{Im}k_0+\gamma \overline{\delta }>0`$ $``$ $`k_0+i(\gamma \delta _m)>k_0+i(\gamma \overline{\delta })`$ (A13) $`\mathrm{Im}k_0+\gamma <0`$ $``$ $`k_0+i(\gamma \delta _m)>k_0+i\gamma `$ (A14) $`\gamma <\mathrm{Im}k_0<(\gamma \overline{\delta })`$ $``$ $`k_0+i(\gamma \delta _m)>|\mathrm{Re}k_0|`$ (A15) For the region $`\mathrm{Im}k_0+\gamma \overline{\delta }>0`$, the expansion in $`k^2`$ of the continued fraction is certainly convergent in the domain $`k/(k_0+i(\gamma \overline{\delta }))<1`$ as it is the domain of absolute convergence of the expansion in $`k^2`$ of $`Q_l((k_0+i(\gamma \overline{\delta }))/k)`$, i.e., in the complex $`k_0`$ plane, the expansion is convergent out of a half disk of radius $`k`$ centered at $`k_0=i(\gamma \overline{\delta })`$. Similarly, for the region $`\mathrm{Im}k_0<\gamma `$, it is certainly convergent out of a half disk centered at $`k_0=i\gamma `$. And for $`\gamma <\mathrm{Im}k_0<(\gamma \overline{\delta })`$, it is convergent for $`|\mathrm{Re}k_0|>k`$. To summarize, the expansion of $`\mathrm{\Pi }_R^t`$ in powers of $`k^2`$ is certainly convergent in the complex $`k_0`$ plane out of a domain made of two half disks and a rectangle (see Fig.3). In particular, for $`\mathrm{Im}k_00`$ the expansion is certainly convergent for $$(\mathrm{Re}k_0)^2+(\mathrm{Im}k_0+\gamma \overline{\delta })^2>k^2\mathrm{with}\overline{\delta }=\mathrm{sup}_m\delta _m<2\gamma /3$$ (A16) ### 3 The Imaginary parts of $`\mathrm{\Pi }_R^t`$ $`\mathrm{\Pi }_R^t`$ has two types of imaginary parts along the real $`k_0`$ axis. i) all along the axis, $`\mathrm{\Pi }_R^t`$ gets an imaginary part from the eigenvalues $`c_m`$ of the collision operator, i.e. from the damping of the color excitations arising from collisions at the scale $`(gT)^1`$. ii) out of the domain of convergence of the expansion in $`k^2`$, $`\mathrm{\Pi }_R^t`$ gets another imaginary part similar to the $`i\pi `$ term of $`Q_{2l}(z)`$. For example, for $`Q_0(z)`$ in the complex $`z`$ plane, $`z>1`$ $`Q_0(z)={\displaystyle \frac{1}{2}}\mathrm{ln}({\displaystyle \frac{z+1}{z1}})`$ (A17) $`z<1`$ $`Q_0(z)=i{\displaystyle \frac{\pi }{2}}+{\displaystyle \frac{1}{2}}\mathrm{ln}({\displaystyle \frac{1+z}{1z}})\mathrm{above}\mathrm{the}\mathrm{cut}z=1\mathrm{to}z=1`$ (A18) From the shape of the domain of convergence of the expansion in $`k^2`$ (see Fig. 3), this imaginary part may only appear in the region $`|\mathrm{Re}k_0|<k`$. It may be interpreted as a Landau-type effect for the $`W`$ field, i.e. the propagating fluctuating $`W`$ field absorbs (emits) a soft gluon $`kg^2T\mathrm{ln}1/g`$ from the plasma. From the inequality (A16) one sees that for $`k<(\gamma \overline{\delta })`$, the expansion in powers of $`k^2`$ is convergent all along the $`k_0`$ axis, i.e. the half disk on Fig. 3 dont intersect the real axis. As a consequence, $`\mathrm{\Pi }^t`$ has no Landau-type imaginary part for $`k<(\gamma \overline{\delta })\gamma /3`$ This appendix discusses the properties of the retarded amplitude $`\mathrm{\Pi }_R^t(k_0,k)`$. A general property of a retarded propagator is that it is analytic in the upper $`k_0`$ plane. So it is for $`\mathrm{\Pi }_R^t`$ (see Fig.3). The advanced amplitude $`\mathrm{\Pi }_A^t(k_0,k)`$ is just the mirror picture, it is analytic in the lower $`k_0`$ plane, its singularities are in the upper plane $$\mathrm{\Pi }_A^t(k_0,k)=[\mathrm{\Pi }_R^t(k_0,k)]^{}$$ Note that $`\mathrm{\Pi }_R^t(k_0,k)`$ and $`\mathrm{\Pi }_A^t(k_0,k)`$ have no common boundary in the complex $`k_0`$ plane. ($`\mathrm{\Pi }_R^t`$ and $`\mathrm{\Pi }_A^t`$ are on different Riemann sheets of the Landau cut that arises at the scale $`kgT`$). When one considers along the real axis $`\mathrm{\Pi }_R^t(k_0,k)\mathrm{\Pi }_A^t(k_0,k)`$ one is substracting two different functions. ### 4 Locations of the singularities We study the tail of the continued fraction $`\mathrm{\Sigma }_m`$ (see (A5)) . As $`l\mathrm{},b_l^{(m)2}1/4`$ i) case $`\gamma 0,\delta _l=0`$ The tail of the continued fraction $`\mathrm{\Sigma }_m`$ obeys $$X=1\frac{\rho ^2}{4X}\mathrm{i}.\mathrm{e}.X=\frac{1\pm \sqrt{1\rho ^2}}{2}$$ Hence one recovers the fact that the continued fraction has a singularity at $`\rho ^2=1=(k/(k_0+i\gamma ))^2`$ ii) case $`\gamma 0,\delta _l0`$ The same argument gives $$\rho _l^2=1=\frac{k^2}{[k_0+i(\gamma \delta _l)][k_0+i(\gamma \delta _{l+1})]}$$ (A19) i.e. $$k_0=i(\gamma \frac{\delta _l+\delta _{l+1}}{2})\pm \sqrt{k^2(\frac{\delta _l\delta _{l+1}}{2})^2}$$ (A20) Since $`(\delta _l+\delta _{l+1})/2<\overline{\delta }`$ and $`(\delta _l\delta _{l+1})^2<\overline{\delta }^2`$ ($`\overline{\delta }`$ is the upper bound of $`\delta _l`$), for $`k>\overline{\delta }/2\gamma /3`$ all the singularities are inside the two regions in the complex $`k_0`$ plane drawn on Fig.3 $$\sqrt{k^2\overline{\delta }^2/4}<|\mathrm{Re}k_0|k,\overline{\delta }>\mathrm{Im}k_0+\gamma >0$$ Very likely, a cut links these two regions. As $`l\mathrm{}`$ the singularities tend towards $`k_0=\pm ki\gamma `$ since $`\delta _{2l}\gamma /(2l)`$ for $`l\mathrm{}`$ ## B The 4-gluon vertex The 4-gluon vertex $`V_{\mu \nu \rho \sigma }^{1234}(P_{1R},P_{2R},P_{3A},P_{4A})`$ is given by Eq. (118) where $``$ is the sum of two terms $`={\displaystyle \frac{1}{2}}N(P_3,P_4)(f^{14m}f^{23m}+f^{13m}f^{24m})`$ (B7) $`<[v_\mu (v.P_1+i\widehat{C})^1v_\nu +v_\nu (v.P_2+i\widehat{C})^1v_\mu ](v.(P_1+P_2)+i\widehat{C})^1`$ $`[v_\rho (v.P_4+i\widehat{C})^1v_\sigma (p_4^0)+v_\sigma (v.P_3+i\widehat{C})^1v_\rho (p_3^0)]>_{allv}`$ $`+<[v_\mu (v.P_1+i\widehat{C})^1v_\nu +v_\nu (v.P_2+i\widehat{C})^1v_\mu ](v.(P_1+P_2)+i\widehat{C})^1`$ $`\{f^{13m}f^{24m}(p_1^0+p_3^0)[N(P_3,P_2+P_4)v_\rho (v.P_4+i\widehat{C})^1v_\sigma `$ $`N(P_4,P_1+P_3)v_\sigma (v.P_3+i\widehat{C})^1v_\rho >]`$ $`+f^{14m}f^{23m}[34]\}`$ where $`[34]`$ means a term obtained from the factor multiplying $`f^{13m}f^{24m}`$ by the exchange of all indices of $`3`$ and $`4`$. The first term in (B7) is the analogue of the first term in (118) with a colour factor symmetric in $`1`$ and $`2`$ rather than antisymmetric. In both terms of $``$, $`X_1^0,X_2^0`$ are later times, $`X_3^0,X_4^0`$ earlier ones, both are essential for the Ward identities to be satisfied. The symmetries $`12`$ and $`34`$ are explicit in (118) and in (B7). The Ward identity relating $`p_1^\mu V_{\mu \nu \rho \sigma }^{1234}`$ to 3-point vertices is written in Eq.(111) where a vertex of type RRA is related to the one of type ARR (as given in (93)) as follows $$V_{\nu \rho \sigma }^{2m4}(P_{2R},P_{3R},P_{4A})=[V_{\sigma \nu \rho }^{42m}(P_{4R},P_{2A},P_{3A})]^+$$ (B8) where the $`+`$ operation reverses the string of operators, changes $`i\widehat{C}`$ into $`i\widehat{C}`$ and reverses the colour order. This is consistent with Eqs.(51,52) since $`if^{42m}`$ is unchanged in the $`+`$ operation. The Ward identity for a leg of type A is $`ip_4^\sigma V_{\mu \nu \rho \sigma }^{1234}(P_{1R},P_{2R},P_{3A},P_{4A})=f^{43m}V_{\mu \nu \rho }^{12m}(P_{1R},P_{2R},(P_3+P_4)_A)`$ (B12) $`+f^{42m}[{\displaystyle \frac{N(P_3,P_2+P_4)}{N(P_3,P_4)}}V_{\mu \nu \rho }^{1m3}(P_{1R},(P_2+P_4)_A,P_{3A})`$ $`+{\displaystyle \frac{N(P_4,P_1+P_3)}{N(P_3,P_4)}}V_{\mu \nu \rho }^{1m3}(P_{1R},(P_2+P_4)_R,P_{3A})]`$ $`+f^{41m}[12]`$ where the 3-point vertices are those of Eq.(93) and Eq.(B8). ## C The eigenvalues of the collision operator $`\widehat{C}^{}`$ ### 1 The eigenvalues as an integral $`\widehat{C}^{}(𝐯,𝐯^{})`$ commutes with the rotations in $`𝐯`$ space, its eigenvalues $`c_{}^{}{}_{l}{}^{}`$ depend on the only available quantity $`𝐯.𝐯^{}`$ $$c_l^{}=<\widehat{C}^{}(𝐯,𝐯^{})P_l(𝐯.𝐯^{})>_{v,v^{}}$$ (C1) where $`P_l`$ is the Legendre polynomial. In this appendix an analytic expression of $`c_l^{}`$ is given for the case of the dominant transverse gluon exchange in terms of the matrix elements of the operator $`(v.K+i\widehat{C})^1`$ $$c_l^{}=m_D^2\frac{g^2NT}{2}\frac{d_4k}{(2\pi )^4}|\mathrm{\Delta }^t(K_R)|^2[N_{(3g)}^{(l)}(K)+N_{(4g)}^{(l)}(K)]$$ (C2) from Eqs (39,42,45) and (83,86), with $`N_{(3g)}^{(l)}(K)=<P_l(𝐯_1.𝐯_2)<v_{1i}^t(v_1.K+i\widehat{C})^1v_{}^{}{}_{1j}{}^{t}v_{1j}^t(v_1.Ki\widehat{C})^1v_{}^{}{}_{1i}{}^{t}>_{v_1^{}}`$ (C4) $`<v_{2i}^t(v_2.K+i\widehat{C})^1v_{}^{}{}_{2j}{}^{t}v_{2j}^t(v_2.Ki\widehat{C})^1v_{}^{}{}_{2i}{}^{t}>_{v_2^{}}>_{v_1,v_2}`$ $$N_{(4g)}^{(l)}(K)=(\frac{2i\mathrm{I}m\mathrm{\Pi }^t}{k_0m_D^2})<P_l(𝐯.𝐯^{})v_i^t[(v.K+i\widehat{C})^1(v.Ki\widehat{C})^1]v_{}^{}{}_{i}{}^{t}>_{v,v^{}}$$ (C5) where $`v_i^t=v_i\widehat{k}_i𝐯.\widehat{𝐤}`$, and the integration range on $`k`$ is limited to $`kg^2NT\mathrm{ln}1/g`$, i.e. $`\mu _3<k<\mu _2`$ where $`\mu _2`$ and $`\mu _3`$ have been defined in Eq.(50). In this range $$\mathrm{\Pi }^t(k_0,k)=\frac{k_0m_D^2}{3}\mathrm{\Sigma }_1(k_0,k)$$ (C6) where $`\mathrm{\Sigma }_1(k_0,k)`$ is a matrix element of $`(v.K+i\widehat{C})^1`$ given in Eq.(A5) whose scale is $`\gamma g^2NT\mathrm{ln}1/g`$ for $`k_0`$ and $`k`$. Then $$|\mathrm{\Delta }^t(K_R)|^2=\frac{1}{(k_0^2k^2+\mathrm{R}e\mathrm{\Pi }^t)^2+(\mathrm{I}m\mathrm{\Pi }^t)^2}\frac{1}{k^4+k_0^2m_D^4(\mathrm{I}m\mathrm{\Sigma }_1/3)^2}$$ (C7) In the space $`(k_0,k)`$, $`|\mathrm{\Delta }^t|^2`$ put a strong weight upon the region $$k_0\frac{m_D^2}{3}\mathrm{I}m\mathrm{\Sigma }_1(k_0,k)k^2$$ (C8) As $`m_D^2\mathrm{\Sigma }_1m_D^2/\gamma T/\mathrm{ln}1/g`$ the weight is on the domain $$k_0\frac{k^2}{T}\mathrm{ln}1/g$$ (C9) The matrix elements of $`(v.K+i\widehat{C})^1`$ all have the same scale $`\gamma `$ for both $`k_0`$ and $`k`$ (see Eq.(A5) and see Sec.C 4 of this Appendix) except for a few that depend on the zero eigenvalue of $`\widehat{C}`$. For those, one may substitute $`N^{(l)}(k_0=0,k)`$ to $`N^{(l)}(k_0,k)`$ in the integrant of (C2) and perform the integration over $`k_0`$ (see Eq. (60)) with the result $$c_l^{}\frac{g^2NT}{2}_{\mu _3}^{\mu _2}\frac{dk}{4\pi ^2}\frac{3}{|\mathrm{I}m\mathrm{\Sigma }_1(k_0=0,k)|}[N_{(3g)}^{(l)}(k_0=0,k)+N_{(4g)}^{(l)}(k_0=0,k)]$$ (C10) The integral will have no infrared divergence if $`N^{(l)}(k_0=0,k)`$ is finite as $`k0`$ since $`\mathrm{I}m\mathrm{\Sigma }_1(k_0=0,k)1/\gamma `$ as $`k0`$. The exceptional cases will be discussed later on. In the following, $`N_{(3g)}^{(l)}(k_0,k)`$ and $`N_{(4g)}^{(l)}(k_0,k)`$ are expressed in terms of matrix elements of $`(v.K+i\widehat{C})^1`$ whose explicit expressions are given in Sec.C 4. Useful properties are: \- The matrix elements of $`(v.K+i\widehat{C})^1`$ are divided into subspaces where $`l_z=m=𝐥.\widehat{𝐤}`$ is fixed and $`l=m,m+1,m+2,\mathrm{}`$. (see Appendix A) \- Matrix elements for $`l_z=m`$ and $`l_z=m`$ are equal. \- For $`k_0=0`$ and all $`k`$, all the relevant matrix elements are imaginary numbers. One will write $$G=(v.K+i\widehat{C})^1\mathrm{a}ndG^+=(v.Ki\widehat{C})^1$$ (C11) $`\widehat{𝐤}`$ is taken as the $`z`$ axis, and in $`N^{(l)}(k_0,k)`$ one writes $$P_l(𝐯.𝐯^{})=\frac{4\pi }{2l+1}\underset{m=l}{\overset{m=l}{}}Y_l^m(𝐯)Y_l^m(𝐯^{})$$ (C12) ### 2 The 3-gluon vertices diagram’s contribution $`N_{(3g)}^{(l)}(K)`$ is given by Eq.(C4) and (C12) is used. Only the subspace $`|m|=1`$ of $`(v.K+i\widehat{C})^1`$ enters. Indeed, $`v_{}^{}{}_{1i}{}^{t}`$ is an element with $`|m|=1`$, $`(v.K+i\widehat{C})^1`$ acts in this subspace, $`v_{1i}^t`$ and $`Y_l^m(𝐯_1)`$ are combined with the use of Clebsh-Gordan coefficients and the $`𝐯_1`$ integration project them into the subspace. Writing for compactness $$l^{}m=1|GG^+|1m=1=l^{}|\mathrm{\Delta }G|1$$ (C13) the result is $`N_{(3g)}^{(l)}={\displaystyle \frac{1}{3(2l+1)^2}}\{[\left({\displaystyle \frac{l(l1)}{2l1}}\right)^{1/2}l1|\mathrm{\Delta }G|1\left({\displaystyle \frac{(l+1)(l+2)}{2l+3}}\right)^{1/2}l+1|\mathrm{\Delta }G|1]^2`$ (C15) $`+[\left({\displaystyle \frac{(l+1)(l+2)}{2l1}}\right)^{1/2}l1|\mathrm{\Delta }G|1\left({\displaystyle \frac{l(l1)}{2l+3}}\right)^{1/2}l+1|\mathrm{\Delta }G|1]^2(1\delta _{l0})(1\delta _{l1})\}`$ The first term comes from the $`m=0`$ term in the sum (C12), the second from the $`|m|=2`$ term. As the matrix elements have the property (see Eq.(C64)) $$1+n|G^+(k_0,k)|1=(1)^{n1}1+n|G(k_0,k)|1$$ (C17) $`l\pm 1|\mathrm{\Delta }G|1`$ is an even function of $`k_0`$ for $`l`$ even, an odd function of $`k_0`$ for $`l`$ odd. $`N_{(3g)}^{(l)}(k_0=0,k)=0`$ for $`l`$ odd, i.e. with the approximation of (C10) the contribution of $`N_{(3g)}^{(l)}`$ to $`c_l^{}`$ vanishes for $`l`$ odd. The matrix elements $`l\pm 1|G|1`$ may be expressed in terms of $`1|G|1=\mathrm{\Sigma }_1`$ (See Eq.(C53)). Explicitely $$N_{(3g)}^{(l=0)}(k_0,k)=\frac{2}{9}[2\mathrm{I}m\mathrm{\Sigma }_1]^2$$ (C18) $$N_{(3g)}^{(l=1)}(k_0,k)=\frac{1}{5^2}\frac{2}{9}k^2[\frac{\mathrm{\Sigma }_1}{(k_0+ic_2)X_2}\mathrm{c}.c.]^2$$ (C19) $`N_{(3g)}^{(l=2)}(k_0,k)={\displaystyle \frac{1}{5^2}}{\displaystyle \frac{2}{9}}\{[\mathrm{\Sigma }_1(1{\displaystyle \frac{3.4}{5.7}}{\displaystyle \frac{k^2}{(k_0+ic_2)(k_0+ic_3)}}{\displaystyle \frac{1}{X_2X_3}})\mathrm{c}.c.]^2`$ (C21) $`+6[\mathrm{\Sigma }_1(1{\displaystyle \frac{2}{5.7}}{\displaystyle \frac{k^2}{(k_0+ic_2)(k_0+ic_3)}}{\displaystyle \frac{1}{X_2X_3}})\mathrm{c}.c.]^2\}`$ where c.c. means complex conjugate, $`\mathrm{\Sigma }_1(k_0,k)`$ is given in (A5), $`X_2,X_3`$ are defined in term of $`\mathrm{\Sigma }_1`$ in (C49,C50), and $`c_l`$ is the $`l`$ eigenvalue of the operator $`\widehat{C}`$. For $`k<<\gamma `$, $`\mathrm{\Sigma }_1(k_0+ic_1)^1`$ and from Eq.(C53) $`l+1|G|1`$ behaves as $`k^l`$, $`l1|G|1`$ as $`k^{l2}`$. One concludes that the contribution of $`N_{(3g)}^{(l)}`$ to the integral (C10) is infrared finite for all $`l`$. ### 3 The 4-gluon vertex diagram’s contribution $`N_{(4g)}^{(l)}(K)`$ is written Eq.(C5) and (C12) is used. $`v_i^tY_l^m(𝐯)`$ has components in the $`m+1`$ and $`m1`$ subspaces of $`G=(v.K+i\widehat{C})^1`$. The result is written as a sum over the different subspaces $`M`$ of $`G`$, the two first terms correspond to $`M=l+1`$ and $`M=l`$, $`\mathrm{\Delta }G=GG^+`$. $`N_{(4g)}^{(l)}=i{\displaystyle \frac{2}{3}}\mathrm{I}m\mathrm{\Sigma }_1{\displaystyle \frac{2}{2l+1}}\{`$ (C25) $`{\displaystyle \frac{l+1}{2l+3}}l+1l+1|\mathrm{\Delta }G|l+1l+1+{\displaystyle \frac{l}{2l+3}}l+1l|\mathrm{\Delta }G|l+1l`$ $`+{\displaystyle \underset{M=0}{\overset{M=l1}{}}}(1{\displaystyle \frac{1}{2}}\delta _{M0})[{\displaystyle \frac{l(l+1)+M^2}{2l+1}}({\displaystyle \frac{1}{2l+3}}l+1M|\mathrm{\Delta }G|l+1M`$ $`+{\displaystyle \frac{1}{2l1}}l1M|\mathrm{\Delta }G|l1M)b_l^{(M)}b_{l1}^{(M)}2l+1M|\mathrm{\Delta }G|l1M(1\delta _{l0})]\}`$ with $`b_l^{(M)}`$ given in (A4). With the property of G (see Eq.(C64)) $$l^{}+nM|G^+(k_0,k)|l^{}M=(1)^{n+1}l^{}+nM|G(k_0,k)|l^{}M$$ (C26) each matrix element of (C25) is an even function of $`k_0`$. The matrix elements in a given subspace $`M`$ may be written in terms of the lowest one $$\mathrm{\Sigma }_M=l=MM|G|l=MM$$ (C27) For example $$N_{(4g)}^{(l=0)}=\frac{2}{9}[2\mathrm{I}m\mathrm{\Sigma }_1]^2$$ (C28) the contribution of $`N_{(4g)}^{(l=0)}`$ and $`N_{(3g)}^{(l=0)}`$ cancel each other and $`c_0^{}=0`$ as it was shown in Sec.IV D by another method. $`N_{(4g)}^{(l=1)}=i{\displaystyle \frac{2}{3}}\mathrm{I}m\mathrm{\Sigma }_1{\displaystyle \frac{2}{3}}\{{\displaystyle \frac{2}{5}}\mathrm{2\; 2}|\mathrm{\Delta }G|\mathrm{2\; 2}+{\displaystyle \frac{1}{5}}\mathrm{2\; 1}|\mathrm{\Delta }G|\mathrm{2\; 1}`$ (C30) $`+{\displaystyle \frac{1}{3}}[\mathrm{0\; 0}|\mathrm{\Delta }G|\mathrm{0\; 0}+{\displaystyle \frac{1}{5}}\mathrm{2\; 0}|\mathrm{\Delta }G|\mathrm{2\; 0}{\displaystyle \frac{2}{\sqrt{5}}}\mathrm{2\; 0}|\mathrm{\Delta }G|\mathrm{0\; 0}]\}`$ With the use of the relations between the matrix elements of $`G`$, it may be written in terms of $`\mathrm{\Sigma }_2,\mathrm{\Sigma }_1,\mathrm{\Sigma }_0`$ associated respectively with the subspaces $`M=2,1,0`$. For example $$\mathrm{2\; 1}|\mathrm{\Delta }G|\mathrm{2\; 1}=5\frac{d_1}{k^2}(d_1\mathrm{\Sigma }_11)=\frac{d_1}{d_2}\frac{\mathrm{\Sigma }_1}{X_2^{(1)}}$$ (C31) where $`d_l=k_0+ic_l`$ and the $`X_l^{(m)}`$ are expressed in terms of $`\mathrm{\Sigma }_m`$ in (C50,C49). $`N_{(4g)}^{(l=1)}(k_0,k)=i{\displaystyle \frac{2}{3}}\mathrm{I}m\mathrm{\Sigma }_1{\displaystyle \frac{2}{3}}\{{\displaystyle \frac{2}{5}}(\mathrm{\Sigma }_2\mathrm{c}.c.)+{\displaystyle \frac{1}{5}}[{\displaystyle \frac{k_0+ic_1}{k_0+ic_2}}{\displaystyle \frac{1}{X_2^{(1)}}}\mathrm{\Sigma }_1\mathrm{c}.c.]`$ (C33) $`+{\displaystyle \frac{1}{3}}[\mathrm{\Sigma }_0(1{\displaystyle \frac{1}{3}}{\displaystyle \frac{k^2}{(k_0+ic_1)(k_0+ic_2)}}{\displaystyle \frac{1}{X_1^{(0)}X_2^{(0)}}}+{\displaystyle \frac{1}{5}}{\displaystyle \frac{k_0}{k_0+ic_2}}{\displaystyle \frac{1}{X_1^{(0)}X_2^{(0)}}})\mathrm{c}.c.]`$ where the $`c_l`$ are the eigenvalues of $`\widehat{C}`$, and the $`X_l^{(m)}`$ are expressed in terms of $`\mathrm{\Sigma }_m`$ in (C49,C50). The infrared sector For $`k<<\gamma `$, the diagonal matrix elements of $`G`$ do not depend on the subspace $`M`$ $$l^{}M|G|lM=l^{}M|(v.K+i\widehat{C})^1|lM(k_0+ic_l)^1\delta _{ll^{}}$$ (C34) With $`_1^{l1}M^2=l(l1)(l1/2)/3`$, one can check that the full expression (C25) reduces to $$N_{(4g)}^{(l)}(k_0,k<<\gamma )=i\frac{2}{3}\mathrm{I}m\mathrm{\Sigma }_1\frac{2}{2l+1}\frac{1}{3}[\frac{l+1}{k_0+ic_{l+1}}+\frac{l}{k_0+ic_{l1}}\mathrm{c}.c.]$$ (C35) a form immediately obtained from the initial expression of $`N_{(4g)}^{(l)}`$, Eq.(C5), if one writes $$𝐯_t.𝐯_{}^{}{}_{t}{}^{}P_l(𝐯.𝐯^{})=\frac{2}{3}𝐯.𝐯^{}P_l(𝐯.𝐯^{})=\frac{2}{3}\frac{1}{2l+1}[(l+1)P_{l+1}+lP_{l1}]$$ (C36) Because of the eigenvalue $`c_0=0`$, care has to be given to the subspace $`M=0`$. As shown in Sec.C 4 of this appendix, the limits $`k0`$ and $`k_00`$ commute for all $`l^{}0|G|l0`$ with $`l`$ and $`l^{}2`$. One concludes that the contribution of $`N_{(4g)}^{(l)}`$ to the eigenvalue $`c_l^{}`$ is infrared finite for $`l>2`$. For the case $`l=2`$, the elements $`\mathrm{1\; 0}|G|l^{}0`$ enter in (C25), the limits $`k0`$ and $`k_00`$ do not commute, these elements vanish as $`k_00`$, $`k`$ fixed (see Eq.(C59)). There is no infrared divergence in $`N_{(4g)}^{(l=2)}`$. The case of $`c_1^{}`$ As the relation (C35) suggests, one is left with the case $`l=1`$ whose explicit expression is written in (C30) or (C33). In (C30), the matrix elements $`2M|\mathrm{\Delta }G|2M2/ic_2`$ as $`k0`$ irrespective of the order of the limits $`k_0`$ and $`k`$, they give a finite contribution to $`c_1^{}`$. The other terms are $`{\displaystyle \frac{1}{3}}[\mathrm{0\; 0}|\mathrm{\Delta }G|\mathrm{0\; 0}{\displaystyle \frac{2}{\sqrt{5}}}\mathrm{0\; 0}|\mathrm{\Delta }G|\mathrm{2\; 0}]=`$ (C38) $`{\displaystyle \frac{1}{3}}\mathrm{\Sigma }_0(1{\displaystyle \frac{k^2}{(k_0+ic_1)(k_0+ic_2)}}{\displaystyle \frac{1}{X_1^{(0)}X_2^{(0)}}}{\displaystyle \frac{4}{15}})\mathrm{c}.c.`$ if one uses (C53) to relate the second matrix element to the first one. With $$\mathrm{\Sigma }_0(k_0=0,k)=\frac{3ic_1}{k^2}X_1^{(0)}$$ (C39) where $`X_l^{(0)}(k_0=0,k)`$ real $`>1`$ from Eqs.(C49,C50), one sees that the integral (C10) has a linear infrared divergence which is solely due to the existence of the matrix element $`\mathrm{0\; 0}|\mathrm{\Delta }G|\mathrm{0\; 0}=2i\mathrm{I}m\mathrm{\Sigma }_0`$ in (C38). Instead of considering $`\mathrm{\Sigma }(k_0=0,k)`$ as it enters (C10), one may wish to go back one step ahead and study the relevant domain $`(k_0,k)`$ in $`k^2𝑑k𝑑k_0`$ of Eq.(C2). Here $$k^2𝑑k𝑑k_0\frac{1}{k^4+k_0^2a^2}\frac{k^2}{k^4+k_0^2b^2}=\pi \frac{dk}{k^2}\frac{1}{a+b}$$ (C40) From $`|\mathrm{\Delta }^t|^2`$ in (C7) one has $`am_D^2/c_1`$, and for $`\mathrm{I}m\mathrm{\Sigma }_0`$ one has $`bc_1`$ since from (C49), for an estimate $$\mathrm{I}m\mathrm{\Sigma }_0\mathrm{I}m\frac{1}{k_0\frac{k^2}{3(k_0+ic_1)}}=\frac{1}{3}\frac{c_1k^2}{(k_0^2\frac{k^2}{3})^2+c_1^2k_0^2}3\frac{c_1k^2}{k^4+9c_1^2k_0^2}$$ (C41) With $`m_D^2=g^2NT^2/3>>c_1^2(g^2NT\mathrm{ln}1/g)^2`$, one has $`a>>b`$ in (C40) and one concludes that the linear divergence is indeed given by $`\mathrm{\Sigma }_0(k_0=0,k)`$. If one sets $`k_0=0`$ and one writes $`\mathrm{\Sigma }_l(k_0=0,k)=i\stackrel{~}{\mathrm{\Sigma }}_l(k)`$, one gets from (C30) $$N_{(4g)}^{(l)}(k_0=0,k)=\stackrel{~}{\mathrm{\Sigma }}_1\frac{4}{9}[\frac{4}{5}\stackrel{~}{\mathrm{\Sigma }}_2+2\frac{c_1}{k^2}(1c_1\stackrel{~}{\mathrm{\Sigma }}_1)+\frac{2}{3}(\stackrel{~}{\mathrm{\Sigma }}_03\frac{c_1}{k^2})(1+\frac{5}{4})+2\frac{c_1}{k^2}]$$ (C42) As $`k0`$, all the terms are finite, except for the term $`2c_1/k^2`$. In the factor $`(1+5/4)=9/4`$, $`1`$ comes from the infrared singular element $`\mathrm{0\; 0}|\mathrm{\Delta }G|\mathrm{0\; 0}`$ and $`5/4`$ from the elements $`\mathrm{2\; 0}|\mathrm{\Delta }G|\mathrm{2\; 0}`$ and $`\mathrm{0\; 0}|\mathrm{\Delta }G|\mathrm{2\; 0}`$. The form (C42) allows an easy comparison with the result of Arnold and Yaffe at the end of their subsec.7 of Sec.II.C. As it has been discussed in Sec.V B , comparing their expressions at the end of their subsecs.7 and 5, one has to substract out the $`\rho `$-dependent part which does not come fron $`<v_lv_iG_0(p)v_jv_l>𝒫_t^{ij}`$ but from $$\frac{2}{3p^2}(43\frac{\sigma _p}{\sigma _0})=\frac{8}{3p^2}\frac{2\mathrm{\Sigma }_1}{p^2}$$ i.e. for the term in $`(\mathrm{\Sigma }_1)`$, writing in their subsec.7’s form $`8/3=2+2/3`$, it remains $$\frac{1}{2}(\mathrm{\Sigma }_0\frac{3}{\rho ^2})+\frac{2}{3\rho ^2}(1\mathrm{\Sigma }_1)+\frac{4}{15}\mathrm{\Sigma }_2$$ which agrees with the convergent term in the bracket in Eq.(C42) when one factors out $`1/3`$. ### 4 Matrix elements of $`G=(v.K+i\widehat{C})^1`$ The eigenvectors of $`\widehat{C}(𝐯.𝐯^{})`$ are $`|lm=\sqrt{4\pi }Y_l^m(𝐯)`$ with the measure $`d\mathrm{\Omega }_𝐯/4\pi `$, its eigenvalues are $`c_l`$. It is convenient to choose $`\widehat{𝐤}`$ as the z axis, then the operator $`(k_0+i\widehat{C}v_zk)`$ changes $`l`$ but does not change $`l_z=m`$, so does its inverse $$l^{}m|(k_0+i\widehat{C}v_zk)^1|lm=\delta _{mm^{}}l^{}m|(k_0+i\widehat{C}v_zk)^1|lm$$ (C43) In a given subspace $`m`$, all matrix elements of $`G`$ may be written in terms of the lowest one $$l=mm|G|l=mm=\mathrm{\Sigma }_m$$ (C44) because of the recursion relations $$l^{}m|GG^1|lm=\delta _{ll^{}}$$ (C45) $`\delta _{ll^{}}`$ $`=`$ $`kb_l^{}l^{}+1|G|lkb_{l^{}1}l^{}1|G|l+d_l^{}l^{}|G|l`$ (C46) $`\delta _{ll^{}}`$ $`=`$ $`kb_ll^{}|G|l+1kb_{l1}l^{}1|G|l+d_ll^{}|G|l`$ (C47) with the definition $$l|k_0+i\widehat{C}|l=k_0+ic_l=d_l$$ (C48) The $`b_l=b_l^{(m)}`$ are the matrix elements of $`v_z`$ defined in (A4), and the reference to the subspace $`m`$ has been dropped. Note that a consequence of (C47) is, as $`k0`$, $`l^{}|G|l\delta _{ll^{}}/d_l`$ independent of $`m`$. It is convenient to exhibit the threshold factors arising from angular momentum conservation. $`\mathrm{\Sigma }_m`$ is the continued fraction explicited in Eq.(A5), it may be written $$\mathrm{\Sigma }_m(k_o,k)=\frac{1}{d_m}\frac{1}{1{\displaystyle \frac{k^2b_m^2}{d_md_{m+1}}}{\displaystyle \frac{1}{X_{m+1}}}}$$ (C49) $$X_{m+1}=1\frac{k^2b_{m+1}^2}{d_{m+1}d_{m+2}}\frac{1}{X_{m+2}}$$ (C50) with $`d_m=k_0+ic_m`$ (see (C48)), $`b_{m+n}=b_{m+n}^{(m)}`$ is the quantity that depends on the subspace $`m`$ (see (A4)) $$b_{0}^{(0)}{}_{}{}^{2}=\frac{1}{3},b_{1}^{(0)}{}_{}{}^{2}=\frac{4}{15},b_{1}^{(1)}{}_{}{}^{2}=\frac{1}{5}$$ (C51) So do the $`X_m`$. Then the solution to the recursion relations in the subspace $`m`$ may be written for any $`lm`$, $`n>0`$ $`l+n|G|l=l|G|l+n=`$ (C53) $`{\displaystyle \frac{b_lb_{l+1}\mathrm{}b_{l+n1}}{X_{l+1}X_{l+2}\mathrm{}X_{l+n}}}{\displaystyle \frac{k^n}{d_{l+1}d_{l+2}\mathrm{}d_{l+n}}}l|G|l`$ and $`l|G|l`$ is related as follows to the lowest $`l=m`$ $$l=m|G|l=m=\mathrm{\Sigma }_m$$ (C54) $$l=m+1|G|l=m+1=\frac{d_m}{d_{m+1}X_{m+1}}\mathrm{\Sigma }_m$$ (C55) $$l=m+2|G|l=m+2=\frac{d_m}{d_{m+2}}\frac{1}{X_{m+1}X_{m+2}}(1\frac{k^2b_m^2}{d_md_{m+1}})\mathrm{\Sigma }_m$$ (C56) $$l=m+3|G|l=m+3=\frac{d_m}{d_{m+3}}\frac{1}{X_{m+1}X_{m+2}X_{m+3}}(1\frac{k^2b_m^2}{d_md_{m+1}}\frac{k^2b_{m+1}^2}{d_{m+1}d_{m+2}})\mathrm{\Sigma }_m$$ (C57) Limits as $`k_0`$ or $`k0`$ i) For any subspace $`m0`$ \- as $`k0`$, all $`X_m1`$, $`\mathrm{\Sigma }_m1/d_m`$, one finds $`l|G|l1/d_l`$ independent of the subspace $`m`$ as expected, and one obtains the correct threshold factor for the non-diagonal matrix elements. \- as $`k_00`$, $`d_m=ic_m`$, $`X_m>1`$, all $`l|G|l`$ are imaginary, and the non-diagonal matrix elements are alternatively real and imaginary. ii) For the subspace $`m=0`$, $`d_0=k_0`$ as $`c_0=0`$, and care is needed when taking the limits \- $`k_00`$ the limit $`k0`$ is as in the other subspaces $`m`$ \- $`k`$ fixed, $`k_00`$, $`d_00`$, $`d_mic_m`$, $`X_m>1`$. From (C49) $$\mathrm{0\; 0}|G|\mathrm{0\; 0}=\mathrm{\Sigma }_0=\frac{d_1X_1}{k^2b_0^2}=i\frac{c_1X_1}{k^2b_0^2}$$ (C58) For $`l2`$, the elements $`l0|G|l0`$ have factors $$\frac{1}{d_l}(d_0(1\frac{k^2b_0^2}{d_0d_1})+\mathrm{})\mathrm{\Sigma }_0_{k_00}\frac{1}{d_l}(\frac{k^2b_0^2}{d_1}+O(k^4))\mathrm{\Sigma }_0$$ so that the limit $`k0`$ again gives $`1/d_l`$. As a consequence, the limits $`k0`$ and $`k_00`$ commute in all matrix elements $`l0|G|l^{}0`$ with $`l`$ and $`l^{}2`$. The exceptions are seen from Eqs.(C53),(C55),(C58). For $`k_00`$, $`k`$ fixed, $$\mathrm{1\; 0}|G|l00\mathrm{f}orl1$$ (C59) $$\mathrm{0\; 0}|G|\mathrm{1\; 0}=\frac{b_0k\mathrm{\Sigma }_0}{d_1X_1}\frac{1}{kb_0}$$ (C60) and $`\mathrm{0\; 0}|G|l0`$ have the anomalous threshold factor $`k^{l2}`$. Another quantity to be used is $$\frac{k_0}{k^2}(k_0\mathrm{\Sigma }_01)=\frac{1}{3}\frac{k_0}{k_0d_1(1{\displaystyle \frac{k^2b_{1}^{(0)}{}_{}{}^{2}}{d_1d_2X_2^{(0)}}}){\displaystyle \frac{k^2}{3}}}$$ (C61) where $`d_1=k_0+ic_1`$. There are two limits: $`k^2<<k_0|k_0+ic_1|`$, the limit is $`1/3d_1`$ $`k_0|k_0+ic_1|<<k^2`$, the limit is $`k_0/k^2`$ Complex conjugation The matrix elements of $`G=(v.K+i\widehat{C})^1`$ and of $`G^+=(v.Ki\widehat{C})^1`$ may be related. Writing $`d_m^{}(k_0)=k_0ic_m=d_m(k_0)`$, one has from Eqs.(C49) to (C54) $`X_m^{}(k_0,k)`$ $`=`$ $`X_m(k_0,k)`$ (C62) $`\mathrm{\Sigma }_m^{}(k_0,k)`$ $`=`$ $`\mathrm{\Sigma }_m(k_0,k)`$ (C63) $`l|G^+(k_0,k)|l+n`$ $`=`$ $`(1)^{n+1}l|G(k_0,k)|l+n`$ (C64) Limit when $`\widehat{C}`$ is replaced by $`ϵ`$ $`l^{}m|(v.K+iϵ)^1|lm=`$ (C66) $`(1)^m{\displaystyle \underset{L=ll^{}}{\overset{L=l+l^{}}{}}}{\displaystyle \frac{1}{k}}Q_L({\displaystyle \frac{k_0+iϵ}{k}})[(2l+1)(2l^{}+1)]^{1/2}ll^{};\mathrm{0\; 0}|L0ll^{};mm|L0`$
warning/0004/cond-mat0004356.html
ar5iv
text
# Cooperative motion in Lennard-Jones binary mixtures below the glass transition \[ ## Abstract Using the activation-relaxation technique (ART), we study the nature of relaxation events in a binary Lennard-Jones system above and below the glass transition temperature ($`T_g`$). ART generates trajectories with almost identical efficiency at all temperature, thus avoiding the exponential slowing down below $`T_g`$ and providing extensive sampling everywhere. Comparing these runs, we find that the number of atoms involved in an event decreases strongly with temperature. In particular, while in the supercooled liquid activated events are collective, involving on average thirty atoms or more, events below T<sub>g</sub> involve mostly single atoms and produce minimal disturbance of the local environment. These results confirm the interpretation and the generality of recent NMR results by Tang et al (Nature 402, 160 (1999)). \] Atomic motion in solids is largely determined by the nature of the local network. While atomic diffusion in crystals is constrained by symmetry and can be described in terms of well-defined displacements leaving the overall structure of the network unaffected, diffusion in disordered materials offers a much more complex picture. These materials present a wide range of local environments and diffusion can take place in principle through an equally large number of mechanisms. This has made understanding the nature of diffusion and relaxation mechanisms in compact glasses, such as metallic and Lennard-Jones glasses, a difficult task. Nevertheless, significant progress regarding the details of the dynamics in these dense glasses has been achieved recently. On the theoretical side, simulations have been used extensively to investigate this problem. Studies on supercooled model binary glasses have established clearly that as a liquid becomes supercooled a change in the dynamics takes place and diffusion starts to proceed by jumps. More recent work has provided further characterization of these mechanisms, showing that in the supercooled regime moves are collective and occur in highly correlated sequences. Experimental measurements are also challenging; diffusion takes place on a long time scale and the data represent only an average over a distribution of barriers and pre-factors. This make it difficult to identify directly specific mechanisms. A search for diffusion mechanisms in multinary metallic glasses, using standard slicing techniques, has led to conflicting results. More direct probes to identify local changes around atoms, such as NMR, have also been used recently. In particular, the work of Tang et al. implies that there is a qualitative change in the diffusion mechanism of Be atoms as the Zr-Ti-Cu-Ni-Be samples are brought below the glass transition temperature ($`T_g`$): from mostly collective, jumps become localized and clearly atomistic. This last result is of great interest because it demonstrates a qualitative distinction between the supercooled regime and the dynamics below $`T_g`$. A numerical reproduction of this phenomenon, necessary to establish the validity of the explanation and its generality, is difficult to achieve using standard techniques: in the low temperature regime, the time scale covered by molecular dynamics is insufficient to ensure a satisfactory exploration of the space of configurations. The activation-relaxation technique (ART) offers a way to go beyond these limitations and to sample the phase space of disordered systems even at low temperatures. We show here that ART can generate trajectories in Lennard-Jones glasses without suffering from exponential slowing down below $`T_g`$. Comparing events above and below $`T_g`$, we also find that the number of atoms involved in relaxation and diffusion decreases significantly with temperature, going from many tens to one or two at the lowest $`T`$ studied here. ART by-passes the description of the phonon vibrations to concentrate on activated mechanisms: it looks directly for paths connecting minima in a high-dimensional energy landscape. Starting from a local minimum, the whole configuration is first pushed away from it, until a negative eigenvalue appear, and then directed to a nearby saddle point – the activation. The configuration is then brought to a new minimum, providing a complete event with initial, saddle and final configurations. The new move is then accepted or rejected with Bolztmann probability ($`\mathrm{exp}(\mathrm{\Delta }E/k_BT`$, where $`\mathrm{\Delta }E`$ is the energy difference between final and initial configuration and $`T`$ the simulation temperature.) A more detailed description of the original algorithm can be found in Ref. . In order to ensure a better control of the trajectory in this dense material, we use here a modified version of the algorithm. In the activation stage, the configuration is now pushed against the force corresponding to the to the lowest (negative) eigenvalue of the hessian matrix, the second derivative of the total configurational energy. Since an exact diagonalization of the $`3N\times 3N`$ matrix is computationally too demanding for the 1000-atom simulation presented here, a Lanczos algorithm is used to project out eigenvectors corresponding to the lowest eigenvalues only. In spite of the efficiency of the Lanczos algorithm, this approach remains numerically more intensive than the standard ART. However, ART nouveau ensures a direct convergence to the saddle point and provides a better control on the trajectory, which is particularly useful in dense systems such as metallic or Lennard-Jones glasses. This algorithm is applied to a 1000-atom Lennard-Jones binary $`A_{80}B_{20}`$ mixture using the parameters introduced in Ref. but with shifted energy and forces to ensure continuity of the energy and the first derivative at the cut-off. Energy has units of $`ϵ_{AA}`$, temperature of $`ϵ_{AA}/k_B`$, length of $`\sigma _{AA}`$ and time of $`(m\sigma _{AA}/48ϵ_{AA})`$. The simulation procedure is as follows. We start with a randomly-packed configuration and first relax it at constant volume and $`T=0.50`$, a temperature slightly above the glass transition, until thermalization, i.e., until the configurational energy reaches a plateau. This takes place in about 5000 ART-events. The last configuration of the initialization run is then used as a starting point for further runs at $`T=`$ 0.25, 0.50 and 1.00, i.e., well below, slightly above and well above $`T_g`$, respectively. $`T_g`$, here, is defined as in Ref. , by a sharp low-temperature break in the inherent-structure energy curve. The overall acceptation ratio for these runs is about 20 %. This includes exchange events, accounting for about 6 % at all temperatures, where two or more atoms switch position, leaving the final configuration structurally indistinguishable from the original. Although physically relevant in the study of self-diffusion, these events do not contribute to the relaxation of the lattice per se and they excluded from the analysis below. Accepted events are therefore only those resulting in a final configuration structurally different for the initial one. The acceptation rate for these events is about 20 % at $`T=1.00`$, 10 % at $`T=0.50`$ and 4-5 % at $`T=0.25`$. Figure 1 shows the energy sequence of the accepted events at the three temperatures considered here. As the model is initially prepared at T=0.50, this sequence is already thermalized. The thermalization at $`T=1.00`$ takes about 300 events while that at $`T=0.25`$ is longer, about 500 events. It is not formally possible to talk of equilibrium below the glass transition, however all quantities described below have been tested over different intervals at $`T=0.25`$ and found to be insensitive to the specific subset of events chosen past event 400. The exact value of the configurational energies after thermalization with ART is significantly lower than that of models relaxed with molecular dynamics. Starting with a configuration equilibrated at $`T=1.5`$ and slowly cooling down, (at $`10^5ϵ_{AA}/`$ time unit), the inherent structures stabilize at an energy per atom of about $`6.738`$ at $`T=1.00`$, $`6.817`$ at $`T=0.50`$ and $`6.858`$ at $`T=0.25`$. The corresponding values for the ART run are $`6.810`$, $`6.844`$ and $`6.870`$, respectively. ART is not expected to fully describe the liquid phase since it is event-based and does not include entropic contributions. This means that distributions, such as those presented below, obtained with ART should be sharper than those generated with MD in the liquid and supercooled liquid. ART and MD should meet in the solid phase where the dynamics is almost exclusively activated. This is the case in other systems and also here, with only a small energy difference between both techniques at $`T=0.25`$. This difference can be explained, in large part, by the much better sampling of the energy landscape achieved by ART. At all temperatures, the sampling of the energy surface is significant. Figure 2 shows the root-mean-square displacement per atom as a function of accepted events. Diffusion is linear in this quantity, following an Einstein-like relation with a “diffusion constant” almost independent of the temperature. In terms of accepted events, i.e., without even including atomic exchanges, the sampling of the phase space proceeds therefore at a constant rate. ART clearly overcomes the exponential slowing down of the dynamics in these metallic glasses. With the properties of the sampling established, we can study the the nature of the relaxation and diffusion mechanisms as a function of temperature. The self part of the van Hove correlation function, $$4\pi r^2G_s(r,n)=\frac{4\pi r^2}{N}\underset{i}{}\delta \left(r|𝐫_i(0)𝐫_i(n)|\right)$$ (1) provides the probability of finding an atom $`r`$ away from it initial position at event $`n`$. In Figure 3(a), the correlation is plotted as a function of event number $`n`$ at two different temperatures. The distributions at all temperatures are significantly sharper than those generally obtained in molecular dynamics . Because ART concentrates exclusively on activated events, the thermal contribution to the peak around $`r=0`$ are eliminated. At $`T=1.00`$ (inset), the distribution is initially bimodal with broad peaks and evolves into a single wide peak, indicating relaxation mechanisms with a range of of lengths. The distribution at $`T=0.25`$, on the other hand, is narrow and evolves by adding new peaks at integer multiples of $`r=1.05\sigma `$. This behavior is underlined in Fig. 3(b), which compares the long time distribution at all temperatures: as the temperature lowers, the atomic displacements become more and more discrete: the jumps take place on a disordered but rigid lattice. The $`T=0.50`$ model, well into the supercooled region, is not yet as discrete: the distribution is broad and smooth beyond the first peak. In order to get a better understanding of the events, it is useful to plot the average number of atoms involved in an event. Figure 4 gives the full distribution as a function of a threshold displacement. As discussed elsewhere, the slope of the curve is universal. There is a strong temperature dependence on the pre-factor, however, which relates to the rigidity of the network sampled locally: the events become more and more collective as the temperature increases. At $`r=0.1\sigma `$, events involve more than 2 times more atoms at $`T=0.50`$ and eight times more atoms at $`T=1.00`$ than at $`T=0.25`$. over, there are essentially no atoms moving by $`0.2<r/\sigma <1.0`$ in the $`T=0.25`$-run, as indicated by a flat curve: atoms moving have to jump by a nearest-neighbor distance, anything less is not possible. A similar phenomenon can be seen in the $`T=0.50`$ curve, but for $`0.6<r/\sigma <1.0`$, and a slope still relatively high: as an atom jumps by a nearest-neighbor distance, atoms in the local environment relax by up to $`rsigma=0.6`$. Although the degree of collectivity in the events varies strongly with temperature, the three curve come together at $`r=1.0\sigma `$. At all temperatures, the activated dynamics is therefore controlled by nearest-neighbor jumps. This characteristic can also be seen in the self-part of the van Hove correlation (Fig 3): for small intervals, $`t=20`$, the $`T=1.00`$ distribution shows a strong peak at $`r=1.05\sigma `$, identical to that at $`T=0.25`$. How do these results fit with those on heterogeneities in glasses obtained by MD? The conclusions of these simulations can be summarized as follows: heterogeneities exist in glasses but they are not associated with a given scale; their spatial extent depends directly on the time scale selected, leading to homogeneities in the long run. Similar results are found here. Comparing the first and the last half of the run at $`T=0.25`$, for example, we find that 226 atoms move by more than $`r=0.2`$ in the first half of the last 400 events, and 212 in the second, with 75 atoms belonging to the two sets, in agreement with MD work. Where ART and MD differ, however, is in the definition of “collective” event. Certain groups have concluded that the dynamics becomes more and more collective as the temperature decreases, in apparent contraction with our results. However, the adjective collective is applied there for events taking place over an extended time period and pertains more to time correlation between events than to the nature of each activated jump. As such, ART and MD conclusions are not contradictory. Further analysis is currently underway to see whether ART produces the same long time correlations. In summary, we have performed a simulation of a binary Lennard-Jones glass above and below the glass transition temperature. We first show that ART can explore the phase space below $`T_g`$ without encountering exponential slowing down. We also find that there is a qualitative change in the nature of the mechanisms as one goes from above to below T<sub>g</sub>. Although all events seem to be centered around a few atoms jumping over a lattice spacing, the total number of atoms involved in an event decreases by an order of magnitude as the temperature goes down. These results support the interpretation of recent NMR experiments. Although these results are likely to be valid for metallic glasses in general, we can expect different behavior from less dense glasses such as network glasses. Acknowledgments. I would like to acknowledge numerous discussions with G. T. Barkema, S. de Leeuw, Y. Limoge, Y. Wu, as well as D. Drabold. This work is supported in part by the NSF, under grant DMR-9805848. Part of these simulations where run on the computers of HP$`\alpha `$C at TU Delft.
warning/0004/hep-ph0004140.html
ar5iv
text
# Padé-Improved Estimates of Hadronic Higgs Decay Rates ## 1. Introduction The Higgs boson characterizes the electroweak symmetry breaking underlying the Standard Model. Its discovery and phenomenology will be of immense importance in clarifying our understanding of this symmetry breaking, as well as in providing vital information as to the nature of beyond-the-Standard-Model physics. The two leading hadronic decay modes of a Weinberg-Salam Higgs boson $`(H)`$ with mass between $`100\mathrm{GeV}`$ and $`160\mathrm{GeV}`$ are the QCD processes $`Hb\overline{b}`$ and $`H`$ two gluons $`(gg)`$. Although the rate for this latter process is known to $`𝒪(\alpha _s^4)`$, such precision incorporates only two non-leading orders of a slowly converging series in the strong coupling. If $`M_H=100\mathrm{GeV}`$, for example, the known order-by-order QCD corrections have been calculated to be $`1+0.66+0.21`$. The work presented here is primarily directed toward obtaining a more precise estimate of the $`Hgg`$ decay rate. We utilize renormalization-group $`(RG)`$ and asymptotic Padé-approximant methods to estimate the full next-order contribution to the underlying correlation function for this process. Such an approach has already been applied to the quark-antiquark scalar-current correlation function underlying the $`Hb\overline{b}`$ rate , a calculation we review in Section 5 of the present paper. As in our prior analysis of the two-gluon decay amplitude of a non-Standard-Model CP-odd Higgs field , the approach we take here is to test asymptotic Padé-approximant estimates against RG-accessible coefficients within the next-order of perturbation theory. We show such estimates to be accurate up to relative errors of only a few percent, supporting the credibility of the same approach in estimating the RG-inaccessible coefficient needed to determine the full $`𝒪(\alpha _s^5)`$ contribution to the $`Hgg`$ rate. We operate within the context of $`\overline{MS}`$ expressions for the $`Hgg`$ and $`Hb\overline{b}`$ rates that explicitly depend on an arbitrary renormalization scale $`\mu `$. It has been argued elsewhere that asymptotic Padé-approximant methods reduce the explicit scale-dependence of perturbative quantities which must ultimately be scale invariant. We find this to be the case for the $`Hgg`$ rate as well, despite residual scale-sensitivity anticipated from the estimated RG-inaccessible coefficient of the nonlogarithmic $`𝒪(\alpha _s^5)`$ term within the next-order correlator. In Section 2, we demonstrate the explicit RG-invariance of the $`Hgg`$ rate, as calculated in in the $`m_b0,M_H^2<<4M_t^2`$ limit. This RG invariance enables one to calculate all the next-order coefficients $`c_k`$ of $`ln^k[\mu ^2/m_t(\mu ^2)]`$ $`(\alpha _s(\mu )/\pi )^5`$ within the calculated rate; only the $`k=0`$ constant term is RG-inaccessible. In Section 3, we review how asymptotic Padé-approximant methods may be utilized to estimate this set of next-order coefficients, and demonstrate close agreement with the RG-determinations of Section 2 over the range $`100M_H175\mathrm{GeV}`$. The RG-inaccessible coefficient $`c_0`$ is also estimated over this range of Higgs masses and is fitted to its anticipated behaviour as a degree-3 polynomial in $`ln[(M_H^2/M_t^2)_{pole}]`$. In Section 4 we examine the residual scale dependence of the Padé-improved $`Hgg`$ decay rate, as obtained in Section 3. Although we anticipate a relative scale dependence comparable to $`c_0(\alpha _s(\mu )/\pi )^3(310\%)`$, since $`c_0`$ cannot be extracted perturbatively from the RG equation (to the order we consider), we find that the residual scale dependence of the decay rate is a full order of magnitude smaller than this estimate for $`\mu `$ between 0.3 $`M_H`$ and $`M_t`$ (we assume $`100\mathrm{GeV}M_H175\mathrm{GeV}`$). Consequently, the scale-dependence of the Padé-derived term $`c_0(\alpha _s(\mu )/\pi )^5`$ very nearly cancels the scale-dependence anticipated from truncation of the perturbative series for the $`Hgg`$ rate, thereby facilitating the (nearly) scale-invariant rate predictions of Table 6. Predictions of the $`Hgg`$ decay rate are tabulated both for the $`m_b0,M_H^2<<4M_t^2`$ limit, and for the leading-order departures from this limit for appropriate constant and running values for $`m_b`$ and $`M_t`$. In the concluding section, we review how prior asymptotic Padé-approximant estimates of the $`𝒪(\alpha _s^4)`$ contribution to the scalar-current correlation function can be incorporated into the $`Hb\overline{b}`$ decay rate. We demonstrate that the resulting next-order contribution, though only 0.01% of the leading term, is somewhat larger than known $`𝒪(\alpha _s^2)`$ power-suppressed contributions . Consequently, the estimated $`𝒪(\alpha _s^4)`$ term enables the $`Hb\overline{b}`$ rate to be estimated to four significant figure accuracy. ## 2. RG-Invariance of the $`Hgg`$ Rate The Higgs $``$ gg decay rate is given explicitly to 3-loop order in by the following expressions: $$\mathrm{\Gamma }_{Hgg}=\frac{\sqrt{2}G_F}{M_H}R(\alpha _s,q^2=M_H^2,\mu ^2,M_t^2),$$ (2.1a) $$R=C_1^2Im<[0_1^{}]^2>,$$ (2.1b) $`C_1={\displaystyle \frac{x^{(6)}}{12}}[1`$ $`+`$ $`x^{(6)}\left({\displaystyle \frac{11}{4}}{\displaystyle \frac{1}{6}}ln\left({\displaystyle \frac{\mu ^2}{M_t^2}}\right)\right).`$ . $`+`$ $`(x^{(6)})^2({\displaystyle \frac{211}{36}}+{\displaystyle \frac{55}{48}}ln\left({\displaystyle \frac{\mu ^2}{M_t^2}}\right)+{\displaystyle \frac{1}{36}}ln^2\left({\displaystyle \frac{\mu ^2}{M_t^2}}\right))+𝒪\left[(x^{(6)})^3\right]]`$ (2.1c) $`Im<[0_1^{}]^2>={\displaystyle \frac{2q^4}{\pi }}[1`$ $`+`$ $`x^{(5)}\left({\displaystyle \frac{149}{12}}+{\displaystyle \frac{23}{6}}ln\left({\displaystyle \frac{\mu ^2}{q^2}}\right)\right).`$ (2.1d) $`+`$ $`(x^{(5)})^2(68.64817+{\displaystyle \frac{1297}{16}}ln\left({\displaystyle \frac{\mu ^2}{q^2}}\right)+{\displaystyle \frac{529}{48}}ln^2\left({\displaystyle \frac{\mu ^2}{q^2}}\right))+𝒪\left[(x^{(5)})^3\right]].`$ In the above expression, $`<[0_1^{}]^2>`$ is the vacuum polarization of the Higgs field induced via the gluon operator $`0_1^{}=G_{a\mu \nu }G_a^{\mu \nu }`$ , and $`x^{(n_f)}\alpha _s^{(n_f)}/\pi `$, where $`\alpha _s^{(n_f)}(\mu )`$ is the running strong coupling with $`n_f`$ active flavours. Five flavours are assumed to be light in both (2.1c) and (2.1d). The $`t`$-quark mass $`M_t`$ appearing in (2.1c) is an RG-invariant pole mass, and $`M_H^2`$ is assumed to be small compared to $`4M_t^2`$ . Our goal here is to use asymptotic Padé-approximant methods in conjunction with the RG-invariance of (2.1a) in order to predict the next order contribution to $`\mathrm{\Gamma }(Hgg)`$. In a previous application of such methods to the inclusive semileptonic $`bu`$ rate, it was found that success in predicting RG-accessible next-order coefficients was greatly enhanced by recasting the entire expression in terms of the running fermion mass. Indeed, such replacement of the $`b`$-quark pole mass with its scale-dependent $`\overline{MS}`$ mass had already been employed by van Ritbergen to avoid a renormalon pole. Consequently, we first re-express the $`H2g`$ rate in terms of the running $`t`$-quark mass $`m_t(\mu )`$, which evolves via a six-active-flavour $`\gamma _m`$-function, and the corresponding six-active-flavour running coupling $`x^{(6)}(\mu )`$. This transformation is facilitated by the following relationships : $$x^{(5)}(\mu )=x^{(6)}(\mu )\frac{(x^{(6)}(\mu ))^2}{6}ln\left(\frac{\mu ^2}{m_t^2(\mu )}\right)+𝒪\left[(x^{(6)})^3\right],$$ (2.2) $$m_t(\mu )/M_t=[1x^{(6)}(\mu )(\frac{4}{3}+ln\left(\frac{\mu ^2}{M_t^2}\right))+𝒪\left[(x^{(6)})^2\right].$$ (2.3) Equation (2.3) implies that the logarithm in (2.1c) may be re-expressed in terms of $`Lln(\mu ^2/m_t^2(\mu ))`$: $$ln\left(\frac{\mu ^2}{M_t^2}\right)=Lx^{(6)}(\mu )\left[2L+8/3\right]+𝒪\left[(x^{(6)})^2\right]$$ (2.4) With $`q^2=M_H^2`$, the logarithm in (2.1d) can also be expressed entirely in terms of the running $`t`$-quark mass and a logarithm $`Tln\left(M_H^2/M_t^2\right)`$ of the ratio of $`RG`$-invariant pole masses: $$ln\left(\frac{\mu ^2}{M_H^2}\right)=LTx^{(6)}(\mu )\left[2L+8/3\right]+𝒪\left[(x^{(6)})^2\right].$$ (2.5) Substitution of (2.2) and (2.5) into (2.1d) with $`q^2=M_H^2`$, and substitution of (2.4) into (2.1c) leads to the following expression for the $`Hgg`$ decay rate: $$\mathrm{\Gamma }_{Hgg}=\frac{\sqrt{2}G_FM_H^3}{72\pi }S[x^{(6)}(\mu ),L(\mu ),T],$$ (2.6) $`S[x,L,T]=x^2(1`$ $`+`$ $`x\left[\left({\displaystyle \frac{215}{12}}{\displaystyle \frac{23T}{6}}\right)+{\displaystyle \frac{7}{2}}L\right].`$ $`+`$ $`x^2\left[\left(146.8912{\displaystyle \frac{4903}{48}}T+{\displaystyle \frac{529}{48}}T^2\right)+\left({\displaystyle \frac{1445}{16}}{\displaystyle \frac{161}{8}}T\right)L+{\displaystyle \frac{147}{16}}L^2\right]`$ . $`+`$ $`x^3[c_0+c_1L+c_2L^2+c_3L^3]+𝒪(x^4)).`$ (2.7) In (2.7), we list the unknown coefficients $`c_0,c_1,c_2`$ and $`c_3`$ of the 4-loop contribution to the rate. Three of these may be extracted by the scale-\[RG-\] invariance of the physical decay rate: $`d\mathrm{\Gamma }/d\mu =0`$. This invariance implies that $`O=\mu {\displaystyle \frac{dS}{d\mu }}[x,L,T],`$ $`=[12\gamma _m(x)]{\displaystyle \frac{S}{L}}+\beta (x){\displaystyle \frac{S}{x}}.`$ (2.8) Both the $`\beta `$ and $`\gamma _m`$ functions in (2.8) are referenced to six active flavours: $$\beta ^{(6)}(x)=\frac{7}{4}x^2\frac{13}{8}x^3+\frac{65}{128}x^4\mathrm{}$$ (2.9) $$\gamma _m^{(6)}(x)=x\frac{27}{8}x^2\mathrm{}.$$ (2.10) One can easily verify from the known terms listed in (2.7) that (2.8) is perturbatively valid to orders $`x^3,x^4`$ (including explicit cancellation of terms involving $`T`$), and $`x^4L`$. The continued perturbative validity of (2.8) to orders $`x^5L^2`$, $`x^5L`$, and $`x^5`$ is sufficient to determine the four-loop coefficients $`c_1,c_2`$, and $`c_3`$: $$c_1=910.3167\frac{16643}{24}T+\frac{3703}{48}T^2,$$ (2.11) $$c_2=\frac{1225}{4}\frac{1127}{16}T,$$ (2.12) $$c_3=\frac{343}{16}.$$ (2.13) The coefficient $`c_0`$ is not RG-accessible to these orders. In the next section, we will utilize asymptotic Padé approximant methods to estimate the four-loop coefficients $`\{c_0,c_1,c_2,c_3\}`$. As in prior work , the accuracy of these predictions in reproducing (2.11 - 2.13) will serve as an indication of the accuracy of our estimate for $`c_0`$. ## 3. Padé-Predictions for the Four-Loop <br>Coefficients The asymptotic Padé-approximant procedure for estimation of the four loop coefficients $`c_0,c_1,c_2,c_3`$ in (2.7) has been delineated in previous work . The series (2.7) may be expressed in the form $$S[x,L,T]=x^2\left[1+R_1[L,T]x+R_2[L,T]x^2+R_3[L,T]x^3+\mathrm{}\right]$$ (3.1) where $`R_1`$ and $`R_2`$ are known, and $`R_3`$ is to be determined: $$R_1[L,T]=\left(\frac{215}{12}\frac{23}{6}T\right)+\frac{7}{2}L$$ (3.2) $$R_2[L,T]=146.8912\frac{4903}{48}T+\frac{529}{48}T^2+\left(\frac{1445}{16}\frac{161}{8}T\right)L+\frac{147}{16}L^2$$ (3.3) $$R_3[L,T]=c_0(T)+c_1(T)L+c_2(T)L^2+c_3L^3.$$ (3.4) Initially, we shall eliminate $`T`$ as a variable by assuming that the Higgs pole mass is 100 GeV, in which case $`T=ln[M_H^2/M_t^2]=2ln(100/175.6)=1.126`$. As described in and , the $`[0|1]`$ Padé-approximant prediction for $`R_2`$ is $$R_2^{[0|1]}=R_1^2$$ (3.5) and the $`[1|1]`$ Padé-approximant prediction for $`R_3`$ is $$R_3^{[1|1]}=R_2^2/R_1.$$ (3.6) If the error of $`[N|1]`$ approximants in predicting $`R_{N+2}`$, the $`N+2`$ term in the perturbative series, is inversely proportional to $`N+1`$ i.e., if $$\frac{R_{N+2}^{[N|1]}R_{N+2}^{exact}}{R_{N+2}^{exact}}=\frac{A}{N+1},$$ (3.7) where $`A`$ is a constant— one may then utilize (3.5) and the exact value for $`R_2`$ within (3.7) to obtain $`A=(R_2^2R_1^2)/R_2`$. Substituting (3.6) and this estimate for $`A`$ into (3.7), one obtains an error-improved estimate for the unknown coefficient $`R_3`$ : $$R_3[L]=2R_2^3[L]/(R_1[L]R_2[L]+R_1^3[L])$$ (3.8) To obtain estimates of the coefficients $`c_0,c_1,c_2,c_3`$ within (3.4), we match the scale dependence of (3.4) to that of (3.8) over the purely perturbative $`L>0`$ region \[corresponding to the ultraviolet scales $`\mu >m_t(\mu )`$\] through use of the moment integrals $$N_k(k+2)_0^1𝑑ww^{k+1}R_3(w),$$ (3.9) where $`w=m_t^2(\mu )/\mu ^2[L=ln(w)]`$. Substitution of (3.4) into the integrand of (3.9) yields the following expressions for the first four moments : $$N_1=c_0+c_1+2c_2+6c_3,$$ (3.10a) $$N_0=c_0+\frac{1}{2}c_1+\frac{1}{2}c_2+\frac{3}{4}c_3,$$ (3.10b) $$N_1=c_0+\frac{1}{3}c_1+\frac{2}{9}c_2+\frac{2}{9}c_3,$$ (3.10c) $$N_2=c_0+\frac{1}{4}c_1+\frac{1}{8}c_2+\frac{3}{32}c_3.$$ (3.10d) However, explicit numerical estimates of these four moments may be obtained via substitution of (3.2) and (3.3) for the respective factors of $`R_1[L]`$ and $`R_2[L]`$ appearing in (3.8), and by subsequent substitution of this estimate for $`R_3[L]`$ into the integrand of (3.9) \[with $`L=ln(w)`$\]. For $`M_H=100\mathrm{GeV}`$ the resulting estimates are $$N_1=5102.9,N_0=3542.9,N_1=3131.8,N_2=2944.7.$$ (3.11) Substitution of these values into (3.10) yields the following predicted values $`(c_i^{Pad\stackrel{´}{e}})`$ for the four-loop terms in the $`Hgg`$ decay rate (2.7): $$c_0^{Pad\stackrel{´}{e}}=2453,c_1^{Pad\stackrel{´}{e}}=1772,c_2^{Pad\stackrel{´}{e}}=378.4,c_3^{Pad\stackrel{´}{e}}=20.27.$$ (3.12) The exact values of $`c_1,c_2`$, and $`c_3`$ were determined via $`RG`$ methods in the previous section. For $`M_H=100\mathrm{GeV}`$ \[i.e. for $`T=1.126`$\], these values are found from (2.11-13) to be $$c_1=1789.02,c_2=385.568,c_3=21.4375.$$ (3.13) Comparing (3.12) and (3.13), one finds the relative error $`\left[\delta c_i(c_i^{Pad\stackrel{´}{e}}c_i)/c_i\right]`$ of the Padé estimates for $`c_1,c_2`$ and $`c_3`$ to be -0.95%, -1.9%, and -5.5%, respectively. Such accuracy in predicting the known four-loop terms in the $`Hgg`$ rate suggests that the prediction for $`c_0`$ in (3.12) is a credible one. One way to test the stability of this prediction is to utilize the true values of $`c_1,c_2,c_3`$ within the moment expressions (3.10) to obtain \[for the numerical values (3.11) already obtained for these moments\] four independent determinations of $`c_0`$. We then find that $`(3.10a):c_0=2412,`$ $`(3.10b):c_0=2436,`$ $`(3.10c):c_0=2444,`$ $`(3.10d):c_0=2447,`$ (3.14) results all within 2% of that in (3.12). An alternative approach to matching the four-loop coefficients $`c_i`$ within (3.4) to the error improved Padé estimate (3.8) is optimize the least-squares function $$\chi ^2[c_0,c_1,c_2,c_3]=_0^1\left[R_3(c_0c_1lnw+c_2ln^2wc_3ln^3w)\right]^2𝑑w,$$ (3.15) with $`R_3`$ in the integral given by (3.8). As before, factors of $`R_1`$ and $`R_2`$ appearing in (3.8) are given explicitly by (3.2) and (3.3) with $`L=ln(w)`$. One then finds for $`M_H=100\mathrm{GeV}`$ $`[T=1.126]`$ that $`\chi ^2(c_0,c_1,c_2,c_3)`$ $`=`$ $`4.06416887810^7+720c_3^2+4c_0c_2+c_0^2+2c_0c_1+24c_2^2+12c_1c_2+12c_0c_3+2c_1^2`$ (3.16) $`+`$ $`240c_2c_3+48c_1c_310205.87c_017507.05c_154106.23c_2234536.67c_3`$ The optimization requirement $$\frac{\chi ^2}{c_i}=0$$ (3.17) yields predictions remarkably close to those of (3.12), $$c_0^{\chi ^2}=2452,c_1^{\chi ^2}=1774,c_2^{\chi ^2}=377.2,c_3^{\chi ^2}=20.45,$$ (3.18) further confirming the stability of the estimation procedure. In Table 1, we have tabulated a set of predictions of the four-loop term $`c_0`$ for values of the Higgs mass between 100 and 175 GeV. Also tabulated are the errors in the predicted values for $`c_1,c_2,c_3`$, relative to the true values for these coefficients, as given in (2.11-13). Estimated values of $`c_1`$ and $`c_2`$ remain within 2% of their true values (2.11-13) over the range of Higgs masses given; the relative error of $`c_3`$ estimates remains below 7% for the same range. This consistency provides further support for the $`c_0`$ estimates presented in the final column. These $`c_0`$ estimates may be utilized to ascertain the convergence of the series (2.7). If we choose $`\mu =m_t(\mu )`$, the logarithmic factors $`L^k`$ within (2.7) all vanish. To evaluate the series $`S/x^2`$ in (2.7), we assume that $`m_t(m_t)M_t=175.6\mathrm{GeV}`$ and that $`\alpha _s(175.6)=0.10915=\pi x(175.6)`$, as evolved from $`\alpha _s(M_z)=0.119`$ , in which case $$S(\mu =175.6\mathrm{GeV})/x^2(175.6)=1+a_0x(175.6)+b_0x^2(175.6)+c_0x^3(175.6).$$ (3.19) For Higgs masses between 100-175 GeV, we tabulate in Table 2 the magnitudes of successive orders in (3.19). The final 4<sup>th</sup> order term is obtained from the appropriate Padé estimate for $`c_0`$ listed in Table 1. Table 2 shows that the four-loop term decreases from 10% to 3% of the leading contribution as $`M_H`$ increases from 100 to 175 GeV. Moreover, the convergence of the series is problematical in the absence of the estimate for the four loop term. Over the range of Higgs masses considered, the three-loop contribution is between 18% and 33% of the leading (one-loop) contribution. The procedures delineated above can also be utilized to predict $`c_0`$’s explicit polynomial dependence of $`T`$, $$c_0(T)=a_0+a_1T+a_2T^2+a_3T^3,$$ (3.20) analogous to the expressions (2.11-13) obtained from $`RG`$-invariance for $`c_1(T)`$, $`c_2(T)`$, and $`c_3(T)`$. To extract the $`T`$ dependence of $`c_0`$, we first incorporate all known $`T`$-dependence into the least- squares function (3.15): $`\chi ^2[c_0(T)]={\displaystyle _0^1}\left[R_3[L,T]\left\{c_0(T)+c_1(T)L+c_2(T)L^2+c_3(T)L^3\right\}\right]^2𝑑w.`$ (3.21) In (3.21), $`Llnw`$. The quantities $`c_1(T)`$, $`c_2(T)`$ and $`c_3(T)`$ are no longer optimizable variables as in (3.15), but are now the explicit polynomials (2.11-13) obtained in the previous section via $`RG`$-methods. The factor $`R_3[L,T]`$ in the integrand of (3.21) is just (3.8) generalized to include the explicit $`T`$-dependence of $`R_1`$ and $`R_2`$ \[eqs. (3.2) and (3.3)\]: $$R_3[L,T]=\frac{2R_2^3[L,T]}{R_1[L,T]R_2[L,T]+R_1^3[L,T]}.$$ (3.22) From (3.21), the requirement $`d\chi ^2/dc_0=0`$ generates $`c_0`$ as a function of $`T`$. Since $`Tln(M_H^2/M_T^2)`$, we restrict our attention to the region $`1T0`$ (i.e. to values for $`M_H`$ between $`M_t`$ and $`M_te^{1/2}=107\mathrm{GeV}`$). A set of values for $`c_0(T)`$ can be obtained via optimization of (3.21) over values of $`T`$ in this region: $`c_0(0)`$ $`=`$ $`735.7,c_0(0.1)=841.7,c_0(0.2)=955.6,c_0(0.3)=1078,`$ $`c_0(0.4)`$ $`=`$ $`1208,c_0(0.5)=1346,c_0(0.6)=1493,c_0(0.7)=1649,`$ $`c_0(0.8)`$ $`=`$ $`1814,;c_0(0.9)=1988,c_0(1.0)=2170.`$ (3.23) We obtain a least-squares fit of these results to the form (3.20) by optimizing $$\chi ^2[a_0,a_1,a_2,a_3]\underset{i=0}{\overset{10}{}}\left[c_0(i/10)\left(a_0\frac{a_1i}{10}+\frac{a_2i^2}{100}\frac{a_3i^3}{1000}\right)\right]^2$$ (3.24) with respect to $`\{a_0,a_1,a_2,a_3\}`$, and find that $$c_0(T)=735.71020T+388.8T^225.41T^3.$$ (3.25) Each coefficient listed above is within 5% relative error of the corresponding coefficient obtained via a least squares fit of the $`c_0`$ values displayed in Table 1 to the polynomial form (3.20): $$c_0(T)=755.91029T+394.3T^226.74T^3.$$ (3.26) Note that (2.11), (2.12), (2.13) and the Padé-estimates (3.25) or (3.26) specify all the logarithmic coefficients within the full four loop contribution (3.4) to the $`Hgg`$ decay rate (2.6,7), and comparison of these results with future perturbative calculations should yield information which can be employed to further improve Padé estimation procedures. The exact $`Hgg`$ rate (2.6) is necessarily a scale invariant physical quantity. The factor $`S[x(\mu ),L(\mu ),T]`$ within (2.6) will exhibit residual scale-dependence \[i.e. $`\mu `$-dependence\] only as a consequence of truncation of the series $`S`$ to a given order of perturbation theory. In the section that follows, the incorporation of the four-loop coefficients $`c_{03}`$ of $`S[x(\mu ),L(\mu ),T]`$ is seen to eliminate virtually all of this residual scale dependence. ## 4. The $`Hgg`$ Rate In Fig. 1, the $`\mu `$-dependence of three- and four-loop expressions for $`S[x^{(6)}(\mu )`$, $`L(\mu ),T]`$ is plotted for the case of a 140 GeV Higgs mass $`[T=2ln(140/175.6)]`$. The four-loop term within $`S`$ is evaluated through use of the Padé estimate $`c_0=1306`$ (Table 1) in conjunction with eqs. (2.11-2.13) for the $`RG`$-accessible coefficients $`c_1,c_2`$, and $`c_3`$. The running coupling $`\alpha _s(\mu )`$ and running mass $`m_t(\mu )`$ occurring within (2.7) are evolved via four-loop $`\beta `$ and $`\gamma `$-functions from physical reference values $`\alpha _s(M_z)=0.119`$ and $`m_t(m_t)=175.6\mathrm{GeV}`$ . Figure 1 shows that the four-loop expression eliminates virtually all of the residual scale dependence still evident in the three-loop rate in the region $`M_t\mu 30\mathrm{GeV}`$. The four-loop rate is observed to have a local minimum at $`\mu =43.5\mathrm{GeV}`$ and a weak local maximum at $`\mu =89.5\mathrm{GeV}`$. The values of $`S[x(\mu ),L(\mu ),T]`$ at both of these points of minimal-sensitivity differ by only 0.2%, indicative of the flatness of $`S`$ between these two points. In Table 3, values for $`S[x(\mu ),L(\mu ),T]`$ as well as the term-by-term series $`=x^2[1+R_1x+R_2x^2+R_3x^3]`$ within $`S`$ are displayed for a variety of $`\mu `$-values of interest. The table displays the relative size of successive terms $`R_nx^n`$ at the minimal-sensitivity points $`\mu =43.5`$ and $`89.5\mathrm{GeV}`$ in addition to the points $`\mu =140\mathrm{GeV}(=M_H)`$ and $`\mu =175\mathrm{GeV}(M_t)`$. The relative magnitude of the four-loop term $`R_3x^3`$ is seen to be less than 1% of the leading term in the series (unity) over the entire region between the minimal-sensitivity values of $`\mu `$. Of particular interest in this range are those values of $`\mu [47.0\mathrm{GeV}`$ and $`73.5\mathrm{GeV}]`$ at which the four-loop term $`R_3x^3`$ effectively vanishes. These values correspond to the two points in Figure 1 at which the 3-loop and 4-loop curves cross. The usual approach towards extracting information from an asymptotic series $`_{n=0}R_nx^n(R_0=1)`$ is to sum only that series’ decreasing terms, i.e. to evaluate $`_{n=0}^n^{}R_nx^n`$ by choosing $`n^{}`$ such that $`|R_n^{}x^n^{}|`$ is a minimum. By choosing $`\mu `$ so as to have $`R_3(\mu )`$ vanish, one can then argue (for this choice of $`\mu `$) that $`n^{}=3`$, suggesting that such a value for $`\mu `$ is optimal for estimating the series from its known terms. Of course, such an interpretation rests on the assumed increase of terms $`|R_nx^n|`$ subsequent to $`n^{}=3`$; all we can really be certain of is that $`|R_4x^4|>|R_3x^3|(=0)`$ at such a value of $`\mu `$. Nevertheless, Table 3 shows that both values of $`\mu `$ for which $`R_30`$ lie between the two minimal-sensitivity points $`[\mu =43.5`$ and $`89.5\mathrm{GeV}]`$, and that the $`\mu `$-sensitive factor $`S[x(\mu ),L(\mu ),T]`$ within the rate (2.6) varies by less than 0.2% over this entire region. Over the full range of $`\mu `$ values displayed in Table 3 $`(43.5\mathrm{GeV}<\mu <175\mathrm{GeV})`$, the four-loop series term $`(R_3x^3)`$ varies between 0 and 5% of the leading one-loop order term (unity). Surprisingly, however, the rate $`S[x(\mu ),L(\mu ),T]`$ displays a relative spread of values $`\mathrm{\Delta }S/S0.4\%`$ over the same range of $`\mu `$, indicative of a substantial reduction in residual scale dependence. In Figures 2 and 3 we exhibit the residual scale dependence of $`S[x(\mu )`$, $`L(\mu ),T]`$ to three and four-loop order for Higgs masses of 100 GeV and 175 GeV, respectively. These figures show the same reduction in scale dependence evident in Figure 1 when $`M_H=140\mathrm{GeV}`$. For the case of $`M_H=100\mathrm{GeV}`$, for example, Table 4 shows that the 4-loop term $`R_3x^3`$ varies between zero and 10% of the leading-order 1-loop term between the local minimum at $`\mu =29\mathrm{GeV}`$ and $`\mu M_t`$. Over this same range of $`\mu `$, the four-loop values of $`S[x(\mu ),L(\mu ),T]`$ remain within 1% of each other. For $`M_H=175\mathrm{GeV}`$, the four-loop term is between zero and 3% of the leading-order term between the local minimum at $`\mu =48.5\mathrm{GeV}`$ and $`\mu M_t`$, whereas the full four-loop expression for $`S[x(\mu ),L(\mu ),T]`$ exhibits a relative spread of only 0.1% \[Table 5\]. Such scale independence is, of course, a reflection of the $`RG`$-invariance (2.8) of $`S[x(\mu ),L(\mu ),T]`$, which has been utilized explicitly to obtain the coefficients $`c_{13}`$. Nevertheless, the coefficient $`c_0`$, which is not perturbatively accessible via (2.8) but is obtainable (at present) only by Padé approximant methods, appears to be precisely what is required to eliminate virtually all residual $`\mu `$-dependence in the rate arising from truncation. For example, if $`M_H=100\mathrm{GeV}`$, the factor $`c_0x^3(\mu )`$ in isolation contributes more than 10% of the leading order contribution (unity) for $`\mu \stackrel{<}{_{}}M_t`$. Moreover, this contribution increases as $`\mu `$ decreases. Nevertheless, the particular choice $`c_0=2453`$ appears to ensure that the overall spread in $`S`$ remains within 1%, despite the potentially large contribution to this spread from $`c_0x^3(\mu )`$. Evidently the $`\mu `$ dependence of this term serves to cancel the residual $`\mu `$-dependence of the remaining terms in the series. <sup>1</sup><sup>1</sup>1The reduction of scale-dependence via Padé approximant methods is discussed in detail by Gardi . This near cancellation of residual scale-dependence makes possible a set of credible predictions for the rate (2.6). In the fourth column of Table 6 we have tabulated the $`Hgg`$ decay rates for $`M_H=100,125,140,150`$, and $`175\mathrm{GeV}`$ in the $`m_b0,M_H^2<<4M_t^2`$ limit. The largest source of uncertainty for these predictions is in the value for $`\alpha _s(M_z)=`$ ($`0.119\pm 0.002`$ ), which should lead to 4% uncertainty in the rates presented in Table 6. Fermion mass effects (i.e. the departures from the $`m_b0,M_H^2<<4M_t^2`$ assumptions implicit in the derivation of the $`Hgg`$ rate) can be accommodated in leading order by replacing the factor of unity in (3.1) with the following $`m_b`$\- and $`M_t`$-sensitive terms : $$11+\delta _m=\frac{9}{16}\left[(A_t+ReA_b)^2+(ImA_b)^2\right],$$ (4.1) $$A_t=2\left[\tau _t+(\tau _t1)\left(sin^1(\sqrt{\tau _t})\right)^2\right]/\tau _t^2,\tau _tM_H^2/4M_t^2,$$ (4.2) $$A_b=2\left[\tau _b+(\tau _b1)f(\tau _b)\right]/\tau _b^2,\tau _bM_H^2/4m_b^2,$$ (4.3) $$f(\tau )\frac{1}{4}\left[ln\left(\frac{1+\sqrt{11/\tau }}{1\sqrt{11/\tau }}\right)i\pi \right]^2.$$ (4.4) The right hand side of (4.1) is easily seen to approach unity when $`m_b0,M_H^2/4M_t^20`$: $$\underset{\tau _t0}{lim}A_t=\frac{4}{3};\underset{\tau _b\mathrm{}}{lim}A_b=0.$$ (4.5) The leading mass correction $`\delta _m`$, as defined in (4.1), is tabulated for various Higgs boson masses (with PDG fermion-mass values $`m_b=4.2\mathrm{GeV}`$, $`M_t=175.6\mathrm{GeV}`$) in the fifth column of Table 6. This lowest-order fermion mass correction can be incorporated into the rate (2.6) by its inclusion into the series (3.1) $$S=x^2[1+R_1x+R_2x^2+R_3x^3+\delta _m].$$ (4.6) The sixth column of Table 6 tabulates the $`Hgg`$ rate with this correction included. Because $`\delta _m`$ is a leading order correction, there is genuine ambiguity as to whether ”physical” or running masses should be incorporated into this correction. The former choice yields a manifestly scale-dependent contribution $`\mathrm{\Delta }S=\delta _m^px^2(\mu )`$ to the rate $`S`$. One could argue for the incorporation of running masses $`m_b(\mu ),m_t(\mu )`$ in (4.2) and (4.3); i.e., $`\mathrm{\Delta }S=\delta _m(\mu )x^2(\mu )`$, with $`\delta _m(\mu )`$ calculated via (4.1-4) with $`\tau _t=M_H^2/4m_t^2(\mu )`$ and $`\tau _b=M_H^2/4m_b^2(\mu )`$, and with $`\mu `$ identified consistently with the minimal-sensitivity scale tabulated in Column 2 of Table 6. In Column 7 of Table 6 we have tabulated $`\delta _m(\mu )`$ utilizing 6-active-flavour running masses referenced to $`m_t(m_t)=175.6\mathrm{GeV}`$, $`m_b(m_b)=4.2\mathrm{GeV}`$, and in Column 8 we have listed the corresponding $`Hgg`$ decay rates. Theoretical uncertainty associated with the masses utilized in the leading correction-factor $`\delta _m`$ is reflected in the differing rates of Column 6 and Column 8. This discrepancy is seen to be at most a 2% effect. The leading order mass-correction factor $`\delta _m`$ in (4.6) is seen from a comparison of Table 6 to Tables 3-5 to be generally smaller than the three-loop term $`R_2x^2`$ in (4.6), but somewhat larger than the four-loop term $`R_3x^3`$ also appearing in (4.6) \[$`\delta _m`$ is smaller than $`R_3x^3`$ when $`\mu =M_t`$, as in Table 2\]. Since the next order of fermion mass corrections is suppressed by an additional power of $`\alpha _s`$, we anticipate such next-to-leading-order fermion mass corrections to be well within 1% of the total rate. In any case, both leading-order mass corrections $`\delta _m`$ and the four-loop contribution $`R_3x^3`$ we obtain are sufficiently small for credible estimates of the $`Hgg`$ rate on the basis of the terms in (4.6). It is also worth noting from Table 6 that $`\delta _m^p`$ itself is only 1% of the leading-order one loop term \[normalized to unity in (4.6)\] when $`M_H=140150\mathrm{GeV}`$, and that $`\delta _m(\mu )`$ is comparably small when $`M_H=150\mathrm{GeV}`$. Hence, the rates tabulated in Table 6 are optimally valid for Higgs masses in the 140-150 GeV range. ## 5. The $`Hb\overline{b}`$ rate The leading QCD scalar-current correlation function contributions to the $`Higgsb\overline{b}`$ decay rate are known from explicit calculation to $`𝒪(\alpha _s^3)`$, with secondary $`𝒪(m_b^2/M_H^2)`$ power corrections known to $`𝒪(\alpha _s^2)`$, as given in : $`\mathrm{\Gamma }(Hb\overline{b})`$ $`=`$ $`{\displaystyle \frac{3G_F}{4\sqrt{2}\pi }}M_Hm_b^2(M_H)[\mathrm{\Pi }_{scalar}(\mu ^2=s=M_H^2).`$ (5.1) $`.+{\displaystyle \frac{m_b^2(M_H)}{M_H^2}}(640{\displaystyle \frac{\alpha _s(M_H)}{\pi }}87.72\left({\displaystyle \frac{\alpha _s(M_H)}{\pi }}\right)^2)].`$ The factor $`\mathrm{\Pi }_{scalar}`$ in (5.1) is the imaginary part of the QCD correlation function for the quark-antiquark scalar current normalized to unity in leading order : $`\mathrm{\Pi }_{scalar}[\mu ,s,x(\mu )]=1`$ $`+`$ $`\left({\displaystyle \frac{17}{3}}+2ln\mu ^2/s\right)x(\mu )`$ (5.2) $`+`$ $`\left(29.1467+{\displaystyle \frac{263}{9}}ln\left({\displaystyle \frac{\mu ^2}{s}}\right)+{\displaystyle \frac{47}{12}}ln^2\left({\displaystyle \frac{\mu ^2}{s}}\right)\right)x^2(\mu )`$ $`+`$ $`\left(41.7576+238.381ln\left({\displaystyle \frac{\mu ^2}{s}}\right)+94.6759ln^2\left({\displaystyle \frac{\mu ^2}{s}}\right)+7.61574ln^3\left({\displaystyle \frac{\mu ^2}{s}}\right)\right)x^3(\mu )`$ $`+`$ $`\left(d_0+d_1ln\left({\displaystyle \frac{\mu ^2}{s}}\right)+d_2ln^2\left({\displaystyle \frac{\mu ^2}{s}}\right)+d_3ln^3\left({\displaystyle \frac{\mu ^2}{s}}\right)+d_4ln^4\left({\displaystyle \frac{\mu ^2}{s}}\right)\right)x^4(\mu )`$ $`+`$ $`\mathrm{}`$ where $`x(\mu )=[\alpha _s(\mu )]_{n_f=5}/\pi `$, corresponding to five active flavours. In ref., a detailed asymptotic Padé-approximant procedure is presented for estimating the coefficients $`d_{04}`$ in (5.2). The methodology is virtually identical to that of Section 3 above, except that five moments are now calculated for $`R_4[w]`$, the coefficient of $`x^4(\mu )`$, based upon the asymptotic error formula prediction $$R_4=\frac{R_3^2[R_2^3+R_1R_2R_32R_1^3R_3]}{R_2[2R_2^3R_1^3R_3R_1^2R_2^2]}.$$ (5.3) The results, as tabulated in Table 3 of ref. , are $$d_0=64.2,d_1=745,d_2=1180,d_3=253,d_4=15.4.$$ (5.4) The factors $`d_{14}`$ may be obtained directly from RG-invariance of the physical rate $$\mu ^2\frac{d}{d\mu ^2}\left[m_b^2(\mu )\mathrm{\Pi }_{scalar}[\mu ,s,x(\mu )]\right]=0,$$ (5.5) leading to the following values (also tabulated) in ): $$d_1=791.52,d_2=1114.7,d_3=260.06,d_4=14.755.$$ (5.6) The strong agreement between (5.6) and (5.4) suggests that the estimate for $`d_0`$ in (5.4) is a credible one. It is interesting to note that a much more naive Padé approach for estimating $`d_0`$, in which (5.3) was applied directly to $`\mathrm{\Pi }_{scalar}(\mu ^2=s=M_H^2)`$, yielded a value for $`d_0`$ (=67.25 ) surprisingly consistent with the estimate in (5.4). Only the latter estimate exhibits sensitivity to the logarithmic terms in $`\mathrm{\Pi }_{scalar}`$, which all vanish when $`\mu ^2=s`$. If we utilize the estimate for $`d_0`$ quoted in (5.4) within the $`Hb\overline{b}`$ rate, we find that the correlation-function factor within (5.1) is given to $`𝒪(\alpha _s^4)`$ by $$\mathrm{\Pi }_{scalar}(\mu ^2=s=M_H^2)=1+\frac{17}{3}x(M_H)+29.1467x^2(M_H)+41.7576x^3(M_H)+\underset{¯}{64}x^4(M_H).$$ (5.7) The underlined coefficient is, of course, $`d_0`$ as estimated in via asymptotic Padé-approximant methods. There is genuine value in having an estimate of this term, as it is generally larger than the $`𝒪(\alpha _s^2m_b^2/M_H^2)`$ final term in (5.1). For example, suppose that $`M_H=130\mathrm{GeV}`$, $`m_b(130\mathrm{GeV})=2.7\mathrm{GeV}`$ and $`x(130\mathrm{GeV})=0.114/\pi `$ (these values are consistent with those employed in Table 1 of ref. ). The relative magnitudes of the correlator contributions 95.5) to the $`Hb\overline{b}`$ rate are seen to be $$\mathrm{\Pi }_{scalar}\left(\mu ^2=s=(130\mathrm{GeV})^2\right)=1+0.2056+0.0384+0.0020+\underset{¯}{0.00011}.$$ (5.8) All but the underlined Padé-estimate term are tabulated in Table 1 of ref. . Comparison to the power-suppressed terms in (5.1), $$\frac{m_b^2(M_H)}{M_H^2}\left(640x(M_H)87.72x^2(M_H)\right)\begin{array}{c}\\ \\ _{M_H=130\mathrm{GeV}}\end{array}0.00260.000620.00005,$$ (5.9) (also tabulated in ) reveals that the $`𝒪(\alpha _s^4)`$ term in (5.8) is double the magnitude of the $`𝒪(\alpha _s^2)`$ term in (5.9). Hence the Padé-estimated $`𝒪(\alpha _s^4)`$ term in (5.7) enables one to utilize the full precision available in the known power-suppressed contributions to (5.1). However, one cannot anticipate such precision experimentally for many years to come. ## 6. Summary In the preceding sections, we have demonstrated how asymptotic Padé-approximant methods may be utilized to estimate the unknown $`𝒪\left(\alpha _s^5\right)`$ coefficients $`\{c_0,c_1,c_2,c_3\}`$ within the $`H2g`$ decay rate (2.7). Such estimates for $`\{c_1,c_2,c_3\}`$ are seen (Table 1) to be within a few percent of the true values for these coefficients, which can be extracted via renormalization-group methods. This accuracy supports corresponding asymptotic Padé-approximant estimates of the renormalization-group-inaccessible coefficients $`c_0`$ presented in Table 1. Moreover, the inclusion of $`𝒪\left(\alpha _s^5\right)`$ terms within (2.7) is seen to virtually eliminate the scale-parameter dependence of the rate over an astonishingly large range of the scale-parameter $`\mu `$, typically $`0.3M_H\stackrel{<}{_{}}\mu \stackrel{<}{_{}}M_t`$. Inclusion of estimated $`𝒪\left(\alpha _s^5\right)`$ corrections in conjunction with leading-order fermion-mass corrections to the rate (Table 6) are seen to reduce the perturbative uncertainty in the $`H2g`$ decay rate from $`𝒪(20\%)`$ to $`𝒪(2\%)`$. Four-loop corrections to the $`Hb\overline{b}`$ decay mode are also presented, which are seen to reduce the very small perturbative uncertainty of this dominant hadronic mode by an additional order of magnitude. ## Acknowledgement We are grateful for support from the Natural Sciences and Engineering Research Council of Canada.
warning/0004/cond-mat0004207.html
ar5iv
text
# Equilibrium and dynamical properties of the ANNNI chain at the multiphase point ## I Introduction The ANNNI chain is one of the simplest systems with competing interactions. It is defined by the following Ising spin Hamiltonian: $`H={\displaystyle \underset{i=1}{\overset{L}{}}}(J_1s_is_{i+1}+J_2s_is_{i+2}),s_i=\pm 1.`$ (1) For $`J_2>0`$, the interactions are competing and one can have different ground states depending on the relative strengths of the interactions. A specially interesting case is the point $`J_1=2J_2`$, the so-called multiphase point, where the ground state is no longer unique. It can be shown that any spin configuration, which does not have three consecutive spins of the same sign, is a ground state. For a chain of length $`L`$, the number of ground states $`\mu ^L`$, where $`\mu =(\sqrt{5}+1)/2`$ is the golden mean. Thus there are an exponentially large number of degenerate ground states and the system has finite zero-temperature entropy per spin. The model has been extensively studied both in one and higher dimensions and is known to have a rich and interesting phase diagram . In this paper we consider some aspects of the equilibrium and dynamical behaviour of the ANNNI chain at the multiphase point. In our equilibrium study we consider the effect of softening the spins, that is allowing them to take continuous instead of discrete values. It is usual in the study of spin models to consider soft-spin versions of discrete spin models. A well-known example is the Ginzberg-Landau Hamiltonian which is a continuum version of the discrete Ising model. Other examples occur in the study of spin glass models. For instance, the soft spin version of the Sherrington-Kirkpatrick (SK) model was studied in the context of dynamics. The reason for going to soft-spin versions is that they are often more amenable to theoretical approaches. It is usually expected that qualitatively the soft and hard spin versions should show similar behaviour. For models with multiple ground states, arising out of frustration, softness may however change the degeneracy completely, as may be seen in a three-spin example, or as in the present case as we shall show here. The effect of spin-softening in systems with competing interactions has been studied earlier by several authors. Seno and Yeomans have looked at the effect of softening spins at the multiphase point of a clock-model. They find, using a perturbative method, that as a softness parameter is varied the system goes through a series of different ground states. In this work we use a similar perturbative method to prove that the macroscopic degeneracy of the ground state in the ANNNI model is lifted by the smallest amount of softness. We then show explicitly how the release of the zero-temperature entropy results in qualitative differences in the low temperature properties of the system. This is similar to the recently observed phenomena of entropy release in spin-ice systems . We also construct an effective hard-spin Hamiltonian to describe the low- temperature properties of the soft-spin model. We have also performed Monte Carlo simulations on the soft-spin model and verified the low-temperature predictions of the effective Hamiltonian. In the second part of the paper we look at the dynamical properties of the system. As noted before, the ANNNI model at the multiphase point has a large number of degenerate ground states. It is, therefore, of interest to look at dynamical properties of the system at low temperatures. Here we use Kawasaki dynamics to evolve the system and consider zero temperature properties only. Thus two nearest neighbor spins flip with a rate $`\gamma `$, provided both magnetization and energy are conserved. This dynamics has been studied earlier by Das and Barma . In this paper we extend their studies by using the correspondence between $`W`$-matrices for stochastic processes and quantum spin chains. The correspondence between the stochastic Fokker-Planck operator and quantum chains has often been exploited to derive dynamical properties. For instance, the scaling, with system size, of the first excited state of the quantum Hamiltonian gives the dynamical exponent of the stochastic process. A well-known example where this correspondence has been used is in exclusion processes , which are stochastic models of hard-core diffusing particles. For such processes, it has been possible to exactly calculate the dynamic exponent by solving the corresponding quantum model, namely the Heisenberg model . The dynamics considered by us is very similar to the symmetric exclusion process (SEP) but with added restrictions on allowed moves. To see this we first note that with the present dynamics, nearest neighbor spins with opposite signs flip, provided that the resulting configuration satisfies the ground state constraint of no three successive spins having same signs. Identifying up spins with particles and down spins with holes we see that the dynamics is equivalent to hard core particles diffusing on a lattice with the constraint that there cannot be three successive particles or holes. An interesting question is whether these rather strong constraints make the system different from the SEP. Earlier numerical work seems to suggest that the dynamics still behaves like the SEP. We note that there have been some other recent studies on exclusion processes with constraints on allowed configurations . These cases are solvable by the Bethe ansatz and show the same behaviour as the unconstrained model. Here we address this question of the effect of the constraints by studying the quantum Hamiltonian. By means of exact numerical diagonalization for finite chains and through analytic bounds, we have tried to understand the differences and similarities of the present Hamiltonian with the Heisenberg Hamiltonian for the SEP. We also discuss the different symmetry properties of the two quantum models. The Heisenberg model has full rotational symmetry and this has several important implications some of which are of direct relevance in understanding the original stochastic process. For example it implies that two-point time correlations in the SEP do not depend on the number of particles. The present model, on the other hand is only invariant under rotations in the $`XY`$ plane. The rest of the paper is divided into two sections. In section (II), we consider equilibrium properties of the soft-spin model while in section (III) we consider the dynamics of the hard-spin model. Section (IV) contains a summary of our main results and a few concluding remarks. ## II Soft-spin ANNNI model We consider the following soft spin version of the ANNNI model: $`H_s={\displaystyle \underset{i}{}}J(2s_is_{i+1}+s_is_{i+2})`$ $`+`$ $`ag(s_i^4/4s_i^2/2),`$ (3) $`s_i(\mathrm{},\mathrm{})`$ where $`a`$ is a dimensionless parameter which controls the amount of softness. In the limit $`a\mathrm{}`$ we get the hard-spin model. We will set $`g=1`$ since there is no loss of generality in doing so. Let us first look at the ground states of the soft-spin Hamiltonian given by Eqn. (3). To do so we look at the extrema of $`H_s`$ which are obtained by setting $`H/s_i=0`$ for all $`i`$. This gives: $`2J(s_{i+1}+s_{i1})+J(s_{i+2}+s_{i2})+a(s_i^3s_i)=0`$ (4) Solving this set of coupled nonlinear equations in general is very difficult. However for small values of the parameter $`1/a`$ we can obtain the solutions perturbatively. For $`a\mathrm{}`$ all configurations, $`\{s_i\}`$, with $`s_i=0,\pm 1`$ are solutions. Those with $`\{s_i=\pm 1\}`$ correspond to the minima. For finite but large $`a`$ we try to obtain the solutions perturbatively with $`1/a`$ acting as the perturbation parameter. We denote the unperturbed minima by the set $`\{t_i=\pm 1\}`$. Let us try the following perturbative expansion: $`s_i={\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}t_i^{(n)}({\displaystyle \frac{1}{a}})^n,`$ (5) where the coefficients $`t_i^{(n)}`$ are independent of $`a`$ and $`t_i^{(0)}t_i=\pm 1`$ correspond to the unperturbed solutions in the limit $`a\mathrm{}`$. Plugging this into Eqn. (4), we get $`J[2(t_{i+1}+t_{i1})+(t_{i+2}+t_{i2})]+{\displaystyle \frac{J}{a}}[2(t_{i+1}^{(1)}+t_{i1}^{(1)})+`$ (6) $`(t_{i+2}^{(1)}+t_{i2}^{(1)})]+2t^{(1)}_i+{\displaystyle \frac{1}{a}}(3t_i(t_i^{(1)})^2+2t_i^{(2)})+O({\displaystyle \frac{1}{a^2}})=0.`$ (7) Equating different powers of $`1/a`$ to zero we then get: $`t_i^{(1)}={\displaystyle \frac{J}{2}}[2(t_{i+1}+t_{i1})+(t_{i+2}+t_{i2})]+O({\displaystyle \frac{1}{a}})`$ (8) $`t^{(2)}={\displaystyle \frac{J}{2}}[2(t_{i+1}^{(1)}+t_{i1}^{(1)})+(t_{i+2}^{(1)}+t_{i2}^{(1)})]{\displaystyle \frac{3}{2}}t_i(t_i^{(1)})^2`$ (9) and so on. Thus we get $`2^L`$ perturbed minima given by the above perturbation series. The energies corresponding to these minima can now be found by putting these solutions into the expression for energy in Eqn. (3). We thus get $`E=E_0+E_1+E_2+O(1/a^2),\mathrm{where}`$ (10) $`E_0={\displaystyle \frac{La}{4}}`$ (11) $`E_1={\displaystyle \underset{i}{}}J(2t_it_{i+1}+t_it_{i+2})`$ (12) $`E_2={\displaystyle \frac{1}{a}}{\displaystyle \underset{i}{}}[(t_i^{(1)})^2+2J(t_it_{i+1}^{(1)}+t_{i+1}t_i^{(1)})+J(t_it_{i+2}^{(1)}+t_{i+2}t_i^{(1)})]`$ (13) $`={\displaystyle \frac{J^2}{2a}}{\displaystyle \underset{i}{}}(5+4t_it_{i+1}+4t_it_{i+2}+4t_it_{i+3}+t_it_{i+4}).`$ (14) In the above expansion, $`E_0`$ corresponds to the unperturbed energy, while $`E_1`$ and $`E_2`$ represent the corrections resulting from the perturbation. In the $`a\mathrm{}`$ limit the term $`E_1`$ causes the energy levels of the $`2^L`$ minima to split, with separation between them $`O(J)`$. We recognize $`E_1`$ as the Hamiltonian for the hard-spin ANNNI model. Thus the lowest energy level is still $`\mu ^L`$-fold degenerate. The term $`E_2`$ then causes a further splitting of the ground states into levels with separation $`O(J^2/a)`$. To see whether or not the macroscopic degeneracy of the ground state survives, we need to consider the interaction Hamiltonian corresponding to the energy term $`E_2`$. Since we are interested in the splitting of the lowest energy level of $`E_1`$, we only consider the restricted subspace of spin configurations which are ground states of $`E_1`$. In this subspace the Hamiltonian corresponding to $`E_2`$ can be rewritten as $`H_2={\displaystyle \frac{3LJ^2}{2a}}{\displaystyle \frac{J^2}{2a}}{\displaystyle \underset{i}{}}(2t_it_{i+2}+4t_it_{i+3}+t_it_{i+4}).`$ (15) Thus all the interactions are ferromagnetic. However the ground state of $`H_2`$ is not the state with all spins up, since this does not belong to the subspace of ground states of $`E_1`$. To find the ground state, we write the second term in $`H_2`$, which we denote by $`h_2`$, in the following form (the constant factor $`J^2/(2a)`$ is suppressed): $`h_2={\displaystyle \underset{i}{}}(2t_it_{i+2}+4t_it_{i+3}+t_it_{i+4})`$ (16) $`={\displaystyle \underset{i=(4n+1)}{}}ϵ(t_i,t_{i+1},t_{i+2},t_{i+3}t_{i+4},t_{i+5},t_{i+6},t_{i+7})`$ (17) $`\mathrm{where}ϵ(\mathrm{t}_1,\mathrm{t}_2,\mathrm{t}_3,\mathrm{t}_4\mathrm{t}_5,\mathrm{t}_6,\mathrm{t}_7,\mathrm{t}_8)=\mathrm{t}_1\mathrm{t}_3+\mathrm{t}_2\mathrm{t}_4+2\mathrm{t}_3\mathrm{t}_5+`$ (18) $`2t_4t_6+t_5t_7+t_6t_8+2t_1t_4+4t_2t_5+4t_3t_6+4t_4t_7+2t_5t_8`$ (19) $`+t_1t_5+t_2t_6+t_3t_7+t_4t_8`$ (20) and the index $`n`$ runs from $`0`$ to $`(L/41)`$ (we take $`L`$ to be an integral multiple of $`4`$). By enumerating the matrix elements $`ϵ(t_1,t_2,t_3,t_4t_5,t_6,t_7,t_8)`$ for all allowed spin configurations we find that the lowest energy configuration is obtained for the periodic sequence $`(\mathrm{})`$ and the five other configurations obtained by translating and flipping this. Thus we find that the infinite degeneracy of the ground state is removed and instead we get a six-fold degenerate ground state. We note that the procedure just outlined provides a straight forward method of finding the ground state of any spin Hamiltonian. By numerically solving Eqn. (4) for small lattice sizes $`(L=12)`$ and finding the minimum energy configurations for $`a`$ large enough $`(a=50)`$, we have verified that the perturbative solutions are quite accurate. The fact that softening of the spins results in removal of the exponential degeneracy of the ground state means that the finite zero temperature entropy is released and we expect it to show up in the behaviour of the low temperature specific heat. This leads to the soft-spin model having low-temperature properties very different from the hard-spin version as we shall now see. We note that the hard spin model is easily solvable by transfer-matrix methods and one can exactly compute various thermodynamic properties. In the soft-spin case the transfer-matrix eigenvalue equation becomes an integral equation which we have not been able to solve. Hence we have studied the model by Monte Carlo simulations. We have used a dynamics which allows three kinds of processes; (i) single spin-flip moves, (ii) moves in which two nearest neighbor spins are simultaneously flipped and (iii) moves which change the length of a spin. All three kinds of processes occur with usual Metropolis rates. The reason for allowing both single and double spin-flips is the following. We find that in the hard-spin case, equilibration times, with a single-spin flip dynamics, become very large at low temperatures. On the other hand, allowing for two-spin flips results in very fast equilibration. This is related to the fact that while the single spin-flip dynamics at $`T=0`$ is non-ergodic, including double-flips makes it ergodic. We expect a similar situation even in the case of soft-spins and so have included both (i) and (ii). Finally (iii) is necessary since the spins are now continuous variables and we need to be able to change their lengths. In order to compare the properties of the soft-spin model with those of the hard-spin one, it is necessary to subtract from the soft-spin free energy a part corresponding to the continuum degrees of freedom. We thus look at the following free energy: $`F=(1/\beta )[\mathrm{ln}Tre^{\beta H_s}+L\mathrm{ln}(2)\mathrm{ln}Tre^{\beta H_g}],`$ (21) where $`H_s`$ is as in Eqn. (3), $`H_g=_ia(s_i^4/4s_i^2/2)`$, and $`Tr`$ indicates integration over all spin variables. We note that the above expression for the free energy is equivalent to writing the partition function in the form $`Z=Tre^{\beta H}P(\overline{s})\mathrm{with}`$ (22) $`P(\overline{s})={\displaystyle \underset{i}{}}{\displaystyle \frac{2e^{\beta a(s_i^4/4s_i^2/2)}}{𝑑s_ie^{\beta a(s_i^4/4s_i^2/2)}}}.`$ (23) $`H`$ being the original hard-spin Hamiltonian and $`P(\overline{s})`$ a probability distribution over the spin variables. In the limit $`a\mathrm{}`$ this exactly reduces to the hard-spin partition function while at $`T\mathrm{}`$ one gets $`Z=2^L`$. ¿From our simulations we get properties corresponding to the first part of the free energy in Eqn. (21). The second part simply corresponds to a noninteracting system and its properties can be easily computed numerically. In Fig. 1 we plot the specific heat data $`C(T)`$ for both the soft-spin and hard-spin models. The hard-spin result is exact and corresponds to infinite system size while the soft-spin data is from simulations on a chain of length $`L=24`$. The values of various parameters used in the simulation were $`a=50`$ and $`J=1`$. The high temperature $`(T>1)`$ data was obtained by averaging over $`10^6`$ Monte Carlo steps while the low temperature data is over $`10^7`$ steps. As expected we find a second peak in the specific heat at low temperatures. For the hard-spin case the total area under the curve for $`C(T)/T`$ is equal to $`\mathrm{ln}(2/\mu )`$. The ground state entropy, $`\mathrm{ln}(\mu )`$, which is released when the spins are softened, is mostly accounted for by the area under the low temperature peak. The low temperature properties are quite well reproduced by the effective Hamiltonian, $`H_2`$, which describes the energy levels in the lowest band. The thermodynamic properties of $`H_2`$ can be exactly calculated by transfer matrix methods, both for finite system sizes and in the infinite size limit. In Fig. 2 we plot the soft-spin low temperature simulation data $`C(T)`$ for two system sizes and compare them with results obtained from the effective Hamiltonian. We see good agreement between the two. We also show the infinite system size $`C(T)`$ curve obtained from the Hamiltonian $`H_2`$. It is interesting to note that the peak value of the specific heat first increases with system size and then starts decreasing beyond a certain size. ## III Kawasaki Dynamics of the hard-spin ANNNI model at the multiphase point As for the usual exclusion process, the quantum Hamiltonian corresponding to our process can be easily written and is given by: $`=𝒫\{{\displaystyle \underset{k=1}{\overset{L}{}}}[(\sigma _k^+\sigma _{k+1}^{}`$ $`+`$ $`\sigma _k^{}\sigma _{k+1}^+)+`$ (25) $`{\displaystyle \frac{1}{2}}(\sigma _k^z\sigma _{k+1}^z1)]P_k\}𝒫`$ where $`\sigma _k^\alpha `$ are the usual Pauli matrices, $`P_k`$ are local projection operators given by $`P_k=(1\sigma _{k2}\sigma _{k1})(1\sigma _{k+2}\sigma _{k+3})/4`$ (26) and $`𝒫=_{k=1}^LP_k`$ is a global projection operator which projects onto the space of allowed states, i.e those that satisfy the ground state constraint. The spin-flip rate, $`\gamma `$, has been set to unity. Alternatively we can write the Fokker-Planck operator in the following form: $`={\displaystyle \underset{k=1}{\overset{L}{}}}(\theta _k+\theta _k^2)\mathrm{where}`$ (27) $`\theta _k=𝒫(\sigma _k^+\sigma _{k+1}^{}+\sigma _k^{}\sigma _{k+1}^+)𝒫.`$ (28) The term $`_k\theta _k^2`$ is the diagonal term since it corresponds to flipping an unequal pair twice. It is important to write the diagonal part carefully. For instance if in Eqn. (25), the local projection operators, $`P_k`$, were not present, the off-diagonal elements of $``$ would still be correct but the diagonal ones would be wrong. We now study the properties of this quantum Hamiltonian. Our interests are (a) to compare the symmetry properties and conservation laws of the present Hamiltonian with that of the Heisenberg model and (b) to obtain results on the energy gap and hence the dynamical exponent. ### A Symmetry properties and conservation laws of the quantum model We first observe that the $`z`$-component of the total spin, $`S^z`$, commutes with $``$. This simply implies conservation of spin or number of particles in the stochastic model. Thus we can classify energy states into sectors labelled by number of particles $`n`$. The constraints on allowed configurations means that for a lattice of length $`L`$ the number of particles can vary over the range $`[L/3]nL[L/3]`$ where $`[L/3]`$ denotes the smallest integer greater than or equal to $`L/3`$. It can be shown that except in the lowest and highest sectors, in every other case the dynamics is ergodic. It then follows from detailed balance that the steady state is one in which all allowed configurations in a given sector occur with equal probability. For the quantum model this means that the ground state in any sector is an equally weighted sum over all states ( For the special case where $`L`$ is a multiple of $`3`$, the lowest and highest sectors have $`3`$-fold degenerate ground states ). The other components of the total angular momentum $`S^x`$ and $`S^y`$ however do not commute with $``$. Thus the present Hamiltonian has $`U(1)`$ symmetry instead of the $`SU(2)`$ symmetry of the Heisenberg model. Also even though the ground states are degenerate, with one state in every $`S^z`$ sector, there is no analogue of the raising/lowering operator $`S^\pm `$. If there were such an operator then the entire eigenvalue spectrum in the $`n`$-particle sector would be a subset of the $`(n1)`$-particle sector (for $`n<L/2`$). By looking at the spectrum for finite sized lattices we have verified that this is not so. To study the presence of long-range order in the ground-state, we have calculated the two-point static correlation functions $`c_z(r)=<\sigma _0^z\sigma _r^z>`$ and $`c_\pm (r)=<\sigma _0^+\sigma _r^{}>`$ in the ground state for the half-filled sector. The simple characterization of the ground states in terms of disallowed subsequences enables calculation of ground-state expectation of any operator by means of transfer matrices. The transfer matrix method sums over all the different particle sectors, but in the thermodynamic limit the half filled sector dominates, and so we get correct results (To compute expectation values in other sectors one would need to introduce a chemical potential). Thus we find that $`c_z(r)=A\mathrm{cos}(\varphi 2\pi r/3)e^{r/\xi }`$ where $`\xi =1/\mathrm{log}((3+\sqrt{(}5))/2)=1.03904\mathrm{}`$ and $`A`$ and $`\varphi `$ are constants that have different values on odd and even sites. Fourier transforming $`c(r)`$ gives the structure factor $`<\sigma ^z(q)\sigma ^z(q)>`$ which has the form shown in Fig. 3. We note that it is non-vanishing at all $`q`$. The off-diagonal correlation can similarly be obtained using transfer matrices but the calculation becomes extremely cumbersome. Instead we have computed this correlation numerically for finite lattices and find that it saturates, for large $`r`$, to a constant value, which is given by $`<\sigma _{}>^2=0.02917\mathrm{}`$ (which has been obtained by using the transfer matrix method). Thus we find that ground-state correlation functions show the same behaviour as in the Heisenberg chain. For the Heisenberg model $`c_z(r)`$ is delta-correlated while $`c_\pm (r)`$ saturates to the value $`1/4`$ (which is much larger than its value in the present model). The presence of off-diagonal long range order means that $`U(1)`$ symmetry is broken in the ground state. This is analogous to the breaking of $`SU(2)`$ symmetry in the ground state of the Heisenberg model. On the other hand, consider the $`XXZ`$ chain defined by the Hamiltonian $`={\displaystyle \underset{k=1}{\overset{L}{}}}(\sigma _k^+\sigma _{k+1}^{}+\sigma _k^{}\sigma _{k+1}^+){\displaystyle \frac{\mathrm{\Delta }}{2}}\sigma _k^z\sigma _{k+1}^z.`$ (29) Away from the two isotropic points ($`\mathrm{\Delta }=\pm 1`$), this has the same symmetry as the present model. It has no long range order in the gapless phase ($`1<\mathrm{\Delta }<1`$) and all correlations $`<\sigma ^\alpha (0)\sigma ^\alpha (r)>`$ have power law decays. In the ferromagnetic phase ($`\mathrm{\Delta }>1`$), the model has a gap and full ferrromagnetic long-range order in the ground state, with ultra-local longitudinal correlations namely $`c_z(r)=1/4`$. Thus we see that as far as ground state correlations are concerned the present model is different from the anisotropic $`XXZ`$ chain even though they have the same symmetry properties. Our model is more similar in properties to the ferromagnet ($`\mathrm{\Delta }=1`$) but has a nontrivial depletion of the condensate, as well as a nontrivial $`<\sigma _0^z\sigma _r^z>`$ correlation. Finally we note that rotational invariance of the Heisenberg model means that two-point time correlations are completely determined by single magnon excitations and so have the same behaviour in any $`S_z`$ sector . This result does not hold in the case of the present model. A second conserved quantity in the model is the total linear momentum. This follows from the translation invariance of $``$. The momentum operator commutes both with $``$ and $`S^z`$ so that in each $`S^z`$ sector energy states can be labelled by their momentum. Clearly the ground state has zero momentum. ### B Results on the energy gap As is well known the first excited state of $``$ determines the decay of correlations for the stochastic process. Thus the energy gap $`\mathrm{\Delta }1/L^z`$ and this determines the dynamic exponent $`z`$. For the SEP, which corresponds to the Heisenberg ferromagnet, it is known that $`z=2`$. This simply reflects the diffusive modes in the dynamics. The dynamics studied here is very similar to the SEP but with the constraints on the allowed number of succesive particles and holes. An interesting question is whether these rather strong constraints change the dynamical exponent. Unlike the Heisenberg model where the Bethe ansatz is applicable and yields information on the eigenvalue spectrum, the Hamiltonian in Eqn. (25) is much more complicated and we have not been able to use the Bethe ansatz. We have looked at the eigenvalue spectrum by numerical diagonalization of $``$ for small system sizes and also through Monte Carlo simulations. We also obtain various analytic bounds on the energy levels. (i)Results of numerical diagonalization of $``$ and Monte Carlo simulations We have carried out exact diagonalization of the Hamiltonian in Eqn. (25) for chains of length upto $`L=22`$ at half filling. The diagonalization has been done in the momentum basis. This makes the Hamiltonian block diagonal and enables us to go to quite large chian sizes. We find that for small $`L`$ the first excited state occurs at total linear momentum $`q=\pi `$ and the gap seems to decreases as $`1/L`$. However from $`L=22`$ onwards, the first excited state shifts to $`q=2\pi /L`$ and the gap at this momentum decreases as $`1/L^2`$. In Fig. 4 we show the numerically obtained gaps at the two momenta as a function of system size. We also plot corresponding upper bounds on the gaps (to be derived in the next section). We note here that though it is usually the first excited state that determines the decay of correlations in the stochastic process, it is possible to construct correlation functions whose decay is governed by some other eigenvalue. As an example consider the operator $`Q=e^{i\frac{\pi }{L}_kk\sigma _k^z}`$. This is the so-called twist operator, first studied by Lieb, Schultz and Mattis . In this case, the decay of the correlation, $`<Q(0)Q(t)>`$ is determined by the lowest eigenvalue at momentum $`\pi `$ since the operator carries momentum $`\pi `$. In Fig. 5 we show the decay constant as determined from the correlation decay for different system sizes and compare them with those obtained from exact diagonalization. The correlation function is obtained from Monte Carlo simulations and can also be used for larger system sizes at which numerical diagonalization becomes too difficult. (ii)Exact Bounds We now find upper bounds on the first excited state. Consider the sector with states which have $`n`$ overturned spins. The bounds are obtained by constructing trial wave functions orthogonal to the ground state in each sector. Thus consider the operators $`\sigma ^z(q)=\frac{1}{\sqrt{L}}_k\sigma _k^ze^{ikq}`$ and the twist operator $`Q`$ defined in the previous section. Under translation these operators transform as $`T\sigma ^z(q)T^{}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{L}}}{\displaystyle \underset{k}{}}\sigma _{k+1}e^{ikq}=e^{iq}\sigma ^z(q)`$ (30) $`TQT^{}`$ $`=`$ $`e^{i\frac{\pi }{L}_kk\sigma _{k+1}^z}=e^{i2\pi d}Q`$ (31) where $`d=n/L`$ is the filling fraction of particles. If $`|0_n>`$ is the ground state in the $`n`$-particle sector, then the states $`\sigma ^z(q)|0_n>`$ and $`Q|0_n>`$ have momenta $`q`$ and $`2\pi d`$ respectively and for $`q0`$ are orthogonal to the ground state, which has zero momentum. Hence the following expectation values give us two different upper bounds on the gap: $`(a)e_z`$ $`=`$ $`{\displaystyle \frac{<\sigma ^z(q)\sigma ^z(q)>}{<\sigma ^z(q)\sigma ^z(q)>}}`$ (32) $`(b)e_Q`$ $`=`$ $`<Q^{}Q>`$ (33) where $`<\mathrm{}>`$ denotes ground state expectations. We now evaluate $`(a)`$ and $`(b)`$. We shall henceforth restrict ourselves to the half-filled sector only, though extensions to other sectors can be done. (a) To evaluate $`e_z`$ we first note that the numerator and denominator in Eqn. (32) can be written in the following equivalent form: $`<\sigma ^z(q)\sigma ^z(q)>`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{l}{}}e^{iql}<[\sigma _1^z,[,\sigma _{l+1}^z]]>`$ (34) $`<\sigma ^z(q)\sigma ^z(q)>`$ $`=`$ $`{\displaystyle \underset{l}{}}e^{iql}<\sigma _1^z\sigma _{l+1}^z>`$ (35) The commutator occurring in the above equation can be evaluated and gives: $`[\sigma _1^z,[`$ $``$ $`,\sigma _{l+1}^z]]=4𝒫[(\sigma _1^+\sigma _2^{}+\sigma _1^{}\sigma _2^+)p_1(\delta _{l,L}+\delta _{l,1})+`$ (37) $`4(\sigma _L^+\sigma _1^{}+\sigma _L^{}\sigma _1^+)p_L(\delta _{l,L1}\delta _{l,L})]𝒫^{}`$ Inserting this in Eqn. (35), and using translational invariance of the ground state we finally obtain: $`<\sigma ^z(q)`$ $``$ $`\sigma ^z(q)>`$ (38) $`=`$ $`4[1\mathrm{cos}(q)]<𝒫P_1(\sigma _1^+\sigma _2^{}+\sigma _1^{}\sigma _2^+)𝒫^{}>`$ (39) $`=`$ $`2[1\mathrm{cos}(q)]<𝒫P_1(1\sigma _1^z\sigma _2^z)𝒫>`$ (40) where the last step has been obtained using the fact that $`<0||0>=0`$. As noted before ground-state expectations of any operator can be computed using transfer matrices. The expectation value on the rhs of Eqn. (40) is thus found to have the limiting value $`(\mathrm{as}L\mathrm{})`$ $`<𝒫P_1(1\sigma _1^z\sigma _2^z)𝒫>=816/\sqrt{5}`$. The Fourier transform of $`c_z(r)`$, which gives the structure factor $`<\sigma ^z(q)\sigma ^z(q)>`$, has already been obtained and was plotted in Fig. 3. We note that it is non-vanishing at all $`q`$. Finally, from Eqns. (32,35) we get $`e_z`$ which is plotted in Fig. 6 along with the exact results from finite size diagonalization. Putting $`q=2\pi /L`$ and putting in all numerical factors, we get the following result: $`\mathrm{\Delta }19.78{\displaystyle \frac{\pi ^2}{L^2}}`$ (41) (b) We now obtain the other bound using the twist operator, $`Q`$. We first note the following properties of $`Q`$: $`Q^{}\sigma _l^+\sigma _{l+1}^{}Q|\{\sigma \}>`$ $`=`$ $`e^{i\frac{2\pi }{L}}\sigma _l^+\sigma _{l+1}^{}|\{\sigma \}>`$ (42) $`Q^{}\sigma _l^{}\sigma _{l+1}^+Q|\{\sigma \}>`$ $`=`$ $`e^{i\frac{2\pi }{L}}\sigma _l^{}\sigma _{l+1}^+|\{\sigma \}>.`$ (43) Using these relations we obtain $`<0|`$ $`Q^{}`$ $`Q|0>`$ (44) $`=`$ $`{\displaystyle \underset{k}{}}<0|Q^{}𝒫(\sigma _k^+\sigma _{k+1}^{}+\sigma _k^{}\sigma _{k+1}^+)P_k𝒫Q|0>`$ (46) $`{\displaystyle \underset{k}{}}{\displaystyle \frac{1}{2}}<0|Q^{}𝒫(\sigma _k^z\sigma _{k+1}^z1)P_k𝒫Q|0>`$ $`=`$ $`\mathrm{cos}(2\pi /L){\displaystyle \underset{k}{}}<0|𝒫(\sigma _k^+\sigma _{k+1}^{}+\sigma _k^{}\sigma _{k+1}^+)P_k𝒫|0>`$ (48) $`{\displaystyle \underset{k}{}}{\displaystyle \frac{1}{2}}<0|𝒫(\sigma _k^z\sigma _{k+1}^z1)P_k𝒫|0>`$ $`=`$ $`{\displaystyle \frac{L}{2}}[1\mathrm{cos}(2\pi /L)]<0|𝒫(1\sigma _k^z\sigma _{k+1}^z)P_k𝒫|0>,`$ (49) where in the last step we have again used $`<O||0>=0`$ and translational invariance of the ground state. The expectation value above has already obtained so that we get, for large $`L`$, the following bound for the gap at momentum $`q=\pi `$. $`\mathrm{\Delta }0.845{\displaystyle \frac{\pi ^2}{L}}`$ (50) In Fig. 4 we have plotted both the bounds and the exact finite size results at $`q=2\pi /L`$ and $`q=\pi `$ as functions of the system size. ## IV Summary In summary we have studied a one-dimensional spin model with competing interactions and studied its low-temperature equilibrium and dynamical properties. In the equilibrium case we have shown that low temperature properties of the soft-spin and hard-spin versions of the model can be very different. The hard-spin version of the model has an infinitely degenerate ground-state. Through a perturbative calculation we have shown, that as soon as we introduce the slightest amount of softness, the degeneracy is lifted. The ground state energy levels split to form a band which is separated from higher levels by $`\mathrm{\Delta }E=O(J)`$. The energy levels within this lowest band are described by an effective hard-spin Hamiltonian, containing ferromagnetic interactions upto fourth neighbour terms. This can be used to approximately derive the low temperature properties of the model. We find reasonably good agreement with results from Monte Carlo simulations of the soft-spin model. Our results indicate that the fixed-length ($`a\mathrm{}`$) limit is a singular one in our model at low temperatures. Since the ground state of the soft-spin model for large but finite $`a`$ is only six-fold degenerate, it would order into one of these six states as $`T0`$. This implies the occurrence of a zero-temperature phase transition and the existence of an appropriately defined correlation length that diverges as $`T`$ goes to zero. In the fixed-length ($`a\mathrm{}`$) limit, on the other hand, averaging over all the degenerate ground states leads to a finite correlation length even at $`T=0`$. These results suggest that it would be interesting to study the effects of softening the spins on the thermodynamic behavior of two- and higher-dimensional hard-spin models with extensive ground-state entropy. A well-known model of this kind is the nearest-neighbor Ising antiferromagnet on a triangular lattice . This model does not exhibit any phase transition at a non-zero temperature. The degeneracy-lifting effect of introducing magnitude fluctuations found in our study suggests that soft-spin versions of this and other similar models may exhibit finite-temperature phase transitions. Further investigation of this question would be very interesting. We believe that the removal of the exponential ground-state degeneracy by the introduction of spin-softness in the model studied here is a special case of a more general phenomenon in which the presence of additional degrees of freedom allows the system to relieve frustration and thus reduce the number of degenerate ground states. Coupling the hard spins to other degrees of freedom, such as elastic variables describing possible deformations of the underlying lattice, would probably have similar effects on the degeneracy of the ground state. It is interesting to note in this context that a “deformable” Ising antiferromagnet on a triangular lattice in which the Ising spins are coupled to elastic degrees of freedom exhibits a Peierls-type phase transition at a non-zero temperature. The ordering of the spins at this transition is accompanied by a distortion of the lattice. In general, it is expected that in real, physical systems, such couplings to other degrees of freedom, however weak, would induce some kind of ordering of the spins as the temperature is reduced toward zero, thereby avoiding the unstable situation of having a non-vanishing entropy per spin at $`T=0`$. Many disordered spin systems, such as spin glasses , exhibit a large number of nearly-degenerate metastable states arising out of frustration. To take an example, the SK model of infinite-range Ising spin glass is known to have an exponentially large number of local minima of the free energy (locally stable solutions of the TAP equations ) at sufficiently low temperatures. These local minima of the free energy become local minima of the energy at $`T=0`$. The presence of a large number (divergent in the thermodynamic limit) of nearly-degenerate metastable states is crucial in the development of the present understanding of the equilibrium and dynamic properties of this system at low temperatures. Our results about the lifting of degeneracy by the introduction of spin-softness raise the following interesting question: would the low-temperature properties of a soft-spin version of the SK model differ in any significant way from those of the original model? While soft-spin versions of the SK model have been used in studies of the dynamics, questions about how the number and properties of the metastable states of this model change as the spins are made soft have not been addressed in detail. Further investigation of these issues would be most interesting. Finally it is interesting to note that a similar way of lowering frustration is to make the coupling constants soft while keeping the spins hard. For example, in the case of the Edwards-Anderson Ising spin glass model, two versions have been studied . One is the $`\pm J`$ model where the nearest-neighbor coupling constants randomly take the discrete values $`\pm J`$ with equal probability. In the other case, the $`J`$s are chosen from a gaussian distribution. In $`d=2`$, both these cases are believed to have zero-temperature phase transitions, but the nature of the transition is different in the two cases. This difference again arises because of the different ground-state degenaracies in the two cases. In the $`\pm J`$ model, the ground-state is exponentially degenerate, while it is unique (modulo a global inversion of all the spins) in the gaussian case. However in higher dimensions where the transition temperature is finite, critical properties near the transition appear to be the same in both cases. In our nonequilibrium studies we considered the Kawasaki dynamics and studied the quantum Hamiltonian corresponding to the Fokker-Planck operator for the stochastic process. The spectrum of the Hamiltonian is obtained by numerical diagonalization of finite chains. An interesting crossover of the first excited state from momentum $`\pi `$ to $`2\pi /L`$ is observed with increase in system size. We have found analytic upper bounds on the gaps at these two momenta. These, along with our numerical diagonalization results suggest that the gap vanishes as $`1/L^2`$ and so the dynamics is diffusive as in SEP. We have also compared the symmetry properties of our Hamiltonian with the Heisenberg model. We find that while the model has the symmetry of the $`XXZ`$ model, its ground-state properties are closer to those of the ferromagnetic isotropic point. In summary we have shown that our model is a very nontrivial cousin of the Heisenberg ferromagnet. The exclusion of three adjacent like spins essentially changes the model dynamics, and results in a nontrivially depleted condensate in $`<\sigma _0^x\sigma _r^x>`$ and a nontrivial gapped $`<\sigma _0^z\sigma _r^z>`$ correlation function. The existence of a groundstate in every $`S^z`$ sector is quite obvious from the stochastic point of view but a nontrivial one within the framework of the quantum system (e.g. the absence of a $`\sigma ^{}`$ operator), and require a deeper understanding. ## V Acknowledgements We thank Mustansir Barma, Dibyendu Das and Deepak Dhar for useful discussions. AD is grateful to the Poornaprajna Institute for financial support.
warning/0004/astro-ph0004305.html
ar5iv
text
# Probing the Properties of Neutron Stars with Type-I X-Ray Bursts ## 1 Introduction If material accretes onto the surface of a neutron star sufficiently slowly, a layer of fuel develops and then suddenly ignites producing a burst of x-rays known as a Type I X-ray burst. These thermonuclear flashes each release about $`10^{39}`$ ergs and repeat on a timescale from hours to days (Lewin, van Paradijs & Taam 1995 ; Bildsten 1998 ). Although only one source (SAX J1808.4-3658) exhibits periodic variation in its quiescent emission (Wijnands & van der Klis 1998 ; Chakrabarty & Morgan 1998 ), during the bursts themselves the emission is quasiperiodic (e.g. Strohmayer, Zhang & Swank 1997 ), apparently due to rotational modulation of the inhomogeneities in the thermonuclear burning. Strohmayer & Markwardt 1999 find that the oscillations in the cooling tails of the X-ray bursts from 4U 1702-429 and 4U 1728-34 are nearly coherent with $`Q4000`$ and consistent with an increase in the modulation frequency as the burning layer cools. Joss 1978 calculated the first numerical models of thermonuclear burning on the surface of a 1.4 M neutron star with a radius of 6.6 km. He found that the initial surface layer of helium expands from a thickness of about three meters to one of thirty meters during the onset of the burst (also Bildsten 1995 ). Strohmayer et al. 1997 argue that this increase in radius is sufficient to account for the frequency shifts observed during the bursts due to the conservation of angular momentum. To use the conservation of angular momentum to explain the change in frequency of the emission as the burning region of the atmosphere (or hotspot) expands and then contracts, one assumes that the specific angular momentum of a fluid element is conserved during the burst. The magnetic fields of the neutron stars are likely to be weak (otherwise there would be periodicities in the persistent emission); consequently, this is a viable assumption. However, the relationship between the specific angular momentum of a fluid element and its position is more complicated in general relativity than in Newtonian theory (e.g. Abramowicz, Miller & Stuchlik 1993 ). In the case of the Schwarzschild spacetime, the angular momentum of a fluid element is $$\mathrm{d}L=\mathrm{d}m\mathrm{\Omega }r_g^2=\mathrm{d}m\mathrm{\Omega }\frac{r^2}{12M/r}$$ (1) where $`\mathrm{\Omega }`$ is the angular velocity of the element as measured by an observer at infinity. This expression and its generalizations result in the reversal of the centrifugal forces (e.g. Abramowicz & Prasanna 1990 ; Abramowicz 1993 ) and gyroscopic precession (e.g. Nayak & Vishveshwara 1998 ) in curved spacetimes. Rather than using Equation 1 to obtain the change in the angular velocity of the hotspot, § 2 calculates the change the angular velocity in a frame that rotates with the neutron star. The reasons for this treatment are two-fold. First, the calculation is quite straightforward and illustrative. Second, the effects of frame dragging are likely to be important for neutron stars, so rather than deriving a new definition for the radius of gyration to include frame dragging, frame dragging may simply be included in the metric. § 3 applies these results to Type I X-ray bursts, and finally § 4 summarizes the conclusions. ## 2 The Coriolis Force in Curved Spacetime Evaluating the Coriolis force in the curved spacetime surrounding a neutron star is straightforward once the metric is specified. The forces that affect the motion of the neutron star atmosphere are most easily expressed in a frame that rotates with the star. In this frame the nonzero components of the metric tensor are $`g_{00}`$ $`=`$ $`1{\displaystyle \frac{2M}{r}}+\mathrm{\Omega }^2r^2\mathrm{sin}^2\theta \left({\displaystyle \frac{I}{r^3}}+1\right)`$ (2) $`g_{03}=g_{30}`$ $`=`$ $`\mathrm{\Omega }r^2\mathrm{sin}^2\theta \left(2{\displaystyle \frac{I}{r^3}}+1\right)`$ (3) $`g_{11}`$ $`=`$ $`\left(1{\displaystyle \frac{2M}{r}}\right)^1`$ (4) $`g_{22}`$ $`=`$ $`r^2`$ (5) $`g_{33}`$ $`=`$ $`r^2\mathrm{sin}^2\theta .`$ (6) where $`x^\mu =[t,r,\theta ,\varphi ]`$. $`J=I\mathrm{\Omega }`$ which accounts for the effects of frame dragging to lowest order in $`\mathrm{\Omega }`$ (Landau & Lifshitz 1987 ), and the various quantities are given in geometric units with $`c=G=1`$. For rotational frequencies , $`\nu <400`$ Hz, this is a good approximation (e.g. Morsink & Stella 1999 ) but for $`\nu 600`$ Hz, the induced quadrupole moment of the star may introduce a significant but not dominant correction to the metric (Cook, Shapiro & Teukolsky 1994 ; Sibgatullin & Sunyaev 1998 ); this issue will be addressed in the more detailed sequel. The Coriolis force for material moving radially manifests itself through the connection term, $`\mathrm{\Gamma }_{01}^3`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left(g^{33}g_{30,1}+g^{30}g_{00,1}\right)`$ (7) $`=`$ $`{\displaystyle \frac{\mathrm{\Omega }}{r}}\left[\left(1{\displaystyle \frac{I}{r^3}}\right)+{\displaystyle \frac{1+2I/r^3}{12M/r}}{\displaystyle \frac{M}{r}}\right],`$ (8) and $`\mathrm{\Gamma }_{11}^3=\mathrm{\Gamma }_{00}^3=0`$. The first term yields the standard “Newtonian” Coriolis force reduced by the effects of frame dragging, and the second term arises because the usually afferent gravitational acceleration has an azimuthal component for material moving radially in the rotating frame. The atmosphere expands with a four velocity $`𝒗`$ whose components in the rotating frame are found by setting $`𝒖\mathbf{}𝒖=1`$, $`𝒗\mathbf{}𝒗=1`$ and $`𝒖\mathbf{}𝒗=\gamma `$ where $`𝒖`$ points along the time axis in the rotating frame (i.e. it is four velocity of an observer stationary in the rotating frame). Due to pressure gradients, the trajectory of gas does not obey the $`t`$ and $`r`$ components of the geodesic equations. In the non-rotating frame, these pressure gradients also have a $`\varphi `$ component. However, in the rotating frame, the gas is free to follow a geodesic in the $`\varphi `$ direction, yielding $$\frac{\mathrm{d}v^3}{\mathrm{d}\tau }=2\beta \gamma ^2\frac{\mathrm{\Omega }}{r}\left(1\frac{3M}{r}\frac{I}{r^3}\right)\left(1\frac{2M}{r}\right)^1+𝒪\left(\mathrm{\Omega }^3\right)$$ (9) where $`\beta `$ is the magnitude of the three-velocity measured locally by an observer stationary in the rotating frame (i.e. $`𝒖`$) and $`\gamma =1/\sqrt{1\beta ^2}`$. If the radial motion is nonrelativistic and small relative to the initial radius (as for the expansion of the atmosphere), the result is easily integrated. After transforming the result to the nonrotating (distant observer’s) frame, it is $$\frac{\mathrm{d}\mathrm{ln}\mathrm{\Omega }_{\mathrm{}}}{\mathrm{d}\mathrm{ln}r}=2\left(1\frac{3M}{r}\frac{I}{r^3}\right)\left(1\frac{2M}{r}\right)^1.$$ (10) This result in the absence of frame dragging (i.e. $`I0`$) could also be obtained by using the definition of Abramowicz, Miller & Stuchlik 1993 for the radius of gyration in general relativity. Comparison of this result with Equation 1 yields the following definition for the radius of gyration with frame dragging included to lowest order in $`\mathrm{\Omega }`$, $$r_g=\frac{r}{\sqrt{12M/r}}\left(1\frac{2M}{r}\right)^{\frac{1}{8}\frac{I}{M^3}}\mathrm{exp}\left(\frac{1}{4}\frac{r+M}{M^2r^2}I\right)$$ (11) As the mass of the star and its moment of inertia vanish, the familiar result obtains. However, in the vicinity of a neutron star, the correction may be large. ## 3 Application to X-Ray Bursts Strohmayer & Markwardt 1999 find that during a X-ray burst, the observed frequency of the modulation exponentially decreases toward a value which is constant from burst to burst for each source. They interpret this asymptotic frequency as the spin frequency of the star. At the beginning of the burst, they argue that the atmosphere expands and slows due to the Coriolis force, and as it cools and deflates the spin frequency increases again through the Coriolis force. This simple inverse correlation between the spin frequency and the thickness of the atmosphere restricts $$R>3M+\frac{I}{R^2}$$ (12) for the neutron stars associated with 4U 1636-54, KS 1731-26, Aql X-1, the galactic center source, 4U 1702-43 and 4U 1728-34 ( $`R`$ and $`M`$ are the mass and radius of the neutron star). For a uniform density sphere in the Newtonian limit $`I=\frac{2}{5}MR^2`$. Ravenhall & Pethick 1994 find that the moment of inertia for the FPS equation of state (Friedman & Pandharipande 1981 ) is well fit by $$I=0.21MR^2\left(1\frac{2M}{R}\right)^1.$$ (13) In the context of this equation of state, $`R>3.49M`$. Cumming & Bildsten 2000 have examined the structure of the atmosphere of the X-ray burst in further detail. By understanding the magnitude of the change in the frequency of X-ray bursts as well as its sign, tighter constraints on the radius of the neutron star may be obtained. The the proper thickness of the layer ($`\mathrm{\Delta }z_{1.89}`$) as calculated by Cumming & Bildsten 2000 for $`g_s=1.89\times 10^{14}`$ cms<sup>-2</sup> and the coordinate thickness ($`\mathrm{\Delta }r`$) are related by $$\frac{\mathrm{\Delta }r}{R}=2.1\times 10^7\text{cm}^1\mathrm{\Delta }z_{1.89}\frac{R}{M}\left(1\frac{2M}{R}\right)$$ (14) where the variation in the surface gravity with mass and radius is included and $`2.1\times 10^7\text{cm}^1=g_sc^2`$. The change in observed frequency $`\mathrm{\Delta }\nu `$ for $`\nu =300`$ Hz is $$\mathrm{\Delta }\nu =0.252\text{Hz}\frac{\mathrm{\Delta }z_{1.89}}{20\text{m}}\frac{\nu }{300\text{Hz}}\frac{R}{M}\left(1\frac{3M}{R}\frac{I}{R^3}\right)$$ (15) It is also straightforward to add the relativistic corrections to the more detailed results of Cumming & Bildsten 2000 (C&B) which yields $$\frac{\mathrm{\Delta }\mathrm{\Omega }_{\mathrm{}}}{\mathrm{\Omega }_{\mathrm{}}}|_{\text{GR}}=\frac{R}{M}\left(\frac{R}{M}|_{\text{C\&B}}\right)^1\left(1\frac{3M}{R}\frac{I}{R^3}\right)\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}|_{\text{C\&B}}$$ (16) where the effects of changing surface gravity are also included by assuming that the scale height of the atmosphere is inversely proportional to the gravitational acceleration; Cumming & Bildsten 2000 assume that $`R/M4.82`$. Figure LABEL:fig:correct shows the corrections to the calculated values of $`\mathrm{\Delta }\mathrm{\Omega }`$ including both the effects of general relativity and varying the surface gravity. If the effects of the curved spacetime are included, the result is given by the middle curve, and if frame dragging is also included the lower curve traces the result. The expected change in the observed angular velocity as the atmosphere expands vanishes for $`R=3M`$ (without frame dragging) and $`R=3.49M`$ (with frame dragging). Table 1 of Cumming & Bildsten 2000 presents the properties of the oscillations from Type I bursts. Specifically, the sources fall into two types those with $`\nu 300`$ Hz and those with $`\nu 600`$ Hz. Both sets of objects typically have $`\mathrm{\Delta }\nu 1`$ Hz. Miller 1999 argues that those objects with faster oscillations simply have two burning fronts active simultaneously, so the actual rotational frequency of the star is $`300`$ Hz and the kinematics of the expanding atmosphere is similar to those with faster oscillations. An alternative possibly is that the 600-Hz oscillations correspond to a 600-Hz rotational frequency and that $`\mathrm{\Delta }\nu /\nu `$ is smaller due to a different value of $`M/R`$. To wit, assuming that the 300-Hz objects are 1.4 M neutron stars with the FPS equation of state (Friedman & Pandharipande 1981 ; Ravenhall & Pethick 1994 ) they have a radius of 10.75 km and $`M/R=0.196`$. Using this data to normalize Equation 16 for $`\mathrm{\Delta }\nu /\nu =6\times 10^3`$ gives $$\frac{\mathrm{\Delta }\mathrm{\Omega }}{\mathrm{\Omega }}|_{\text{C\&B}}=13.5\times 10^3$$ (17) which is about twice as large as the value that Cumming & Bildsten 2000 find for their Newtonian models. Using this value, Equation 16 can be used to infer a value of $`M/R`$ for the fast rotators of 0.234, modestly different from that of the slow rotators (0.246 if frame dragging is neglected). Without the general relativistic corrections, $`M/R`$ for the fast rotators would be 0.391 which is more compact than is achieved for any standard equation of state. Using the value of $`M/R`$ which includes frame dragging yields $`M=1.6\text{M}_{}`$ for the fast rotators and $`R=10.3`$ km. If more detailed models argue that the normalization (Equation 17) is unphysically large, the conclusion would be that the objects have $`R3M`$. Specifically, if the modeling of the evolving atmosphere by Cumming & Bildsten 2000 is precisely correct, $`R=8.1M`$ for the slowly rotating bursters and $`R=5.7M`$ for the fast rotators. If the mass of the neutron stars exceeds a solar mass, this requires a very hard equation of state such as the TI and MF results (Pandharipande, Pines & Smith 1976 ). ## 4 Conclusions The Coriolis force in the curved spacetime near a neutron star is substantially smaller than the Newtonian value and reverses its direction as the radius of the star approaches $`3GM/c^2`$. For typical neutron-star parameters of $`R=10`$ km and $`M=1.4`$ M, the fully relativistic Coriolis force is only 30% of the Newtonian value. This contrasts with the assertion of Strohmayer 1999 that the correction to the Coriolis force is the same order as the gravitational redshift, i.e. 20 – 30% due to time dilation alone. Although time dilation does contribute, it does not dominate. Time dilation alone would not predict, the change in the sign of the Coriolis force for $`r3M`$. The bulk of the effect arises from the fact that the specific angular momentum of a fluid element has a minimum at the photon circular orbit, which has $`r=3M`$ for a non-rotating star. This also manifests itself through the reversal of the centrifugal force also at $`r=3M`$ (e.g. Abramowicz & Prasanna 1990 ; Abramowicz 1990 , Abramowicz 1993 ; Sonego & Massar 1996 ) and the fact that compact stars may have moments of inertia which exceed $`\frac{2}{5}MR^2`$, the Newtonian value for a sphere with constant density (e.g. Hartle 1967 ; Chandrasekhar & Miller 1974 ). Because the Coriolis force sensitively depends on the ratio of the mass to the radius of a relativistic star, Type I X-ray bursts provide a strong constraint on the equation of state of the underlying neutron star. Specifically, since the oscillation frequency of the burst increases as the atmosphere cools and presumably shrinks, the radius of the star must exceed $`3GM/c^2`$. If frame dragging is included, the constraint is stronger, $`R>3.49GM/c^2`$. Furthermore, this sensitivity provides a simple explanation for the fact that the quickly rotating bursters, 4U 1636-54, KS 1731-26, Aql X-1 and the galactic center source, spin up by the same amount as the slowly rotating bursters, 4U 1702-43 and 4U 1728-34. The fast rotators are simply slightly more massive, 1.6 M versus 1.4 M. Further studies of the thermonuclear burning in the atmospheres of Type I X-ray bursts will make them a precise probe of the spacetime geometry surrounding rotating neutron stars. General relativity presents gravity as an inertial force; therefore, it is not surprising that Newtonian notions of inertial forces do not apply in strong gravitational fields. Gravity strongly affects the Coriolis force which is important in the evolution of Type I X-Ray bursts; consequently, these bursts provide a strong constraint on the spacetime geometry surrounding accreting neutron stars. I would like to acknowledge a Lee A. DuBridge Postdoctoral Scholarship for support. I would like to thank Lars Bildsten, Greg Ushomirsky, Alessandra Buonanno, Lior Burko, Scott Hughes, Eugene Chiang and Yoram Lithwick for useful discussions.
warning/0004/hep-ph0004164.html
ar5iv
text
# 1 Introduction ## 1 Introduction With the prospect of a new generation of ongoing kaon experiments, a number of rare kaon decays have been suggested to test the Cabibbo-Kobayashi-Maskawa (CKM) paradigm. However it is sometimes a hard task to extract the short-distance contribution, which depends on the CKM matrix, because of large theoretical uncertainties in the long-distance contribution to the decays . To avoid this difficulty, much of recent theoretical work as well as experimental attention has been on searching for the two modes: $`K^+\pi ^+\nu \overline{\nu }`$ and $`K_L\pi ^0\nu \overline{\nu }`$. It is believed that the long-distance contributions in these two modes are much smaller than the short-distance ones and therefore they are negligible . It has been shown that the decay branching ratio of $`K^+\pi ^+\nu \overline{\nu }`$ is close to $`10^{10}`$ arising dominated from the short-distance loop contributions containing virtual charm and top quarks. This decay is a CP conserving process and probably the cleanest one, in the sense of theoretical uncertainties, to study the absolute value of the CKM element $`V_{td}`$. Currently, the E787 group at BNL has seen one event for the decay with the branching ratio of $`B(K^+\pi ^+\nu \overline{\nu })=1.5\begin{array}{c}+3.5\\ 1.3\end{array}\times 10^{10}`$, which is consistent with the standard model prediction, and it is expected that there will be several events when the analysis is complete. The approved experiments of E949 at BNL and E905 at FNAL will have the sensitivities of $`10^{11}`$ and $`10^{12}`$, respectively. On the other hand, the decay $`K_L\pi ^0\nu \overline{\nu }`$ depending on the imaginary part of $`V_{td}`$ is a CP violating process and offers a clear information about the origin of CP violation. In the standard model, it is dominated by the Z-penguin and W-box loop diagrams with virtual top quark and the decay branching ratio is found to be at the level of $`10^{12}`$ , whereas the current experimental limit is less than $`5.9\times 10^7`$ given by the experiment of E799 at FNAL . Several dedicated experimental searches for this decay mode are underway at KEK, BNL and FNAL, respectively. However, from an experimental point of view very challenging efforts are necessary to perform the experiments. This is because all the final state particles are neutral and the only detectable particles are $`2\gamma `$’s from $`\pi ^0`$. As an alternative search, it was proposed to use the decay of $`K_L\pi ^+\pi ^{}\nu \overline{\nu }`$. But, the decay branching ratio is small and the background for $`\pi ^+\pi ^{}`$ is large . In this paper, we study the radiative decay of $`K_L\gamma \nu \overline{\nu }`$, where there is one photon at the final states. The mode has been considered previously in Refs. and it is believed that the decay is short distance dominated . However, the decay branching ratio predicted in Ref. and does not agree with each other. Furthermore, all the discussions were confined in the CP conserving contribution due to the vector part of the structure-dependent amplitudes . The decay branching ratio was found at the levels of $`10^{11}`$ and $`10^{13}`$ in Refs. and , respectively, which are about two orders of magnitude different. To clarify the issue, we will re-examine the decay by using the form factors of $`K\gamma `$ transition calculated directly in the entire physical range of momentum transfer within the light front framework. We will study both CP conserving and violating contributions to the decay branching ratio, respectively. The paper is organized as follows. In Sec. 2, we present the relevant effective Hamiltonian for the radiative decay of $`K_L\gamma \nu \overline{\nu }`$ and study the form factors in the $`K^0\gamma `$ transition within the light front framework. In Sec. 3, we calculate the decay branching ratio. We also compare our result with those in literature . We give our conclusions in Sec. 4. ## 2 Effective Hamiltonian and Form Factors The processes of $`K_L\nu _l\overline{\nu }_l\gamma `$ ($`l=e,\mu ,\tau `$), arise from the box and $`Z`$-penguin diagrams that contribute to $`sd\nu _l\overline{\nu }_l`$ with the photon emitting from the charged particles in the diagrams. The effective Hamiltonian for $`sd\nu \overline{\nu }`$ at the quark level in the standard model is given by $`H_{eff}(sd\nu \overline{\nu })`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}{\displaystyle \frac{\alpha }{2\pi \mathrm{sin}^2\theta _W}}{\displaystyle \underset{l=e,\mu ,\tau }{}}\left(\lambda _cX_{NL}^l+\lambda _tX(x_t)\right)`$ (1) $`\overline{d}\gamma _\mu (1\gamma _5)s\overline{\nu }_l\gamma ^\mu (1\gamma _5)\nu _l,`$ where $`x_t=m_t^2/M_W^2`$, $`\lambda _i=V_{is}^{}V_{id}(i=c,t)`$ represent the products of the CKM matrix elements, and the functions of $`X_{NL}^l`$ and $`X(x_t)`$ correspond to the top and charm contributions in the loops with the next-to-leading logarithmic approximation, respectively, and their expressions can be found in Ref. . In the Wolfenstein parameterization, we have $`Re\lambda _c`$ $`=`$ $`\lambda (1{\displaystyle \frac{\lambda ^2}{2}}),`$ $`Re\lambda _t`$ $`=`$ $`(1{\displaystyle \frac{\lambda ^2}{2}})A^2\lambda ^5(1\rho +{\displaystyle \frac{\lambda ^2}{2}}\rho ),`$ $`Im\lambda _c`$ $`=`$ $`Im\lambda _t=A^2\lambda ^5\eta .`$ (2) For Phenomenological applications, we use $`X(x_t)`$ $`=`$ $`\eta _XX_0(x_t)`$ (3) where $`\eta _X`$ $`=`$ $`0.994,`$ $`X_0(x_t)`$ $`=`$ $`{\displaystyle \frac{x_t}{8}}\left[{\displaystyle \frac{2+x_t}{1x_t}}+{\displaystyle \frac{3x_t6}{(1x_t)^2}}\mathrm{ln}x_t\right],`$ (4) with the $`\overline{MS}`$ definition of the top-quark mass, $`m_t\overline{m}_t(m_t)=(166\pm 5)GeV`$. For the charm sector, from the Table 1 in Ref. , for example, one has $`X_{NL}^{e,\mu }`$ $`=`$ $`11.00\times 10^4,`$ $`X_{NL}^\tau `$ $`=`$ $`7.47\times 10^4,`$ (5) with the central values of the QCD scale $`\mathrm{\Lambda }=\mathrm{\Lambda }_{\overline{MS}}^{(4)}=(325\pm 80)MeV`$ and the charm quark mass $`m_c=\overline{m}_c(m_c)=(1.30\pm 0.05)GeV`$. From the effective Hamiltonian in Eq. (1), we see that to find the decay rate, we have to evaluate the hadronic matrix element: $`<\gamma |J_\mu |K^0>`$, where $`J_\mu =\overline{d}\gamma _\mu (1\gamma _5)s`$. The element can be parameterized as follows: $`<\gamma (q)|\overline{d}\gamma ^\mu \gamma _5s|K^0(p+q)>`$ $`=`$ $`e{\displaystyle \frac{F_A}{M_K}}\left[ϵ^\mu (pq)(ϵ^{}p)q^\mu \right]`$ $`<\gamma (q)|\overline{d}\gamma ^\mu s|K^0(p+q)>`$ $`=`$ $`ie{\displaystyle \frac{F_V}{M_K}}ϵ^{\mu \alpha \beta \gamma }ϵ_\alpha ^{}p_\beta q_\gamma `$ (6) where $`q`$ and $`p+q`$ are photon and $`K`$-meson four momenta, $`F_A`$ and $`F_V`$ are form factors of axial-vector and vector, respectively, and $`ϵ`$ is the photon polarization vector. The form factors of $`F_A`$ and $`F_V`$ in Eq. (6) can be calculated in the light front quark model at the time-like momentum transfers in which the physically accessible kinematic region is $`0p^2p_{\mathrm{max}}^2`$ and they are found to be $`F_A(p^2)`$ $`=`$ $`4M_K{\displaystyle \frac{dx^{}d^2k_{}}{2(2\pi )^3}\mathrm{\Phi }(x,k_{}^2)\frac{1}{1x}}`$ (7) $`\times \left\{{\displaystyle \frac{1}{3}}{\displaystyle \frac{m_s+Bk_{}^2\mathrm{\Theta }}{m_s^2+k_{}^2}}{\displaystyle \frac{2}{3}}{\displaystyle \frac{m_dAk_{}^2\mathrm{\Theta }}{m_d^2+k_{}^2}}\right\},`$ $`F_V(p^2)`$ $`=`$ $`4M_K{\displaystyle \frac{dx^{}d^2k_{}}{2\left(2\pi \right)^3}\mathrm{\Phi }(x,k_{}^2)\frac{1}{1x}}`$ (8) $`\left\{{\displaystyle \frac{1}{3}}{\displaystyle \frac{m_s(1x)(m_sm_d)k_{}^2\mathrm{\Theta }}{m_s^2+k_{}^2}}{\displaystyle \frac{2}{3}}{\displaystyle \frac{m_dx\left(m_sm_d\right)k_{}^2\mathrm{\Theta }}{m_d^2+k_{}^2}}\right\},`$ where $`A`$ $`=`$ $`(12x^{})x(m_sm_d)2x^{}m_d,`$ $`B`$ $`=`$ $`[(12x^{})x1]m_s+(12x^{})(1x)m_d,`$ $`\mathrm{\Phi }(x,k_{}^2)`$ $`=`$ $`N\left({\displaystyle \frac{2x(1x)}{M_0^2(m_dm_s)^2}}\right)^{1/2}\sqrt{{\displaystyle \frac{dk_z}{dx}}}\mathrm{exp}\left({\displaystyle \frac{\stackrel{}{k}^2}{2\omega _K^2}}\right),`$ $`\mathrm{\Theta }`$ $`=`$ $`{\displaystyle \frac{1}{\mathrm{\Phi }(x,k_{}^2)}}{\displaystyle \frac{d\mathrm{\Phi }(x,k_{}^2)}{dk_{}^2}},`$ $`x`$ $`=`$ $`x^{}\left(1{\displaystyle \frac{p^2}{M_K^2}}\right),\stackrel{}{k}=(\stackrel{}{k}_{},\stackrel{}{k}_z),`$ (9) with $`N`$ $`=`$ $`4\left({\displaystyle \frac{\pi }{\omega _K^2}}\right)^{\frac{3}{4}},`$ $`k_z`$ $`=`$ $`\left(x{\displaystyle \frac{1}{2}}\right)M_0+{\displaystyle \frac{m_s^2m_d^2}{2M_0}},`$ $`M_0^2`$ $`=`$ $`{\displaystyle \frac{k_{}^2+m_d^2}{x}}+{\displaystyle \frac{k_{}^2+m_s^2}{1x}},`$ (10) and $`\omega _K`$ being chosen to be $`0.34GeV`$ fixed by the decay constant of $`f_K=160MeV`$. To illustrate the form factors, we input the values of $`m_d=0.3`$, $`m_s=0.4`$, and $`M_K=0.5`$ in $`GeV`$ to integral whole range of $`p^2`$. It is interesting to note that at $`p^2=0`$, we get that $`(F_A(0),F_V(0))=(0.0429,0.0915)`$ comparing with $`(0.0425,0.0945)`$ found in the chiral perturbation theory at the one-loop level . ## 3 Decay Branching Ratios From the effective Hamiltonian for $`K^0\gamma \nu \overline{\nu }`$ in Eq. (1) and the form factors defined in Eq. (6), we can write the amplitude of $`K^0\gamma \nu \overline{\nu }`$ as $`M(K^0\gamma \nu \overline{\nu })`$ $`=`$ $`i{\displaystyle \frac{G_F}{\sqrt{2}}}{\displaystyle \frac{\alpha }{2\pi \mathrm{sin}^2\theta _W}}{\displaystyle \underset{l=e,\mu ,\tau }{}}\left(\lambda _cX_{NL}^l+\lambda _tX(x_t)\right)`$ (11) $`ϵ_\mu ^{}H^{\mu \nu }\overline{u}(p_{\overline{\nu }})\gamma _\nu (1\gamma _5)v(p_\nu ),`$ with $`H_{\mu \nu }`$ $`=`$ $`{\displaystyle \frac{F_A}{M_K}}(p^{}qg_{\mu \nu }+p_\mu ^{}q_\nu )+iϵ_{\mu \nu \alpha \beta }{\displaystyle \frac{F_V}{M_K}}q^\alpha p^{}_{}{}^{}\beta .`$ (12) where $`p^{}`$ is the four momentum of $`K^0`$ and the form factors $`F_{A,V}`$ are given by Eqs. (7) and (8), respectively. Since $`K_LK_2=(K^0\overline{K}^0)/\sqrt{2}`$, we may write $`(K_L\gamma \nu \overline{\nu })`$ $`=`$ $`_{CPC}+M_{CPV}`$ (13) where $`_{CPC}`$ and $`_{CPV}`$ are the amplitudes corresponding to CP conserving and violating contributions, respectively, which are given by $`_{CPC}`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}{\displaystyle \frac{\alpha }{2\pi \mathrm{sin}^2\theta _W}}{\displaystyle \frac{2}{\sqrt{2}}}{\displaystyle \underset{l=e,\mu ,\tau }{}}\left(Re\lambda _cX_{NL}^l+Re\lambda _tX(x_t)\right)`$ (14) $`ϵ^{\mu \nu \alpha \beta }{\displaystyle \frac{F_V}{M_K}}ϵ_\mu ^{}q_\alpha p_{}_{}{}^{}\beta \overline{u}(p_{\overline{\nu }})\gamma _\nu (1\gamma _5)v(p_\nu ),`$ and $`_{CPV}`$ $`=`$ $`{\displaystyle \frac{G_F}{\sqrt{2}}}{\displaystyle \frac{\alpha }{2\pi \mathrm{sin}^2\theta _W}}{\displaystyle \frac{6}{\sqrt{2}}}Im\lambda _tX(x_t)`$ (15) $`{\displaystyle \frac{F_A}{M_K}}ϵ_\mu ^{}(p^{}qg^{\mu \nu }+p^\mu q^\nu )\overline{u}(p_{\overline{\nu }})\gamma _\nu (1\gamma _5)v(p_\nu ).`$ Here we have neglected the imaginary part of $`Im\lambda _c`$ for $`_{CPV}`$. To evaluate the branching ratio, one needs to replace $`p^2`$ into $`(p^{},q)`$. In the physical allowed region of $`K_L\gamma \nu \overline{\nu }`$, one has that $`0p^2M_K^2.`$ (16) In the $`K_L`$ rest frame, the partial decay rate of $`K_L\gamma \nu \overline{\nu }`$ is given by $`d^2\mathrm{\Gamma }={\displaystyle \frac{1}{(2\pi )^3}}{\displaystyle \frac{1}{8M_K}}^2dE_\gamma dE_\nu .`$ (17) where we have used two variables to describe the kinematic of the decay. For convention, we define $`x_\gamma =2E_\gamma /M_K`$ and $`x_\nu =2E_\nu /M_K`$ as the normalized energies of the photon and and neutrino, respectively, and we have the form $`p^2`$ $`=`$ $`M_K^2(1x_\gamma ).`$ (18) The differential decay rate is then given by $`{\displaystyle \frac{d^2\mathrm{\Gamma }}{dx_\gamma dx_\nu }}`$ $`=`$ $`{\displaystyle \frac{M_K}{256\pi ^3}}\left|\right|^2.`$ (19) By integrating the variable $`x_\nu `$, we obtain $`{\displaystyle \frac{d\mathrm{\Gamma }}{dx_\gamma }}`$ $`=`$ $`{\displaystyle \frac{d\mathrm{\Gamma }_{CPC}}{dx_\gamma }}+{\displaystyle \frac{d\mathrm{\Gamma }_{CPV}}{dx_\gamma }},`$ (20) where $`{\displaystyle \frac{d\mathrm{\Gamma }_{CPC}}{dx_\gamma }}`$ $`=`$ $`{\displaystyle \frac{4\alpha }{3}}\left({\displaystyle \frac{G_F\alpha }{16\pi ^2\mathrm{sin}^2\theta _W}}\right)^2{\displaystyle \underset{l=e,\mu ,\tau }{}}\left(Re\lambda _cX_{NL}^l+Re\lambda _tX(x_t)\right)^2`$ (21) $`|F_V|^2x_\gamma ^3(1x_\gamma )M_K^5,`$ and $`{\displaystyle \frac{d\mathrm{\Gamma }_{CPV}}{dx_\gamma }}`$ $`=`$ $`4\alpha \left({\displaystyle \frac{G_F\alpha }{16\pi ^2\mathrm{sin}^2\theta _W}}\right)^2\left(Im\lambda _tX(x_t)F_A\right)^2x_\gamma ^3(1x_\gamma )M_K^5.`$ (22) To illustrate the numerical results, we use $`m_d=0.3GeV`$, $`m_s=0.4GeV`$, $`m_t=166GeV`$, $`m_c=1.30GeV`$, $`M_K=0.5GeV`$, $`\mathrm{\Lambda }=325MeV`$, $`\alpha (M_Z)=1/128`$, $`\mathrm{sin}^2\theta _W=0.23`$, $`\omega =0.34`$, and the CKM parameters of $`\lambda =0.22`$, $`A=0.83`$, $`\rho =0.13`$, and $`\eta =0.34`$. The differential decay branching ratios of $`dB(K_L\gamma \nu \overline{\nu })_{CPC}/dx_\gamma `$ and $`dB(K_L\gamma \nu \overline{\nu })_{CPV}/dx_\gamma `$ as a function of $`x_\gamma =2E_\gamma /M_K`$ are shown in Figs. 1 and 2, respectively. The decay branching ratios are found to be $`B(K_L\gamma \nu \overline{\nu })_{CPC}`$ $`=`$ $`1.0\times 10^{13},`$ (23) $`B(K_L\gamma \nu \overline{\nu })_{CPV}`$ $`=`$ $`1.5\times 10^{15}.`$ (24) From Eqs. (23) and (24), we find that the CP conserving contribution to the decay branching ratio is about a factor of $`67`$ larger than that from CP violating one. It is clear that the numerical values in (23) and (24) depend on the values of the CKM parameters of $`\rho `$ and $`\eta `$, respectively. Nevertheless, one could conclude that a measurement of the decay would determine the real part of $`V_{td}`$. We now compare our numerical result of the CP conserving contribution in Eq. (23) with those in Refs. and . Our value is about two orders of magnitude and a factor 2 smaller than that in and , respectively. The main reason for the former difference is due to a factor 2 was missed in Eq. (28) of Ref. , whereas that for the later one is unclear. It seems that one needs to re-study the approach in Ref. . Finally, we remark that the ratio between the CP conserving and CP violating branching rates agree with that estimated in Ref. . ## 4 Conclusions We have studied the CP conserving and violating contributions to the decay of $`K_L\gamma \nu \overline{\nu }`$ in the standard model. With the form factors for $`K\gamma `$ transitions calculated directly in the entire physical range of momentum transfer within the light front framework, we have shown that the CP conserving part is much larger than that from CP violating one. We have found that the decay branching ratio is at the level of $`10^{13}`$, which could be accessible at a future kaon project such as the KAMI at FNAL . Acknowledgments This work is supported by the National Science Council of the ROC under contract number NSC89-2112-M-007-013. Fig. 1. The differential decay branching ratios $`dB(K_L\gamma \nu \overline{\nu })_{CPC}/dx_\gamma `$ as a function of $`x_\gamma =2E_\gamma /M_K`$. Fig. 2. The differential decay branching ratios $`dB(K_L\gamma \nu \overline{\nu })_{CPV}/dx_\gamma `$ as a function of $`x_\gamma =2E_\gamma /M_K`$.
warning/0004/hep-ph0004139.html
ar5iv
text
# 1 Introduction ## 1 Introduction The ongoing and forthcoming high-statistics B-physics experiments at BaBar, BELLE, HERA-B, the Tevatron, and the LHC experiments ATLAS, CMS and LHCb will probe the flavor sector of the Standard Model (SM) with high precision. These experiments may reveal physics beyond the SM, and a substantial theoretical effort is devoted to calculating the observables that will be tested in various scenarios of new physics. A common feature of all popular weakly-coupled extensions of the SM is an enlarged Higgs sector. In this paper we study the type-II two-Higgs-doublet model (2HDM), which has the same particle content and tree-level Yukawa couplings as the Higgs sector of the Minimal Supersymmetric Model (MSSM). If the ratio $`\mathrm{tan}\beta `$ of the two Higgs vacuum expectation values is large, the Yukawa coupling to $`b`$ quarks is of order one and large effects on $`B`$ decays are possible. Direct searches for the lightest neutral MSSM Higgs particle have begun to constrain the low $`\mathrm{tan}\beta `$ region in the MSSM, because the theoretically predicted mass range increases with $`\mathrm{tan}\beta `$. Hence observables allowing us to study the complementary region of large $`\mathrm{tan}\beta `$ are increasingly interesting. A further theoretical motivation to study the large $`\mathrm{tan}\beta `$ case is SO(10) grand unified theories : they unify the top and bottom Yukawa couplings at high energies, corresponding to $`\mathrm{tan}\beta `$ of order 50. The leptonic decay $`B_d^{}\mathrm{}^+\mathrm{}^{}`$, where $`d^{}=d`$ or $`s`$ and $`\mathrm{}=e`$, $`\mu `$ or $`\tau `$, is especially well suited to the study of an enlarged Higgs sector with large $`\mathrm{tan}\beta `$. In the SM the decay amplitude suffers from a helicity-suppression factor of $`m_{\mathrm{}}/m_b`$, which is absent in the Higgs-mediated contribution. This helicity suppression factor numerically competes with the suppression factor of $`(m_{\mathrm{}}/M_W)\mathrm{tan}\beta `$ stemming from the Higgs Yukawa couplings to the final state leptons, so that one expects the new contributions in the 2HDM to be similar in size to those of the SM. Earlier papers have already addressed the decay $`B_d^{}\mathrm{}^+\mathrm{}^{}`$ in the 2HDM or the full MSSM . Yet the presented results differ analytically and numerically substantially from each other, so that we have decided to perform a new analysis. This paper is organized as follows. In section 2 we give a brief review of the type-II 2HDM. In section 3 we review the SM calculation of the decay $`B_d^{}\mathrm{}^+\mathrm{}^{}`$ and describe our calculation of the relevant 2HDM diagrams. We finish section 3 by combining the results for the 2HDM diagrams and giving compact expressions for the branching fractions. In section 4 we compare our result with previous calculations. In section 5 we present a numerical analysis of our result and estimate the reach of future experiments. We present our conclusions in section 6. Finally the appendix contains a discussion of trilinear Higgs couplings. ## 2 The two-Higgs-doublet model In this paper we study the type-II 2HDM. The model is reviewed in detail in ref. . The 2HDM contains two complex SU(2) doublet scalar fields, $$\mathrm{\Phi }_1=\left(\begin{array}{c}\varphi _1^+\\ \varphi _1^0\end{array}\right),\mathrm{\Phi }_2=\left(\begin{array}{c}\varphi _2^+\\ \varphi _2^0\end{array}\right),$$ (1) which acquire the vacuum expectation values (vevs) $`\varphi _i^0=v_i`$ and break the electroweak symmetry. The Higgs vevs $`v_1`$ and $`v_2`$ are constrained by the $`W`$ boson mass, $`M_W^2=\frac{1}{2}g^2(v_1^2+v_2^2)=\frac{1}{2}g^2v_{SM}^2`$, where $`v_{SM}=174`$ GeV is the SM Higgs vev. Their ratio is parameterized by $`\mathrm{tan}\beta =v_2/v_1`$. Since in this paper we are not interested in CP-violating quantities, we assume CP is conserved by the Higgs sector for simplicity. The mass eigenstates are then given as follows. The charged Higgs states are $`G^+`$ $`=`$ $`\varphi _1^+\mathrm{cos}\beta +\varphi _2^+\mathrm{sin}\beta `$ $`H^+`$ $`=`$ $`\varphi _1^+\mathrm{sin}\beta +\varphi _2^+\mathrm{cos}\beta ,`$ (2) and their hermitian conjugates. The CP-odd states are $`G^0`$ $`=`$ $`\varphi _1^{0,i}\mathrm{cos}\beta +\varphi _2^{0,i}\mathrm{sin}\beta `$ $`A^0`$ $`=`$ $`\varphi _1^{0,i}\mathrm{sin}\beta +\varphi _2^{0,i}\mathrm{cos}\beta ,`$ (3) where we use the notation $`\varphi _i^0=v_i+\frac{1}{\sqrt{2}}(\varphi _i^{0,r}+i\varphi _i^{0,i})`$ for the real and imaginary parts of $`\varphi _i^0`$. The would-be Goldstone bosons $`G^\pm `$ and $`G^0`$ are eaten by the $`W`$ and $`Z`$ bosons. The CP-even states mix by an angle $`\alpha `$ giving two states, $`H^0`$ $`=`$ $`\varphi _1^{0,r}\mathrm{cos}\alpha +\varphi _2^{0,r}\mathrm{sin}\alpha `$ $`h^0`$ $`=`$ $`\varphi _1^{0,r}\mathrm{sin}\alpha +\varphi _2^{0,r}\mathrm{cos}\alpha .`$ (4) In order to avoid large flavor-changing neutral Higgs interactions we require natural flavor conservation . We impose the discrete symmetry $`\mathrm{\Phi }_1\mathrm{\Phi }_1`$, $`\mathrm{\Phi }_2\mathrm{\Phi }_2`$ (which is softly broken by dimension-two terms in the Higgs potential), with the SU(2) singlet fermion fields transforming as $`dd`$, $`uu`$, $`ee`$. These transformation rules define the type-II 2HDM and determine the Higgs-fermion Yukawa couplings. The Yukawa terms in the Lagrangian are: $$_{\mathrm{Yuk}}=Y_d\overline{Q}\mathrm{\Phi }_1dY_u\overline{Q}\mathrm{\Phi }_2^cuY_l\overline{L}\mathrm{\Phi }_1e+\mathrm{h}.\mathrm{c}.$$ (5) where $`\mathrm{\Phi }^c=i\tau _2\mathrm{\Phi }^{}`$. Down-type quarks and charged leptons (up-type quarks) are given mass by their couplings to $`\mathrm{\Phi }_1`$ ($`\mathrm{\Phi }_2`$). Most of the Higgs couplings needed in our calculation are given in ref. . In addition we must consider the trilinear $`H^+H^{}H`$ couplings ($`H=h^0,H^0`$) which were first given in ref. and are discussed in appendix A. ## 3 Effective hamiltonian for $`𝐁\mathrm{}^+\mathrm{}^{}`$ The decay $`B_d^{}\mathrm{}^+\mathrm{}^{}`$ proceeds through loop diagrams and is of fourth order in the weak coupling. In both the SM and 2HDM, the contributions with a top quark in the loop are dominant, so that one may describe the decay at low energies of order $`m_b`$ by a local $`\overline{b}d^{}\overline{\mathrm{}}\mathrm{}`$ coupling via the effective hamiltonian, $$H=\frac{G_F}{\sqrt{2}}\frac{\alpha _{EM}}{2\pi \mathrm{sin}^2\theta _W}\xi _t\left[C_SQ_S+C_PQ_P+C_AQ_A\right].$$ (6) Here $`G_F`$ is the Fermi constant, $`\alpha _{EM}`$ is the electromagnetic fine structure constant and $`\theta _W`$ is the Weinberg angle. The CKM elements are contained in $`\xi _t=V_{tb}^{}V_{td^{}}`$. The operators in (6) are $$Q_S=m_b\overline{b}P_Ld^{}\overline{\mathrm{}}\mathrm{},Q_P=m_b\overline{b}P_Ld^{}\overline{\mathrm{}}\gamma _5\mathrm{},Q_A=\overline{b}\gamma ^\mu P_Ld^{}\overline{\mathrm{}}\gamma _\mu \gamma _5\mathrm{},$$ (7) where $`P_L=(1\gamma _5)/2`$ is the left-handed projection operator. We have neglected the right-handed scalar quark operators because they give contributions only proportional to the $`d^{}`$ mass. The vector leptonic operator $`\overline{\mathrm{}}\gamma ^\mu \mathrm{}`$ does not contribute to $`B_d^{}\mathrm{}^+\mathrm{}^{}`$, because it gives zero when contracted with the $`B_d^{}`$ momentum. Finally, no operators involving $`\sigma _{\mu \nu }=i[\gamma _\mu ,\gamma _\nu ]/2`$ contribute to $`B_d^{}\mathrm{}^+\mathrm{}^{}`$. Because $`m_bM_W,m_t,M_{H^+}`$, there are highly separated mass scales in the decay $`B_d^{}\mathrm{}^+\mathrm{}^{}`$. Short-distance QCD corrections can therefore contain large logarithms like $`\mathrm{log}(m_b/M_W)`$, which must be summed to all orders in perturbation theory with the help of renormalization group techniques. The calculation of the diagrams in the full high-energy theory gives the initial condition for the Wilson coefficients at a high renormalization scale $`\mu `$ on the order of the heavy masses in the loops. The hadronic matrix elements, however, are calculated at a low scale $`\mu =𝒪(m_b)`$ characteristic of the $`B_d^{}`$ decay. The renormalization group evolution down to this low scale requires the solution of the renormalization group equations of the operators $`Q_A`$, $`Q_S`$ and $`Q_P`$. Yet the operator $`Q_A`$ has zero anomalous dimension because it is a ($`VA`$) quark current, which is conserved in the limit of vanishing quark masses. Similarly, the operators $`Q_S`$ and $`Q_P`$ have zero anomalous dimension because the anomalous dimensions of the quark mass $`m_b(\mu )`$ and the (chiral) scalar current $`\overline{b}P_Ld^{}(\mu )`$ add to zero. Hence the renormalization group evolution is trivial: if the bottom quark mass in $`Q_S`$ and $`Q_P`$ is normalized at a low scale $`\mu =𝒪(m_b)`$, then no large logarithms appear in the effective hamiltonian or in the decay rate. In the SM, the dominant contributions to this decay come from the $`W`$ box and $`Z`$ penguin diagrams shown in fig. 1. These diagrams were first calculated in and give a non-negligible contribution only to the Wilson coefficient $`C_A`$. There is no contribution from a photonic penguin because of the photon’s pure vector coupling to leptons. There are also contributions to the Wilson coefficient $`C_S`$ from a SM Higgs penguin and to the Wilson coefficient $`C_P`$ from the would-be neutral Goldstone boson penguin , but these contributions to the amplitude are suppressed by a factor of $`m_b^2/M_W^2`$ relative to the dominant contributions and can be ignored. The SM decay amplitude is then given by the Wilson coefficient $$C_A=2Y(x_t),$$ (8) where $`x_t=m_t^2(m_t)/M_W^2=4.27\pm 0.26`$ and $`m_t`$ is evaluated in the $`\overline{\mathrm{MS}}`$ scheme at $`\mu =m_t`$, giving $`m_t(m_t)=166`$ GeV. The function $`Y(x_t)`$ is given by $`Y(x_t)=Y_0(x_t)+\frac{\alpha _s}{4\pi }Y_1(x_t)`$, where $`Y_0(x_t)`$ gives the leading order (LO) contribution calculated in and $`Y_1(x_t)`$ incorporates the next-to-leading (NLO) QCD corrections and is given in . The NLO corrections increase $`Y(x_t)`$ by about 3%, if $`m_t`$ is normalized at $`\mu =m_t`$. Numerically, $$Y(x_t)=0.997\left[\frac{m_t(m_t)}{166\mathrm{GeV}}\right]^{1.55},$$ (9) where we have parameterized the dependence on the running top quark mass in the $`\overline{\mathrm{MS}}`$ scheme. We limit our consideration to the case of large $`\mathrm{tan}\beta `$, for which the 2HDM contributions to this decay are significant. In the large $`\mathrm{tan}\beta `$ limit, the Wilson coefficients $`C_P`$ and $`C_S`$ receive sizeable contributions from the box diagram involving $`W`$ and $`H^+`$ and the penguins and fermion self-energy diagrams with neutral Higgs boson exchange shown in fig. 2. There are no new contributions to $`C_A`$ in the 2HDM, which therefore retains its SM value. We have calculated the individual diagrams explicitly in a general $`R_\xi `$ gauge, keeping only the terms proportional to $`\mathrm{tan}^2\beta `$. Although each diagram that involves a $`W^\pm `$ or $`G^\pm `$ boson is gauge-dependent, their sum is gauge-independent. For compactness, we give the results of the individual diagrams below in the ’t Hooft-Feynman gauge. ### 3.1 Box diagram The box diagram in fig. 2 gives the following contribution to $`C_S`$ and $`C_P`$: $`C_S^{box}`$ $`=`$ $`C_P^{box}={\displaystyle \frac{m_{\mathrm{}}}{2M_W^2}}\mathrm{tan}^2\beta B_+(x_{H^+},x_t).`$ (10) Here $`x_{H^+}=M_{H^+}^2/M_W^2`$ and $`x_t`$ was defined after (8) in terms of $`m_t(m_t)`$. Strictly speaking, in a LO calculation like ours, one is not sensitive to the renormalization scheme and we could equally well use the top quark pole mass in $`x_t`$. However, experience with NLO calculations in the SM shows that with the definition of $`m_t`$ adopted here, higher-order QCD corrections are small in leptonic decays. Finally the loop function $`B_+`$ in (10) reads $`B_+(x,y)`$ $`=`$ $`{\displaystyle \frac{y}{xy}}\left[{\displaystyle \frac{\mathrm{log}y}{y1}}{\displaystyle \frac{\mathrm{log}x}{x1}}\right].`$ (11) $`B_+(x_{H^+},x_t)`$ also contains the contribution from internal up and charm quarks with $`m_c=m_u=0`$ from the implementation of the GIM mechanism. The effect of a nonzero charm quark mass is negligibly small. ### 3.2 Penguins The penguin diagram with $`H^+`$ and $`W^+`$ in the loop (see fig. 2) contributes $`C_S^{peng,1}`$ $`=`$ $`{\displaystyle \frac{m_{\mathrm{}}}{2}}\mathrm{tan}^2\beta P_+(x_{H^+},x_t)\left[{\displaystyle \frac{\mathrm{sin}^2\alpha }{M_{h^0}^2}}+{\displaystyle \frac{\mathrm{cos}^2\alpha }{M_{H^0}^2}}\right],`$ $`C_P^{peng,1}`$ $`=`$ $`{\displaystyle \frac{m_{\mathrm{}}}{2}}\mathrm{tan}^2\beta P_+(x_{H^+},x_t){\displaystyle \frac{1}{M_{A_0}^2}}.`$ (12) Here again all three quark flavors enter the result from the GIM mechanism, and the effect of nonzero charm quark mass is negligible. By contrast, in the penguin diagram involving $`H^+`$ and $`G^+`$ in the loop only the internal top quark contribution is relevant, because the coupling of $`G^+`$ to quarks is proportional to either of the quark masses and we neglect $`m_s`$. This diagram gives $`C_S^{peng,2}`$ $`=`$ $`{\displaystyle \frac{m_{\mathrm{}}}{2}}\mathrm{tan}^2\beta P_+(x_{H^+},x_t)[{\displaystyle \frac{\mathrm{sin}^2\alpha }{M_{h^0}^2}}{\displaystyle \frac{(M_{H^+}^2M_{h_0}^2)}{M_W^2}}`$ $`+{\displaystyle \frac{\mathrm{cos}^2\alpha }{M_{H^0}^2}}{\displaystyle \frac{(M_{H^+}^2M_{H_0}^2)}{M_W^2}}],`$ $`C_P^{peng,2}`$ $`=`$ $`{\displaystyle \frac{m_{\mathrm{}}}{2}}\mathrm{tan}^2\beta P_+(x_{H^+},x_t){\displaystyle \frac{1}{M_{A^0}^2}}\left[{\displaystyle \frac{M_{H^+}^2M_{A^0}^2}{M_W^2}}\right].`$ (13) The results in (12) and (13) involve the loop function $`P_+(x,y)`$ $`=`$ $`{\displaystyle \frac{y}{xy}}\left[{\displaystyle \frac{x\mathrm{log}x}{x1}}{\displaystyle \frac{y\mathrm{log}y}{y1}}\right].`$ (14) ### 3.3 Self-energies Before we write down the result from the diagrams with self-energies in the external quark lines, we discuss how these contributions come into play. A treatment of flavor-changing self-energies has been proposed in and analyzed in some detail in . In these works flavor-changing self-energies have been discussed in a different context, the renormalization of the $`W`$-boson coupling to quarks. In counterterms have been chosen in such a way that the flavor-changing self-energies vanish if either of the involved quarks is on-shell. For the down-type quarks these counterterms form two $`3\times 3`$ matrices in flavor space, one for the left-handed quark fields and one for the right-handed ones, and similarly for the up-type quarks. It was argued in that the anti-hermitian parts of the counterterm matrices for the left-handed fields can be absorbed into a renormalization of the CKM matrix, and the hermitian parts of the matrices can be interpreted as wave function renormalization matrices $`Z_{ij}^L`$ and $`Z_{ij}^R`$ with $`i,j=d,s,b`$ for our case of external down-type quarks. However, it has also been argued that the on-shell scheme of is not gauge invariant. In addition we find that the approach of leads to an inconsistency in our calculation. We cannot cancel the anti-hermitian parts of the self-energies in the external lines with the counterterms for the CKM matrix, because unlike in the case of the $`W`$ coupling renormalization there is no tree-level coupling of a neutral scalar or vector boson to $`\overline{b}d^{}`$ to be renormalized. Hence we cannot absorb the anti-hermitian parts of the flavor-changing self-energy matrices into counterterms and they do contribute to our calculation. The absorption of the hermitian parts into wave function counterterm matrices is optional, because the introduction of wave function counterterms only trivially shuffles self-energy contributions into vertex counterterms. It is most straightforward then to avoid the issue of counterterms altogether by simply calculating the fermion self-energy diagrams as they are shown in fig. 2. This calculation is straightforward because the internal $`b`$ quark line is off-shell and therefore it does not contribute to the 1-particle pole of the $`s`$ quark and needs not be truncated. It is crucial to note that one must start with $`m_sm_b`$ in the diagrams with external self-energies in fig. 2 to properly account for the quark propagator $`1/(m_bm_s)`$, and then take the limit $`m_b,m_s0`$ (except in the $`\mathrm{tan}\beta `$-enhanced Higgs couplings, of course) at the end. FCNC transitions become meaningless for degenerate quark masses, and one obtains an incorrect result if one starts with $`m_s=m_b`$ and regulates the propagator pole with an off-shell momentum $`p`$ with $`p0`$. In this respect the Higgs exchange diagrams in fig. 2 differ from the situation with $`\gamma `$\- or $`Z`$\- exchanges, where both methods give the same correct result. Further we note that for $`m_sm_b`$ one must include flavor-changing self-energies in external lines with a factor of 1 rather than of 1/2 as in the flavor-conserving case. This is due to the fact that flavor-conserving self-energies come from the residue of the one-particle pole, while in our approach flavor-changing self-energies are part of the non-truncated Green’s function. By close inspection of the formulae in one also recovers this “factor of 1 rule” from the expressions for the wave function renormalization matrices derived in . There are two fermion self-energy diagrams that contribute in the 2HDM, one with a would-be Goldstone boson $`G^+`$ and one with the physical charged Higgs $`H^+`$. Their sum is ultraviolet-finite. This is different from the SM case, in which the UV divergence of the $`G^+`$ diagram cancels with the UV divergence of a SM Higgs vertex diagram involving a $`G^+`$ and a top quark in the loop. As in the penguin diagrams involving $`H^+`$ and $`G^+`$ in the loop, only the internal top quark contributions to the self-energy are relevant here, because the coupling of $`H^+`$ or $`G^+`$ to quarks is proportional to either of the quark masses and we neglect $`m_s`$. The self-energy diagrams add the following term to the Wilson coefficients: $`C_S^{self}`$ $`=`$ $`{\displaystyle \frac{m_{\mathrm{}}}{2}}\mathrm{tan}^2\beta \left(x_{H^+}1\right)P_+(x_{H^+},x_t)\left[{\displaystyle \frac{\mathrm{sin}^2\alpha }{M_{h^0}^2}}+{\displaystyle \frac{\mathrm{cos}^2\alpha }{M_{H^0}^2}}\right],`$ $`C_P^{self}`$ $`=`$ $`{\displaystyle \frac{m_{\mathrm{}}}{2}}\mathrm{tan}^2\beta \left(x_{H^+}1\right)P_+(x_{H^+},x_t){\displaystyle \frac{1}{M_{A_0}^2}}.`$ (15) ### 3.4 2HDM Wilson coefficients and branching ratios Adding (10), (12), (13) and (15) we obtain the 2HDM Wilson coefficients in (6): $$C_S=C_P=\frac{m_{\mathrm{}}}{2M_W^2}\mathrm{tan}^2\beta \frac{\mathrm{log}r}{r1},$$ (16) where $`rx_{H^+}/x_t=M_{H^+}^2/m_t^2(m_t)`$. Note that the dependence on the masses of the neutral Higgs bosons from the penguin and fermion self-energy diagrams has dropped out in their sum without invoking any relation between the mixing angle $`\alpha `$ and the Higgs masses. The result depends on only two of the 2HDM parameters: $`\mathrm{tan}\beta `$ and $`M_{H^+}`$. The two hadronic matrix elements involved are related by the field equation of motion $`\mathrm{\hspace{0.17em}0}|\overline{b}\gamma ^\mu \gamma _5d^{}(x)|B_d^{}(P_{B_d^{}})`$ $`=`$ $`if_{B_d^{}}P_{B_d^{}}^\mu e^{iP_{B_d^{}}x}`$ $`\mathrm{\hspace{0.17em}0}|\overline{b}\gamma _5d^{}(x)|B_d^{}(P_{B_d^{}})`$ $`=`$ $`if_{B_d^{}}{\displaystyle \frac{m_{B_d^{}}^2}{m_b+m_d^{}}}e^{iP_{B_d^{}}x}.`$ (17) The resulting decay amplitude is $$\left|𝒜\right|=\frac{G_F}{\sqrt{2}}\frac{\alpha _{EM}m_{B_d^{}}f_{B_d^{}}}{4\pi \mathrm{sin}^2\theta _W}\left|\xi _t\left[\left(m_{B_d^{}}C_S\right)\overline{\mathrm{}}\mathrm{}+\left(m_{B_d^{}}C_P\frac{2m_{\mathrm{}}}{m_{B_d^{}}}C_A\right)\overline{\mathrm{}}\gamma _5\mathrm{}\right]\right|.$$ (18) Here $`f_{B_d^{}}`$ is the $`B_d^{}`$ decay constant, normalized according to $`f_\pi =132`$ MeV. The corresponding branching ratio is $`(B_d^{}\mathrm{}^+\mathrm{}^{})`$ $`=`$ $`{\displaystyle \frac{G_F^2\alpha _{EM}^2}{32\pi ^2\mathrm{sin}^4\theta _W}}{\displaystyle \frac{m_{B_d^{}}^3\tau _{B_d^{}}f_{B_d^{}}^2}{8\pi }}|\xi _t|^2\sqrt{1{\displaystyle \frac{4m_{\mathrm{}}^2}{m_{B_d^{}}^2}}}`$ (19) $`\times \left[\left(1{\displaystyle \frac{4m_{\mathrm{}}^2}{m_{B_d^{}}^2}}\right)m_{B_d^{}}^2C_S^2+\left(m_{B_d^{}}C_P{\displaystyle \frac{2m_{\mathrm{}}}{m_{B_d^{}}}}C_A\right)^2\right],`$ where $`\tau _{B_d^{}}`$ is the $`B_d^{}`$ lifetime. Numerically, the branching fractions are given by $`(B_d\mathrm{}^+\mathrm{}^{})`$ $`=`$ $`(3.0\times 10^7)\left[{\displaystyle \frac{\tau _{B_d}}{1.54\mathrm{ps}}}\right]\left[{\displaystyle \frac{f_{B_d}}{210\mathrm{MeV}}}\right]^2\left[{\displaystyle \frac{|V_{td}|}{0.008}}\right]^2{\displaystyle \frac{m_{\mathrm{}}^2}{m_{B_d}^2}}\sqrt{1{\displaystyle \frac{4m_{\mathrm{}}^2}{m_{B_d}^2}}}`$ $`\times \left[\left(1{\displaystyle \frac{4m_{\mathrm{}}^2}{m_{B_d}^2}}\right)\left({\displaystyle \frac{m_{B_d}^2\mathrm{tan}^2\beta }{8M_W^2}}{\displaystyle \frac{\mathrm{log}r}{r1}}\right)^2+\left({\displaystyle \frac{m_{B_d}^2\mathrm{tan}^2\beta }{8M_W^2}}{\displaystyle \frac{\mathrm{log}r}{r1}}Y(x_t)\right)^2\right],`$ $`(B_s\mathrm{}^+\mathrm{}^{})`$ $`=`$ $`(1.1\times 10^5)\left[{\displaystyle \frac{\tau _{B_s}}{1.54\mathrm{ps}}}\right]\left[{\displaystyle \frac{f_{B_s}}{245\mathrm{MeV}}}\right]^2\left[{\displaystyle \frac{|V_{ts}|}{0.040}}\right]^2{\displaystyle \frac{m_{\mathrm{}}^2}{m_{B_s}^2}}\sqrt{1{\displaystyle \frac{4m_{\mathrm{}}^2}{m_{B_s}^2}}}`$ (20) $`\times \left[\left(1{\displaystyle \frac{4m_{\mathrm{}}^2}{m_{B_s}^2}}\right)\left({\displaystyle \frac{m_{B_s}^2\mathrm{tan}^2\beta }{8M_W^2}}{\displaystyle \frac{\mathrm{log}r}{r1}}\right)^2+\left({\displaystyle \frac{m_{B_s}^2\mathrm{tan}^2\beta }{8M_W^2}}{\displaystyle \frac{\mathrm{log}r}{r1}}Y(x_t)\right)^2\right].`$ It is a well known property of the 2HDM that there exists a limit in which the particles $`A^0`$, $`H^0`$, and $`H^+`$ become very heavy and decouple from processes at the electroweak energy scale while $`h^0`$ remains light and its couplings approach those of the SM Higgs particle . In the limit of large $`M_{H^+}`$, $`C_P`$ and $`C_S`$ fall as $`M_{H^+}^2`$. Thus the deviation of the branching fractions from their SM prediction falls as $`M_{H^+}^2`$ in the large $`M_{H^+}`$ limit, and the effects of the enlarged Higgs sector decouple. Next we discuss the accuracy of the large $`\mathrm{tan}\beta `$ approximation. Subleading terms in $`\mathrm{tan}\beta `$ could be enhanced by powers of $`m_t/m_b`$ compared to our result in (16), as is the case for the SM contribution. Such terms indeed occur, but they are suppressed by two powers of $`\mathrm{cot}\beta `$ compared to the SM terms in (16). Hence the formulae above are sufficient for all purposes; e.g. if $`\mathrm{tan}\beta =50`$ the terms subleading in $`\mathrm{tan}\beta `$ give a correction only of $`𝒪(2\%)`$. If $`\mathrm{tan}\beta `$ is between a few and 15 the 2HDM corrections are small and experimentally hard to resolve, so that an error of order $`\mathrm{cot}\beta `$ is tolerable as well. ## 4 Comparison with other calculations ### 4.1 The analyses of He et al. and of Savage In the paper of He, Nguyen and Volkas , the decays $`B\mathrm{}^+\mathrm{}^{}`$, $`BK\mathrm{}^+\mathrm{}^{}`$ and $`bs\mathrm{}^+\mathrm{}^{}`$ are analyzed in both type-I and type-II 2HDMs. In , the only diagrams considered are the box diagram involving two charged Higgs bosons and the $`A^0`$ penguin involving $`H^+`$ and $`W^+`$ in the loop. Although the calculations of are performed in the ’t Hooft-Feynman gauge, the $`A^0`$ penguin involving $`H^+`$ and $`G^+`$ in the loop is not considered. Similarly, in the paper of Savage , the decay $`B\mu ^+\mu ^{}`$ is considered in the 2HDM, with and without tree-level flavor-changing neutral Higgs couplings. Only the contribution of the $`A^0`$ penguin is considered. In both and , several diagrams that are important at large $`\mathrm{tan}\beta `$ and required in order to obtain a gauge-independent result are neglected. ### 4.2 The analysis of Skiba and Kalinowski In the paper of Skiba and Kalinowski , the decay $`B_s\tau ^+\tau ^{}`$ is analyzed in the type-II 2HDM. In additional penguin diagrams are considered that are not proportional to $`\mathrm{tan}^2\beta `$, but rather contain one or no powers of $`\mathrm{tan}\beta `$. We have neglected these contributions in our analysis, because they are not relevant for the interesting case of large $`\mathrm{tan}\beta `$. These additional penguin diagrams can be important for small values of $`\mathrm{tan}\beta `$ and in regions of the parameter space where some of the Higgs quartic couplings are very large resulting in large trilinear $`H^+H^{}H`$ couplings ($`H=h^0,H^0,A^0`$). Considering only terms proportional to $`\mathrm{tan}^2\beta `$, our results for the individual diagrams agree with those of , with two important exceptions. First, the authors of incorrectly conclude that the box diagram involving $`H^+`$ and $`W^+`$ is negligible and therefore neglect it. If we neglect the box diagram, we find that the sum of the remaining contributions proportional to $`\mathrm{tan}^2\beta `$ is gauge-dependent. In the ’t Hooft-Feynman gauge employed in the omitted diagram gives the dominant contribution, affecting the numerical result substantially. Second, our result for the $`A^0`$ penguin diagram involving $`H^+`$ and $`G^+`$ in the loop differs from that of by a sign. Our sign is required for the gauge-independence of $`C_P`$. ### 4.3 The analyses of Huang et al. and Choudhury et al. In the paper of Dai, Huang and Huang , the Wilson coefficients in (6) are calculated in the type-II 2HDM at large $`\mathrm{tan}\beta `$. As in our calculation, only the diagrams proportional to $`\mathrm{tan}^2\beta `$ are considered. However, the authors of consider only the penguin and fermion self-energy diagrams with neutral Higgs boson exchange and neglect the box diagram with a $`W`$ and charged Higgs boson in the loop. Still, after leaving out the box diagram, our results for $`C_S`$ and $`C_P`$ in the ’t Hooft-Feynman gauge do not agree with those of . This is partly due to a typographical error in , which is corrected in . There are two remaining discrepancies. First, our result for the $`A^0`$ penguin diagram involving $`H^+`$ and $`W^+`$ in the loop differs from that of by a sign. Second, in a contribution from the $`h^0`$ and $`H^0`$ penguin diagrams with two $`H^+`$ bosons in the loop is included. This diagram is included in because it apparently receives a factor of $`\mathrm{tan}\beta `$ from the trilinear $`H^+H^{}H`$ couplings ($`H=h^0,H^0`$). We argue in appendix A that the trilinear couplings should not be considered $`\mathrm{tan}\beta `$ enhanced. Therefore we conclude that the penguin diagram with two $`H^+`$ bosons in the loop should not be included in the $`𝒪(\mathrm{tan}^2\beta )`$ calculation because it is of subleading order in $`\mathrm{tan}\beta `$. In the Wilson coefficients in (6) are calculated for supersymmetric models with large $`\mathrm{tan}\beta `$. If the diagrams involving supersymmetric particles are neglected, this calculation reduces to that for the type-II 2HDM with parameters constrained by supersymmetric relations. Again, in only the diagrams with neutral Higgs boson exchange are considered, and the box diagram with a charged Higgs boson and $`W`$ boson is not included. Leaving out the box diagram, our result for $`C_S`$ and $`C_P`$ in the ’t Hooft-Feynman gauge does not agree with the non-SUSY part of that of . This discrepancy arises because our result for the penguin diagrams involving $`H^+`$ and $`W^+`$ in the loop differs from that of by a sign. Once SUSY relations are imposed on the Higgs sector, it is clear that the penguin diagrams with two $`H^+`$ bosons in the loop are not of order $`\mathrm{tan}^2\beta `$, and the authors of have omitted these diagrams, as we did. A final critical remark concerns the treatment of the renormalization group in the paper by Choudhury and Gaur . They include an additional renormalization group factor to account for the running of the Wilson coefficients. Yet these authors have overlooked that the running of the (chiral) scalar quark current in $`Q_S`$ and $`Q_P`$ (see (7)) is compensated by the running of the $`b`$-quark mass multiplying the currents as explained in sect. 3. This leads to an underestimate of the Wilson coefficients by roughly 23%. In conclusion, the papers in , and disagree with each other, and our calculation does not agree with any of them. None of the results in passes the test of gauge-independence and, in our opinion, each contains mistakes. ## 5 Phenomenology As can be seen from the numerical coefficients in (20), $`(B_s\mathrm{}^+\mathrm{}^{})`$ is significantly larger than the corresponding branching fraction for $`B_d`$. This is due primarily to the relative sizes of $`|V_{ts}|`$ and $`|V_{td}|`$. As a result, even though the production rate of $`B_s`$ is three times smaller than that of $`B_d`$ at the Tevatron, the bounds on the leptonic branching fractions of $`B_s`$ are much closer to the SM predictions than those of $`B_d`$ . For this reason we concentrate on the decays of $`B_s`$. Because of the suppression of the branching fractions by $`m_{\mathrm{}}^2/m_B^2`$, clearly the decay to $`\tau `$ pairs is the largest of the leptonic branching fractions in both the SM and the 2HDM. However, this decay is very difficult to reconstruct experimentally (due to the two missing neutrinos), and as a result the experimental limits on $`B`$ decays to $`\tau `$ pairs are very weak. Therefore in our numerical analysis we focus on the decay $`B_s\mu ^+\mu ^{}`$, for which the experimental bound is the closest to the SM prediction. The best experimental bound comes from CDF , where one candidate event for $`B\mu ^+\mu ^{}`$ has been reported; this event was consistent with the expected background and lay in the overlapping part of the search windows for $`B_d`$ and $`B_s`$. The corresponding 95% confidence level upper bound on the $`B_s\mu ^+\mu ^{}`$ branching fraction is $$(B_s\mu ^+\mu ^{})<2.6\times 10^6(\mathrm{expt}).$$ (21) The SM prediction for the branching fraction is $$(B_s\mu ^+\mu ^{})=4.1\times 10^9(\mathrm{SM}),$$ (22) where we have taken the central values for all inputs in (20) and ignored the 2HDM contributions as well as the errors in the hadronic parameters. Because the 2HDM Wilson coefficients in (16) depend on only two of the parameters of the 2HDM, $`\mathrm{tan}\beta `$ and $`M_{H^+}`$, the behavior of the result in different parts of parameter space is easy to understand. In regions of the parameter space with a large 2HDM contribution to $`(B_d^{}\mathrm{}^+\mathrm{}^{})`$ compared to the SM contribution, we may neglect $`Y(x_t)`$ in (20). Then the result is particularly simple: the branching fractions are proportional to $`\mathrm{tan}^4\beta \mathrm{log}^2r/(r1)^2`$. We can see from (20) and the value of $`Y(x_t)`$ given in (9) that the interference between the SM and 2HDM contributions to the branching fractions is destructive. The effect of the destructive interference can clearly be seen in fig. 3. In fig. 3 we plot the predicted value of $`(B_s\mu ^+\mu ^{})`$ in the 2HDM as a function of $`M_{H^+}`$, for various values of $`\mathrm{tan}\beta `$. For comparison we show the constraint on $`M_{H^+}`$ from $`bs\gamma `$ , obtained from the current 95% confidence level experimental upper bound of $`(bs\gamma )<4.5\times 10^4`$ from the CLEO experiment. For very large $`\mathrm{tan}\beta `$ and relatively light $`H^+`$, the 2HDM contribution dominates and the branching fraction is significantly enhanced compared to its SM value. As the 2HDM contribution becomes smaller due to increasing $`M_{H^+}`$ or decreasing $`\mathrm{tan}\beta `$, the branching fraction drops, eventually falling below the SM prediction due to the destructive interference. If we ignore the kinematical factor of $`(14m_{\mathrm{}}^2/m_{B_d^{}}^2)`$ in front of $`C_S`$ in (19) (which is a good approximation for $`B_d^{}\mu ^+\mu ^{}`$ but not for $`B_d^{}\tau ^+\tau ^{}`$) then we can easily show that the branching fraction in the 2HDM crosses the SM value when $$\frac{m_{B_s}^2\mathrm{tan}^2\beta }{8M_W^2}\frac{\mathrm{log}r}{r1}=Y(x_t),$$ (23) and reaches a minimum of half the SM value when $$\frac{m_{B_s}^2\mathrm{tan}^2\beta }{8M_W^2}\frac{\mathrm{log}r}{r1}=\frac{1}{2}Y(x_t).$$ (24) These correspond to $`\mathrm{tan}^2\beta \mathrm{log}r/(r1)=1790`$ and 893, respectively. As a numerical example, taking $`\mathrm{tan}\beta =60`$ and $`M_{H^+}=175`$ GeV (500 GeV), we find $`(B_s\mu ^+\mu ^{})=1.8\times 10^8`$ ($`2.1\times 10^9`$). In fig. 4 we plot the regions of $`M_{H^+}`$ and $`\mathrm{tan}\beta `$ parameter space that will be probed as the sensitivity to the decay $`B_s\mu ^+\mu ^{}`$ improves at the Tevatron Run II. The contours shown (from left to right) were chosen as follows. The current upper bound on $`(B_s\mu ^+\mu ^{})`$ is $`2.6\times 10^6`$ from CDF with about 100 pb<sup>-1</sup> of integrated luminosity. This bound excludes a small region of parameter space with very high $`\mathrm{tan}\beta `$ and very light $`H^+`$, shown by the solid line at the far left of fig. 4. Such low $`H^+`$ masses are already excluded by the constraint from $`bs\gamma `$ . The rest of the contours in fig. 4 show the regions that we expect to be probed at the Tevatron Run II and extended Run II with various amounts of integrated luminosity, assuming that the background for this process remains negligible. In this case the sensitivity to the branching fraction should scale with the luminosity. If there is background however, then the sensitivity will scale only with the square root of the luminosity. With 2 fb<sup>-1</sup> from each of the two detectors, the sensitivity should improve by a factor of 40 over the current sensitivity, to $`6.5\times 10^8`$, shown by the short dashes in fig. 4. For the values of $`\mathrm{tan}\beta `$ that we consider, this sensitivity will still only probe values of $`M_{H^+}`$ already excluded by $`bs\gamma `$. We also show two contours for the expected sensitivity with 10 fb<sup>-1</sup> and 20 fb<sup>-1</sup> of integrated luminosity per detector (dotted and dot-dashed lines in fig. 4, respectively). These correspond to an extended Run II of the Tevatron. With 10 fb<sup>-1</sup> we expect the sensitivity to reach a branching fraction of $`1.3\times 10^8`$, allowing one to begin to probe $`H^+`$ masses above the current bound from $`bs\gamma `$ for $`\mathrm{tan}\beta >54`$. With 20 fb<sup>-1</sup> we expect a reach of $`6.5\times 10^9`$, less than a factor of two above the predicted SM branching fraction. This would allow one to probe $`H^+`$ masses above the current bound from $`bs\gamma `$ for $`\mathrm{tan}\beta >47`$. In particular, for $`\mathrm{tan}\beta =60`$, a non-observation of $`B_s\mu ^+\mu ^{}`$ at this sensitivity would rule out $`H^+`$ lighter than 260 GeV. Looking farther into the future, the experiments at the CERN LHC expect to observe the following numbers of signal (background) events for $`B_s\mu ^+\mu ^{}`$ after three years of running at low luminosity, assuming the SM branching fraction : ATLAS: 27 (93); CMS: 21 (3); and LHCb: 33 (10). Since the suppression of this branching fraction in the 2HDM is at most a factor of two, the LHC experiments will be able to observe this decay for any configuration of the 2HDM at large $`\mathrm{tan}\beta `$. For e.g. $`\mathrm{tan}\beta =60`$ and $`M_{H^+}<285`$ GeV, the branching fraction in the 2HDM is enhanced by 30% or more compared to the SM. Similarly, for $`\mathrm{tan}\beta =60`$ and 375 GeV $`<M_{H^+}<`$ 1 TeV, the branching fraction is suppressed by 30% or more compared to the SM<sup>3</sup><sup>3</sup>3We do not consider charged Higgs masses larger than 1 TeV for naturalness reasons.. In these regions, we expect the LHC to be able to distinguish the 2HDM from the SM. In the region of large $`M_{H^+}`$, the dependence on $`M_{H^+}`$ is very weak; hence the LHC measurement will give powerful constraints on $`\mathrm{tan}\beta `$ in the large $`\mathrm{tan}\beta `$ region. We have made no attempt to simulate the experimental background for this decay in order to obtain an accurate estimate of the reach of the Tevatron Run II. Neither have we taken into account the theoretical uncertainty. We expect the largest theoretical uncertainty to come from uncertainties in the input parameters, primarily the $`B`$ meson decay constants and CKM matrix elements in (20). These uncertainties will be reduced as the $`B`$ physics experiments progress and lattice calculations improve. Also QCD corrections to the 2HDM contribution will arise at NLO and require the calculation of two-loop diagrams. In the SM, the NLO corrections increase the decay amplitude by about 3%, and therefore increase the branching fraction by about 6%. We expect the NLO corrections to the 2HDM contribution to be of the same order, in which case our conclusions are not significantly modified. In order to evaluate the usefulness of $`B_s\mu ^+\mu ^{}`$ as a probe of the 2HDM, we must compare it to other measurements that constrain the 2HDM in the large $`\mathrm{tan}\beta `$ regime. As the statistics of $`B`$ physics experiments improve, the measurement of $`bs\gamma `$ will improve as well. If $`B_s\mu ^+\mu ^{}`$ is to be a useful probe of the 2HDM, it must be sensitive to a range of parameter space not already explored by $`bs\gamma `$ at each integrated luminosity. Fortunately, $`B_s\mu ^+\mu ^{}`$ is complementary to $`bs\gamma `$ because of the different dependence on $`\mathrm{tan}\beta `$. While the 2HDM contributions to $`bs\gamma `$ are independent of $`\mathrm{tan}\beta `$ for $`\mathrm{tan}\beta `$ larger than a few, the 2HDM contributions to $`B_s\mu ^+\mu ^{}`$ depend sensitively on $`\mathrm{tan}\beta `$. This makes $`B_s\mu ^+\mu ^{}`$ an especially sensitive probe of the 2HDM in the large $`\mathrm{tan}\beta `$ regime, while $`bs\gamma `$ will remain more sensitive for moderate and small $`\mathrm{tan}\beta `$. Finally, a fit to the $`Z`$ decay data in the 2HDM puts weak constraints on the $`H^+`$ mass for very large $`\mathrm{tan}\beta `$: $`M_{H^+}>40`$ GeV at 95% confidence level for $`\mathrm{tan}\beta =100`$. The fit gives no constraint for $`\mathrm{tan}\beta <94`$. ## 6 Conclusions In this paper we have analyzed the decays $`B_d^{}\mathrm{}^+\mathrm{}^{}`$ in the type-II 2HDM with large $`\mathrm{tan}\beta `$. Although these decays have been studied in a 2HDM before, the previous analyses omitted the box diagram involving $`W`$ and $`H^+`$, which is the dominant contribution at large $`\mathrm{tan}\beta `$ in the ’t Hooft-Feynman gauge and is needed for gauge independence. We showed that when all the contributions are properly included in the large $`\mathrm{tan}\beta `$ limit, the resulting expressions for the branching fractions are quite simple and depend only on $`\mathrm{tan}\beta `$ and the charged Higgs mass. These 2HDM contributions can enhance or suppress the branching fractions by a significant amount compared to their SM values, providing tantalizing search possibilities with the potential to probe large parts of the large $`\mathrm{tan}\beta `$ parameter space of the 2HDM. We have focused in our numerical analysis on $`B_s\mu ^+\mu ^{}`$, for which the the experimental sensitivity is best. We find that for reasonable values of $`\mathrm{tan}\beta `$ up to 60 and charged Higgs masses above the lower bound set by $`bs\gamma `$, $`(B_s\mu ^+\mu ^{})`$ can be increased by up to a factor of five above its SM expectation or suppressed by up to a factor of two, depending on the charged Higgs mass. Although very high statistics will be needed to observe this decay, it promises to be an experimentally and theoretically clean probe of new Higgs physics. Acknowledgments We are grateful to Jan Kalinowski and Witold Skiba for helpful discussions, and to Jonathan Lewis for discussions on the reach of the Tevatron in Run II. We would also like to thank the conveners of the B-Physics at Run II Workshop at Fermilab for facilitating useful interaction among theorists and experimentalists. Finally we owe a debt of gratitude to Piotr Chankowski and Łucja Sławianowska for pointing out an error in the relative sign between the SM- and 2HDM-induced contributions to the decay rate in an earlier version of this manuscript, and for confirming our result for the 2HDM Wilson coefficients. ## Appendix A Trilinear Higgs couplings The trilinear $`H^+H^{}H`$ couplings ($`H=h^0,H^0`$) for the non-supersymmetric 2HDM were first presented in . These couplings depend strongly on the model considered. For the most general CP-conserving 2HDM with natural flavor conservation, the $`H^+H^{}H`$ couplings are given by $`igQ_H/M_W`$, where $`Q_{h^0}`$ $`=`$ $`v_{SM}^2[\mathrm{sin}(\beta \alpha )(2\lambda _3+\lambda _4)\mathrm{sin}\beta \mathrm{cos}\beta \mathrm{cos}(\alpha +\beta )\lambda _5`$ $`+2\mathrm{sin}\beta \mathrm{cos}\beta (\mathrm{sin}\alpha \mathrm{sin}\beta \lambda _1+\mathrm{cos}\alpha \mathrm{cos}\beta \lambda _2)]`$ $`Q_{H^0}`$ $`=`$ $`v_{SM}^2[\mathrm{cos}(\beta \alpha )(2\lambda _3+\lambda _4)\mathrm{sin}\beta \mathrm{cos}\beta \mathrm{sin}(\alpha +\beta )\lambda _5`$ (25) $`+2\mathrm{sin}\beta \mathrm{cos}\beta (\mathrm{cos}\alpha \mathrm{sin}\beta \lambda _1+\mathrm{sin}\alpha \mathrm{cos}\beta \lambda _2)].`$ The $`H^+H^{}A^0`$ coupling is zero. Here $`v_{\mathrm{SM}}=174`$ GeV is the SM Higgs vev, and the $`\lambda _i`$ are the scalar quartic couplings of the Higgs potential given in . To write these couplings in terms of Higgs masses and mixing angles, one must make an assumption to eliminate one of the independent $`\lambda _i`$. In , formulae are presented for the two cases $`\lambda _1=\lambda _2`$ and $`\lambda _5=\lambda _6`$. The formulae in (25) agree with in these two cases. At large $`\mathrm{tan}\beta `$, (25) reduces to $`Q_{h^0}`$ $``$ $`v_{\mathrm{SM}}^2\mathrm{cos}\alpha (2\lambda _3+\lambda _4)[1+𝒪(\mathrm{cot}\beta )]`$ $`Q_{H^0}`$ $``$ $`v_{\mathrm{SM}}^2\mathrm{sin}\alpha (2\lambda _3+\lambda _4)[1+𝒪(\mathrm{cot}\beta )].`$ (26) These couplings are not enhanced at large $`\mathrm{tan}\beta `$. Considering instead the case $`\lambda _1=\lambda _2`$ and writing the trilinear couplings in terms of Higgs masses and mixing angles, we find at large $`\mathrm{tan}\beta `$, $`Q_{h^0}`$ $``$ $`\frac{1}{2}M_{H^0}^2\mathrm{tan}\beta \mathrm{cos}^2\alpha \mathrm{sin}\alpha [1+𝒪(\mathrm{cot}\beta )]`$ $`Q_{H^0}`$ $``$ $`\frac{1}{2}M_{h^0}^2\mathrm{tan}\beta \mathrm{cos}\alpha \mathrm{sin}^2\alpha [1+𝒪(\mathrm{cot}\beta )].`$ (27) Naively, one would conclude that these couplings are enhanced at large $`\mathrm{tan}\beta `$. This is incorrect because the angle $`\alpha `$ depends on $`\mathrm{tan}\beta `$. At large $`\mathrm{tan}\beta `$ we have $$\mathrm{tan}2\alpha =\frac{2(4\lambda _3+\lambda _5)}{\lambda _54(\lambda _2+\lambda _3)}\mathrm{cot}\beta [1+𝒪(\mathrm{cot}^2\beta )].$$ (28) Thus for generic values of the $`\lambda _i`$, $`\mathrm{sin}\alpha \mathrm{cot}\beta `$, and the $`\mathrm{tan}\beta `$ enhancement in (27) is cancelled.
warning/0004/hep-th0004038.html
ar5iv
text
# 1 Introduction ## 1 Introduction In recent studies of nonperturbative aspects of superstring theory, type IIB superstring is found to provide the simplest setting. However it is difficult to obtain a realistic unified theory in IIB superstring at least perturbatively. Therefore we may expect that an entirely new type of nonperturbative compactification mechanism of IIB superstring exits. On the other hand, a new compactification mechanism which involves branes has been proposed. Since branes naturally appear in superstring theory, such a mechanism is expected to apply for IIB superstring theory. Noncommutative Yang-Mills theory (NCYM) has been obtained by compactifying IIB matrix model on noncommutative tori. We can simply obtain $`\stackrel{~}{d}`$ dimensional NCYM by expanding IIB matrix model around $`\stackrel{~}{d}`$ dimensional noncommuting backgrounds. In IIB matrix model, the dynamical variables are the Hermitian matrices which are interpreted as the space-time coordinates. A $`\stackrel{~}{d}`$ dimensional noncommuting background corresponds to a $`\stackrel{~}{d}`$ dimensional noncommutative space-time. The simplest idea for compactification in IIB matrix model is to postulate that the compactification down to $`\stackrel{~}{d}`$ dimensions is realized by expanding the model around $`\stackrel{~}{d}`$ dimensional backgrounds. In this paper we point out that Newton’s force law may be obtained with four dimensional NCYM with maximal SUSY ($`NCYM_4`$). Our argument is based on its dual supergravity description. We argue that there exists a massless bound state in the effective Hamiltonian of supergravity which gives rise to Newton’s force law a la Randall and Sundrum. Therefore $`NCYM_4`$ may be regarded as a four dimensional compactification of IIB superstring. The remarkable feature is that it compactifies ten dimensional superstring straight down to four dimensions. The compactification of matrix models has been the outstanding problem. We argue that we can obtain four dimensional gauge theory and gravitation with four dimensional noncommutative backgrounds in IIB matrix model. In this sense, we have identified the most satisfactory compactification mechanism of matrix models. It is also possible to obtain $`d+1`$ dimensional NCYM theories by expanding BFSS matrix model around $`d`$ dimensional noncommutative backgrounds in an analogous way. It may be interesting to investigate such theories through supergravity approach. However it is beyond the scope of this paper. In the large $`N`$ expansion of gauge theory, Feynman diagrams can be classified with their world sheet topology. This is a generic feature of matrix valued field theory. On the other hand, string theory is perturbatively defined in terms of field theory on the world sheet. String theory may be nonperturbatively formulated in the large $`N`$ limit of matrix models in view of these remarkable correspondences. IIB matrix model is such a proposal which is a large $`N`$ reduced model of maximally supersymmetric gauge theory: $$S=\frac{1}{g^2}Tr(\frac{1}{4}[A_\mu ,A_\nu ][A^\mu ,A^\nu ]+\frac{1}{2}\overline{\psi }\mathrm{\Gamma }^\mu [A_\mu ,\psi ]).$$ (1.1) Here $`\psi `$ is a ten dimensional Majorana-Weyl spinor field, and $`A_\mu `$ and $`\psi `$ are $`N\times N`$ Hermitian matrices. With vanishing fermionic backgrounds, the equations of motion are: $$[A_\mu ,[A_\mu ,A_\nu ]]=0.$$ (1.2) The following solutions correspond to BPS-saturated backgrounds: $$[A_\mu ,A_\nu ]=cnumberC_{\mu \nu }.$$ (1.3) Since we interpret $`A_\mu `$ as space-time coordinates due to $`𝒩`$=2 SUSY, we expect to obtain $`\stackrel{~}{d}`$ dimensional space-time with $`\stackrel{~}{d}`$ dimensional solutions of this type. We further expect to obtain $`\stackrel{~}{d}`$ dimensional gauge theory. Since matrices form noncommutative but associative algebra, we expect a deep connection to noncommutative geometry . In fact we have obtained NCYM of 16 supercharges with these backgrounds. Ordinary gauge theory appears as the low energy effective theory. Since short open strings correspond to gauge particles, we indeed find another evidence that IIB matrix model can describe infinite numbers of fundamental strings. We have further pointed out that NCYM contains nonlocal degrees of freedom which may be interpreted as long open strings. We have indeed shown that they give rise to gravitational interactions at the one loop level as it is expected in superstring theory. Since NCYM seems to contain the both gauge theory and gravitation, it is very likely that it is equivalent to superstring theory in a particular background. The major issue here is the renormalizability of NCYM . We have shown that the high energy behavior of NCYM is equivalent to large $`N`$ gauge theory by using the bi-local field representation. Although it also exhibits long range interactions which we interpret as gravitation, it is very likely that NCYM exits at least for $`\stackrel{~}{d}4`$. NCYM is often argued to be the low energy limit of string theory with constant $`b_{\mu \nu }`$ field . However long range interactions are found due to the presence of long ‘open strings’ which might signal the presence of ‘closed strings’ . These issues are currently under active investigations . In this paper we propose that NCYM is superstring theory on the von Neumann lattice whose effective string scale $`\alpha _{eff}^{}`$ is set by $`C_{\mu \nu }`$. The organization of this paper is as follows. In section 2, we argue that Newton’s force law is obtained with $`NCYM_4`$. Since it is a nonperturbative problem, we study its dual supergravity description. In section 3, we briefly summarize our formulation of NCYM as twisted reduced models. In section 4, we estimate the string tension of NCYM using the formalism of section 3. We find that our estimate is consistent with string theoretic expectations in section 2. In section 5, we investigate the graviton exchange process by the one loop perturbation theory. We conclude in section 6 with discussions. ## 2 $`NCYM_4`$ as a unified theory In this section, we argue that $`NCYM_4`$ contains four dimensional gauge theory and gravitation. It is clear that NCYM contains ordinary gauge theory since the noncommutative phases become ineffective at tree level in the low energy limit. The remarkable possibility is that it may also contain gravitation. We first observe the long range interaction at the one loop level which is specific in NCYM. It can be interpreted as gravitational interaction in IIB superstring as it is explained in section 5. In string theory, closed string exchanges should be visible at open string one loop level. Therefore this phenomenon is another stringy feature of NCYM. Although we can see the glimpse of closed strings at the one loop level, we need to understand the quantum effects to all orders to investigate the gravitational sector of $`NCYM_4`$. In order to study such a problem, we recall the supergravity solution of $`m`$ coincident D3-branes with the constant NS B field strength $`b`$: $`e^\varphi `$ $`=`$ $`g_{\mathrm{}}{\displaystyle \frac{(1+\frac{g_{\mathrm{}}m\alpha ^2(1+\alpha ^2b^2)}{r^4})}{(1+\frac{g_{\mathrm{}}m\alpha ^2}{r^4})}},`$ $`{\displaystyle \frac{1}{\alpha ^{}}}ds^2`$ $`=`$ $`{\displaystyle \frac{1}{\alpha ^{}}}(1+{\displaystyle \frac{g_{\mathrm{}}m\alpha ^2(1+\alpha ^2b^2)}{r^4}})^{\frac{1}{2}}({\displaystyle \frac{d\stackrel{}{x}^2}{1+\frac{g_{\mathrm{}}m\alpha ^2}{r^4}}}+dr^2+r^2d\mathrm{\Omega }_5^2),`$ $`B_2`$ $`=`$ $`{\displaystyle \frac{b}{(1+\frac{g_{\mathrm{}}m\alpha ^2}{r^4})}}dxdy+{\displaystyle \frac{b}{(1+\frac{g_{\mathrm{}}m\alpha ^2}{r^4})}}dzd\tau ,`$ $`C_2`$ $`=`$ $`i{\displaystyle \frac{1}{g_{\mathrm{}}(1+\alpha ^2b^2)}}B_2,`$ $`C_0`$ $`=`$ $`i{\displaystyle \frac{b^2}{g_{\mathrm{}}(1+\alpha ^2b^2)}}{\displaystyle \frac{1}{(1+\frac{g_{\mathrm{}}m(1+\alpha ^2b^2)}{r^4})}},`$ $`F_{0123r}`$ $`=`$ $`4i{\displaystyle \frac{1}{(1+\frac{g_{\mathrm{}}m\alpha ^2}{r^4})^2}}{\displaystyle \frac{m}{r^5}}.`$ (2.1) Here $`g_{\mathrm{}}`$ is the dilaton expectation value at $`r=\mathrm{}`$. $`\stackrel{}{x}`$ denotes four dimensional space-time coordinates in this section. These background fields appear in the Euclidean IIB supergravity action: $`S_{IIB}`$ $`=`$ $`S_{NS}+S_R+S_{CS},`$ $`S_{NS}`$ $`=`$ $`{\displaystyle \frac{1}{2}}{\displaystyle d^{10}x\sqrt{g}e^{2\varphi }(R+4_\mu \varphi ^\mu \varphi \frac{1}{2}H_{3}^{}{}_{}{}^{2})},`$ $`S_R`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle d^{10}x\sqrt{g}(F_{1}^{}{}_{}{}^{2}+\stackrel{~}{F}_{3}^{}{}_{}{}^{2}+\frac{1}{2}\stackrel{~}{F}_{5}^{}{}_{}{}^{2})},`$ $`S_{CS}`$ $`=`$ $`{\displaystyle \frac{1}{4}}{\displaystyle C_4}H_3F_3,`$ (2.2) where $`\stackrel{~}{F}_3`$ $`=`$ $`F_3C_0H_3,`$ $`\stackrel{~}{F}_5`$ $`=`$ $`F_5{\displaystyle \frac{1}{2}}C_2H_3+{\displaystyle \frac{1}{2}}B_2F_3.`$ (2.3) We identify the $`r`$ dependent metric $`g_{\alpha \beta }`$ in eq.(2.1) as the four dimensional metric for fundamental strings: $$g_{\alpha \beta }=(1+\frac{g_{\mathrm{}}m\alpha ^2(1+\alpha ^2b^2)}{r^4})^{\frac{1}{2}}(\frac{1}{1+\frac{g_{\mathrm{}}m\alpha ^2}{r^4}})\delta _{\alpha \beta }.$$ (2.4) We postulate that D3-branes are located at the maximum of $`g_{\alpha \beta }`$ , namely at the ‘boundary’ $`r=(g_{\mathrm{}}m\alpha ^2)^{1/4}`$. Since open strings live on the D3-branes, we identify the open string metric $`G_{\alpha \beta }`$ with $`g_{\alpha \beta }`$ at the ‘boundary’ as $`G_{\alpha \beta }(1+\alpha ^2b^2)^{\frac{1}{2}}\delta _{\alpha \beta }.`$ (2.5) Eq.(2.1) indicates that fundamental string metric grows at smaller $`r`$. This phenomenon may be interpreted that closed strings become dynamical due to the quantum effects in NCYM. The graviton exchanges we find at the one loop level in section 5 support such an interpretation. We consider the case that the noncommutativity scale $`l_{NC}`$ is much smaller than the string scale $`l_s`$. Let us focus on the physics at the noncommutativity scale by letting $`\alpha ^{}\mathrm{}`$ but keeping $`b,rO(1)`$. Although it might appear to be a strange limit, it is equivalent to consider the standard $`\alpha ^{}ϵ^{1/2},x_\mu ϵ^{1/2}\stackrel{~}{x}_\mu ,bϵ^1`$ limit. The remarkable point is that open string metric is given by eq.(2.5) as $`\alpha ^{}b`$ in this limit. The Polyakov action for fundamental strings becomes in such a limit as $`{\displaystyle \frac{1}{\alpha ^{}}}{\displaystyle d^2zG_{\alpha \beta }x^\alpha \overline{}x^\beta }+{\displaystyle d^2zb_{\alpha \beta }_0x^\alpha _1x^\beta }+\mathrm{}`$ (2.6) $``$ $`b{\displaystyle d^2z(_0x_\alpha _0x^\alpha _1x_\alpha _1x^\alpha )}+{\displaystyle d^2zb_{\alpha \beta }_0x^\alpha _1x^\beta }+\mathrm{}.`$ So the Hamiltonian for open strings behaves like $$cp^2+\underset{k0}{}kn_k,$$ (2.7) where $`c=1/b`$ and $`n_k`$ denote the number operators of the oscillator modes. Here we find that the noncommutativity scale now acts as the effective string scale! Supergravity description of $`U(m)`$ $`NCYM_4`$ may be obtained by considering large $`\alpha ^{}`$ and small $`g_{\mathrm{}}`$ limit while keeping $`\lambda =g_{\mathrm{}}\alpha ^2b^2`$ fixed: $`e^\varphi `$ $`=`$ $`({\displaystyle \frac{m\lambda ^2}{r^4}}){\displaystyle \frac{1}{(1+\frac{m\lambda }{r^4})}},`$ $`{\displaystyle \frac{1}{\alpha ^{}}}ds^2`$ $`=`$ $`({\displaystyle \frac{m\lambda }{r^4}})^{\frac{1}{2}}({\displaystyle \frac{d\stackrel{}{x}^2}{1+\frac{m\lambda }{r^4}}}+dr^2+r^2d\mathrm{\Omega }_5^2),`$ $`B_2`$ $`=`$ $`{\displaystyle \frac{1}{(1+\frac{m\lambda }{r^4})}}dxdy+{\displaystyle \frac{1}{(1+\frac{m\lambda }{r^4})}}dzd\tau ,`$ $`C_2`$ $`=`$ $`i{\displaystyle \frac{1}{\lambda }}B_2,`$ $`C_0`$ $`=`$ $`i{\displaystyle \frac{r^4}{m\lambda ^2}},`$ $`F_{0123r}`$ $`=`$ $`4i{\displaystyle \frac{1}{(1+\frac{m\lambda }{r^4})^2}}{\displaystyle \frac{m}{r^5}}.`$ (2.8) Here we have also put $`b=1`$ which implies that the noncommutativity scale $`l_{NC}`$ is $`O(1)`$. Since we are looking at the vicinity of the D3-branes in this limit, we expect to find massless open strings. However we also find oscillator modes since the effective string scale is set by $`l_{NC}`$ as in eq.(2.7). We recall that there is a crossover at the noncommutativity scale $`l_{NC}`$ in NCYM. When we consider the Wilson loops, we find that the planar diagrams dominate at larger momentum scale than $`1/l_{NC}`$ and the diagrams of all topology contribute in the opposite limit. It may be interpreted that the string coupling (dilaton expectation value) is scale dependent. It is because in our IIB matrix model conjecture, the tree level string theory is considered to be obtained by summing planar diagrams and string perturbation theory is identified with the topological expansion of the matrix model. With this interpretation, the string coupling grows as the relevant momentum scale is decreased while it vanishes in the opposite limit. In the $`D`$ brane interpretation, the small momentum region corresponds to the vicinity of the brane, while the large momentum region corresponds to the region far from the brane since the Higgs expectation value plays the same role with the momentum scale. In this sense $`e^\varphi `$ in eq.(2.8) behaves just like the $`NCYM_4`$ . The small $`r`$ behavior of eq.(2.8) is identical with ordinary $`AdS`$/CFT correspondence if we identify $`\lambda `$ as the coupling of ordinary Yang-Mills theory. This result is reasonable since the low energy limit of $`NCYM_4`$ contains ordinary gauge theory with precisely the same relation between the coupling constants. It is because in string theory the coupling of $`NCYM_4`$ is given by $`\lambda =g_{\mathrm{}}\alpha ^{}b^2`$ when $`\alpha ^{}`$ is large. We may now resort to the standard argument to justify the supergravity description as follows. Since $`m\lambda `$ sets the radius of ‘$`AdS_5`$’ and $`S_5`$, supergravity description is valid in the strong ’t Hooft coupling limit of $`NCYM_4`$. The mass scale for the Kaluza-Klein modes can be estimated to be of order $`1/(m\lambda )^{1/4}l_{NC}`$. We need to consider large $`m`$ limit also in order to keep the dilaton expectation value to be small. As we have argued, the mass scale of the oscillator modes is set by the effective string scale as $`1/l_{NC}`$. In order to investigate the gravitational interaction, we introduce external energy momentum tensor $`T_{\mu \nu }`$. As an explicit example, we may consider the photon-photon scattering on the ‘brane’ as in section 5. Having such a case in mind, we assume that the indices of the nonvanishing components of $`T_{\alpha \beta }`$ run over four dimensional space-time coordinates. We also assume that it is traceless in the four dimensional subspace. There is an ambiguity concerning its dilaton dependence. It may be natural to assume that it contains the factor $`e^\varphi `$ from string theory point of view. However such an ambiguity does not change the main conclusions in this section. We may adopt the coordinate system where the five dimensional subspace $`(\stackrel{}{x},\rho )`$ is conformally flat $$\frac{1}{\alpha ^{}}ds^2=(m\lambda )^{\frac{1}{2}}(A(\rho )(d\stackrel{}{x}^2+d\rho ^2)+d\mathrm{\Omega }_5^2).$$ (2.9) Since $$\rho =𝑑r\sqrt{1+\frac{m\lambda }{r^4}},$$ (2.10) we find that $$A(\rho )1/\rho ^2,\rho \pm \mathrm{}.$$ (2.11) It has the unique maximum at $`\rho =0`$ ($`r=(m\lambda )^{1/4}`$). It is illustrated in Figure 1. Our strategy is to expand the metric and equations of motion around the classical solution to the first order of the fluctuation. In the following investigation, we use the formalism developed in . <sup>5</sup><sup>5</sup>5We use the sign conventions for the curvature tensors of Misner, Thorn and Wheeler in this paper while those of ’t Hooft and Veltman have been used in . Although the signs of the Riemann tensor are the same, those of the Ricci tensor and scalar curvature are the opposite in these conventions. We parametrize the metric as $`g_{\mu \rho }(e^h)_{}^{\rho }{}_{\nu }{}^{}`$ where $`g_{\mu \nu }`$ is the background metric and $`h_{}^{\rho }{}_{\rho }{}^{}=0`$ (traceless). The tensor indices are raised and lowered by the background metric. This formalism explicitly separates $`h_{}^{\mu }{}_{\nu }{}^{}`$ from the conformal mode of the metric. The equation of the motion with respect to $`h_{}^{\mu }{}_{\nu }{}^{}`$ is $$R_{}^{\mu }{}_{\nu }{}^{}+2^\mu _\nu \varphi \frac{1}{2}H_3^\mu H_{3\nu }+\mathrm{}=\kappa ^2T_{}^{\mu }{}_{\nu }{}^{},$$ (2.12) where $`\kappa =e^\varphi `$. We have suppressed the contributions from the R-R sector. The advantage to consider the four dimensional traceless energy momentum tensor $`T_{}^{\mu }{}_{\nu }{}^{}`$ is that all other equations of motion are satisfied to the first order of $`h_{}^{\mu }{}_{\nu }{}^{}`$. In this sense it minimally excites gravitons. Eq. (2.12) is expanded to the first order of $`h_{}^{\mu }{}_{\nu }{}^{}`$ as $`{\displaystyle \frac{1}{2}}^\rho _\rho h_{}^{\mu }{}_{\nu }{}^{}{\displaystyle \frac{1}{2}}h^{\mu \rho },_{\rho \nu }{\displaystyle \frac{1}{2}}h_{\nu \rho },^{\rho \mu }`$ (2.13) $`h_{}^{\mu }{}_{\nu }{}^{},^\rho _\rho \varphi {\displaystyle \frac{1}{2}}H_{}^{\mu }{}_{\rho \tau }{}^{}H_{\nu }^{}{}_{}{}^{\sigma \tau }h_{}^{\rho }{}_{\sigma }{}^{}{\displaystyle \frac{1}{4}}H_{}^{\rho }{}_{\sigma \tau }{}^{}H_{\nu }^{}{}_{}{}^{\sigma \tau }h_{}^{\mu }{}_{\rho }{}^{}+\mathrm{}`$ $`=`$ $`\kappa ^2T_{}^{\mu }{}_{\nu }{}^{}.`$ In the coordinate system of eq.(2.9), it is consistent to assume that the tensor indices of the nonvanishing $`h_{}^{\alpha }{}_{\beta }{}^{}`$ resides in the four dimensional space-time since the tensor indices of $`T_{}^{\alpha }{}_{\beta }{}^{}`$ are also four dimensional. It is also consistent to assume that $`h_{}^{\alpha }{}_{\alpha }{}^{}=0`$ since $`T_{}^{\alpha }{}_{\alpha }{}^{}=0`$. We also adopt the $`h^{\alpha \gamma },_\gamma =0`$ gauge. Our strategy is to first study the following free equation of motion for $`h_{}^{\alpha }{}_{\beta }{}^{}`$ in the ten dimensional curved space-time: $$\frac{1}{2}^\mu _\mu h_{}^{\alpha }{}_{\beta }{}^{}=\kappa ^2T_{}^{\alpha }{}_{\beta }{}^{}.$$ (2.14) Eq.(2.14) can be rewritten as: $`Hh_{}^{\alpha }{}_{\beta }{}^{}`$ $`=`$ $`2A\kappa ^2T_{}^{\alpha }{}_{\beta }{}^{},`$ $`H`$ $`=`$ $`{\displaystyle \frac{A}{\sqrt{g}}}_\mu \sqrt{g}g^{\mu \nu }_\nu .`$ (2.15) The Hamiltonian is $$H=\frac{1}{(m\lambda )^{\frac{1}{2}}}(\stackrel{}{}^2+\frac{^2}{\rho ^2}+\frac{3}{2}\frac{A^{}}{A}\frac{}{\rho }+A\widehat{L}^2),$$ (2.16) where $`A^{}=A/\rho `$. The symbols $`\stackrel{}{}^2`$ and $`\widehat{L}^2`$ denote the Laplacians on $`R^4`$ and $`S^5`$ respectively. We can further simplify eq. (2.15) by the similarity transformation as $`\stackrel{~}{H}A^{\frac{3}{4}}h_{\alpha \beta }`$ $`=`$ $`2\kappa ^2A^{\frac{7}{4}}T_{\alpha \beta },`$ $`\stackrel{~}{H}`$ $`=`$ $`A^{3/4}HA^{3/4}`$ (2.17) $`=`$ $`{\displaystyle \frac{1}{(m\lambda )^{\frac{1}{2}}}}(\stackrel{}{}^2+{\displaystyle \frac{^2}{\rho ^2}}{\displaystyle \frac{3}{4}}{\displaystyle \frac{A^{\prime \prime }}{A}}+{\displaystyle \frac{3}{16}}{\displaystyle \frac{A^2}{A^2}}+A\widehat{L}^2).`$ We concentrate on the $`S`$ wave on $`S^5`$ in what follows. The eigenfunction of $`\stackrel{~}{H}`$ is found to be $`exp(i\stackrel{}{k}\stackrel{}{x})\stackrel{~}{\varphi }_l`$ with the eigenvalue $`(m\lambda )^{\frac{1}{2}}(\stackrel{}{k}^2+E_l)`$. Note that $`E_l`$ acts as the four dimensional mass of the various modes. It can be obtained by solving the following quantum mechanics problem $$(\frac{^2}{\rho ^2}+\frac{3}{4}\frac{A^{\prime \prime }}{A}\frac{3}{16}\frac{A^2}{A^2})\stackrel{~}{\varphi }=E\stackrel{~}{\varphi }.$$ (2.18) Let us introduce a super-charge $$Q=\frac{}{i\rho }\sigma _1+\frac{3}{4}\frac{A^{}}{A}\sigma _2.$$ (2.19) Since the relevant Hamiltonian can be embedded in $`Q^2`$, we only need to solve $`Q\stackrel{~}{\varphi }=0`$ to find the zeromodes. The solution is $$\stackrel{~}{\varphi }=e^{\frac{3}{4}{\scriptscriptstyle 𝑑\rho ({\scriptscriptstyle \frac{A^{}}{A}})}}.$$ (2.20) We conclude that there is a single zero energy bound state with the conformal factor $`A`$ of our type as follows $$\stackrel{~}{\varphi }_0=A^{\frac{3}{4}}.$$ (2.21) Such a zero mode corresponds to a massless field in four dimensions. The propagator in this basis is $$G(x,y)=\underset{n}{}<x|n>\frac{1}{E_n}<n|y>,$$ (2.22) where $`|n>`$ is the eigenstate of $`\stackrel{~}{H}`$ with the eigenvalue $`E_n`$. We may adopt the vacuum saturation type approximation by only considering $`<x|n>=exp(i\stackrel{}{k}\stackrel{}{x})\stackrel{~}{\varphi }_0`$ as the intermediate states. In this way we obtain the propagator of massless fields in four dimensions: $`G(x,y)`$ $``$ $`{\displaystyle d^4kexp(i\stackrel{}{k}(\stackrel{}{x}\stackrel{}{y}))\frac{1}{k^2}\stackrel{~}{\varphi }_0(\rho )\stackrel{~}{\varphi }_0(\rho ^{})}`$ (2.23) $``$ $`{\displaystyle \frac{1}{(\stackrel{}{x}\stackrel{}{y})^2}}\stackrel{~}{\varphi }_0(\rho )\stackrel{~}{\varphi }_0(\rho ^{}).`$ As the final result of these investigations, we find the following gravitational interaction: $`{\displaystyle \frac{1}{2}}{\displaystyle d^{10}xA^{\frac{5}{2}}(\rho )T_{\mu }^{}{}_{}{}^{\nu }(\stackrel{}{x},\rho )h_{}^{\mu }{}_{\nu }{}^{}(\stackrel{}{x},\rho )}`$ (2.24) $`=`$ $`{\displaystyle d^{10}xd^{10}yA^{\frac{7}{4}}(\rho )T_{\alpha }^{}{}_{}{}^{\beta }(\stackrel{}{x},\rho )\stackrel{~}{\varphi }_0(\rho )\frac{1}{(\stackrel{}{x}\stackrel{}{y})^2}\stackrel{~}{\varphi }_0(\rho ^{})\kappa ^2(\rho ^{})T_{}^{\alpha }{}_{\beta }{}^{}(\stackrel{}{y},\rho ^{})A^{\frac{7}{4}}(\rho ^{})}`$ $``$ $`\overline{\kappa }^2{\displaystyle d^4xd^4y\stackrel{~}{T}_{\alpha }^{}{}_{}{}^{\beta }(\stackrel{}{x})\stackrel{~}{T}_{}^{\alpha }{}_{\beta }{}^{}(\stackrel{}{y})\frac{1}{(\stackrel{}{x}\stackrel{}{y})^2}}.`$ We have introduced the four dimensional energy momentum tensor $$\stackrel{~}{T}_{}^{\alpha }{}_{\beta }{}^{}(\stackrel{}{x})=𝑑\rho 𝑑\mathrm{\Omega }_5T_{}^{\alpha }{}_{\beta }{}^{}(\stackrel{}{x},\rho )A^{\frac{5}{2}}(\rho ).$$ (2.25) The interaction between them is of the four dimensional graviton exchange type with the gravitational coupling $`\stackrel{~}{\kappa }=\kappa (0)`$. Here we remark on the Hermiticity of the Hamiltonians in eqs. (2.16) and (2.18). The latter is Hermitian with respect to the trivial norm since $`{\displaystyle 𝑑\rho \stackrel{~}{\varphi }(\frac{}{i\rho }i\frac{3}{4}\frac{A^{}}{A})(\frac{}{i\rho }+i\frac{3}{4}\frac{A^{}}{A})\stackrel{~}{\varphi }}`$ (2.26) $`=`$ $`{\displaystyle 𝑑\rho (\frac{}{i\rho }i\frac{3}{4}\frac{A^{}}{A})\stackrel{~}{\varphi }(\frac{}{i\rho }+i\frac{3}{4}\frac{A^{}}{A})\stackrel{~}{\varphi }}.`$ It is translated into the Hermiticity condition on the former after the similarity transformation as $$𝑑\rho A^{\frac{3}{2}}(\frac{}{i\rho })\varphi (\frac{}{i\rho })\varphi 𝑑\rho \sqrt{g}g^{\rho \rho }\frac{}{\rho }\varphi \frac{}{\rho }\varphi ,$$ where $`\varphi =A^{3/4}\stackrel{~}{\varphi }`$. Therefore our Hamiltonian is positive definite with respect to the natural norm defined by the string frame metric $`g_{\mu \nu }`$. In this sense our important physical input is our identification of the physical metric with string frame metric $`g_{\mu \nu }`$. We have checked that our zero mode remains the exact zero mode after taking account of other terms in eq.(2.13). Therefore we argue that we can obtain four dimensional gravity with $`NCYM_4`$ a la Randall-Sundrum. Since not only the metric but also the dilaton expectation value (string coupling) rapidly decay in the large $`r`$ region, we expect that there is essentially nothing outside the noncommutativity scale transverse to the ‘brane’. In fact we have postulated this kind of ‘compactification’ mechanism in the matrix models . We have expected that four dimensional gravitation is obtained if the eigenvalue distribution of the matrices are four dimensional. It is because the matrices represent space-time coordinates in our proposal. In our interpretation, there is simply nothing outside the support of the eigenvalue distributions, not even space-time. We observe that this Euclidean solution can be analytically continued into Minkowski space-time only in the small $`r`$ region. One possible interpretation of such a solution is to maintain that Minkowski space-time appears from $`NCYM_4`$ as its low energy approximation. We may identify the noncommutativity scale with Planck scale if we apply this model to our space-time. Although the Lorentz invariance is broken at the noncommutativity scale in this model, such a possibility is not excluded by the experiments. Therefore $`NCYM_4`$ is a candidate of the unified theory of interactions. We explain in the subsequent sections that IIB matrix model naturally provides us with such a theory. We still need to solve many problems such as breaking SUSY and finding chiral fermions to construct a realistic unified theory. We hope that these problems can be solved by further investigations in IIB matrix model. ## 3 Noncommutative field theories as twisted reduced models In this section we briefly recapitulate our formulation of NCYM through large $`N`$ reduced models. We have pointed out that well-known twisted reduced models <sup>6</sup><sup>6</sup>6The relevance of reduced models and string theory was first recognized in. are equivalent to NCYM. This connection is further studied in. We consider $`d`$ dimensional $`U(n)`$ gauge theory coupled to adjoint matter as an example: $$S=d^dx\frac{1}{g^2}Tr(\frac{1}{4}[D_\mu ,D_\nu ][D_\mu ,D_\nu ]+\frac{1}{2}\overline{\psi }\mathrm{\Gamma }_\mu [D_\mu ,\psi ]),$$ (3.1) where $`\psi `$ is a Majorana spinor field. The corresponding reduced model is $$S=\frac{1}{g^2}Tr(\frac{1}{4}[A_\mu ,A_\nu ][A_\mu ,A_\nu ]+\frac{1}{2}\overline{\psi }\mathrm{\Gamma }_\mu [A_\mu ,\psi ]).$$ (3.2) Now $`A_\mu `$ and $`\psi `$ are $`n\times n`$ Hermitian matrices and each component of $`\psi `$ is $`d`$-dimensional Majorana-spinor. We expand $`A_\mu =\widehat{p}_\mu +\widehat{a}_\mu `$ around the following classical solution $$[\widehat{p}_\mu ,\widehat{p}_\nu ]=iB_{\mu \nu },$$ (3.3) where $`B_{\mu \nu }`$ are $`c`$-numbers. We assume the rank of $`B_{\mu \nu }`$ to be $`\stackrel{~}{d}`$ and define its inverse $`C^{\mu \nu }`$ in $`\stackrel{~}{d}`$ dimensional subspace. The directions orthogonal to the subspace is called the transverse directions. $`\widehat{p}_\mu `$ satisfy the canonical commutation relations and they span the $`\stackrel{~}{d}`$ dimensional phase space. The semiclassical correspondence shows that the volume of the phase space is $`V_p=n(2\pi )^{\stackrel{~}{d}/2}\sqrt{detB}`$. We Fourier decompose $`\widehat{a}_\mu `$ and $`\widehat{\psi }`$ fields as $`\widehat{a}`$ $`=`$ $`{\displaystyle \underset{k}{}}\stackrel{~}{a}(k)exp(iC^{\mu \nu }k_\mu \widehat{p}_\nu ),`$ $`\widehat{\psi }`$ $`=`$ $`{\displaystyle \underset{k}{}}\stackrel{~}{\psi }(k)exp(iC^{\mu \nu }k_\mu \widehat{p}_\nu ),`$ (3.4) where $`exp(iC^{\mu \nu }k_\mu \widehat{p}_\nu )`$ is the eigenstate of adjoint $`P_\mu =[\widehat{p}_\mu ,]`$ with the eigenvalue $`k_\mu `$. The Hermiticity requires that $`\stackrel{~}{a}^{}(k)=\stackrel{~}{a}(k)`$ and $`\stackrel{~}{\psi }^{}(k)=\stackrel{~}{\psi }(k)`$. We can construct a map from a matrix to a function as $$\widehat{a}a(x)=\underset{k}{}\stackrel{~}{a}(k)exp(ikx),$$ (3.5) where $`kx=k_\mu x^\mu `$. By this construction, we obtain the $``$ product $`\widehat{a}\widehat{b}`$ $``$ $`a(x)b(x),`$ $`a(x)b(x)`$ $``$ $`exp({\displaystyle \frac{C^{\mu \nu }}{2i}}{\displaystyle \frac{^2}{\xi ^\mu \eta ^\nu }})a(x+\xi )b(x+\eta )|_{\xi =\eta =0}.`$ (3.6) The operation $`Tr`$ over matrices can be exactly mapped onto the integration over functions as $$Tr[\widehat{a}]=\sqrt{detB}(\frac{1}{2\pi })^{\frac{\stackrel{~}{d}}{2}}d^{\stackrel{~}{d}}xa(x).$$ (3.7) The twisted reduced model can be shown to be equivalent to NCYM by the the following map from matrices onto functions $`\widehat{a}`$ $``$ $`a(x),`$ $`\widehat{a}\widehat{b}`$ $``$ $`a(x)b(x),`$ $`Tr`$ $``$ $`\sqrt{detB}({\displaystyle \frac{1}{2\pi }})^{\frac{\stackrel{~}{d}}{2}}{\displaystyle d^{\stackrel{~}{d}}x}.`$ (3.8) The following commutator is mapped to the covariant derivative: $$[\widehat{p}_\mu +\widehat{a}_\mu ,\widehat{o}]\frac{1}{i}_\mu o(x)+a_\mu (x)o(x)o(x)a_\mu (x)[D_\mu ,o(x)],$$ (3.9) We may interpret the newly emerged coordinate space as the semiclassical limit of $`\widehat{x}^\mu =C^{\mu \nu }\widehat{p}_\nu `$. Therefore we can interpret $`A_\mu `$ as momenta as well in IIB matrix model with noncommutative backgrounds since $`\widehat{x}`$ and $`\widehat{p}`$ are linearly related. It is the reflection of the remarkable T-duality property of the theory. The space-time translation is realized by the following unitary operator: $$exp(i\widehat{p}d)\widehat{x}^\mu exp(i\widehat{p}d)=\widehat{x}^\mu +d^\mu .$$ Applying the rule eq.(3.8), the bosonic action becomes $`{\displaystyle \frac{1}{4g^2}}Tr[A_\mu ,A_\nu ][A_\mu ,A_\nu ]`$ (3.10) $`=`$ $`{\displaystyle \frac{\stackrel{~}{d}nB^2}{4g^2}}\sqrt{detB}({\displaystyle \frac{1}{2\pi }})^{\frac{\stackrel{~}{d}}{2}}{\displaystyle }d^{\stackrel{~}{d}}x{\displaystyle \frac{1}{g^2}}({\displaystyle \frac{1}{4}}[D_\alpha ,D_\beta ][D_\alpha ,D_\beta ]`$ $`+{\displaystyle \frac{1}{2}}[D_\alpha ,\phi _\nu ][D_\alpha ,\phi _\nu ]+{\displaystyle \frac{1}{4}}[\phi _\nu ,\phi _\rho ][\phi _\nu ,\phi _\rho ])_{}.`$ In this expression, the indices $`\alpha ,\beta `$ run over $`\stackrel{~}{d}`$ dimensional world volume directions and $`\nu ,\rho `$ over the transverse directions. We have replaced $`a_\nu \phi _\nu `$ in the transverse directions. Inside $`()_{}`$, the products should be understood as $``$ products and hence commutators do not vanish. The fermionic action becomes $`{\displaystyle \frac{1}{g^2}}Tr\overline{\psi }\mathrm{\Gamma }_\mu [A_\mu ,\psi ]`$ (3.11) $`=`$ $`\sqrt{detB}({\displaystyle \frac{1}{2\pi }})^{\frac{\stackrel{~}{d}}{2}}{\displaystyle d^{\stackrel{~}{d}}x\frac{1}{g^2}(\overline{\psi }\mathrm{\Gamma }_\alpha [D_\alpha ,\psi ]+\overline{\psi }\mathrm{\Gamma }_\nu [\phi _\nu ,\psi ])_{}}.`$ We therefore find noncommutative U(1) gauge theory. In order to obtain NCYM with $`U(m)`$ gauge group, we need to consider new classical solutions which are obtained by replacing each element of $`\widehat{p}_\mu `$ by the $`m\times m`$ unit matrix: $$\widehat{p}_\mu \widehat{p}_\mu \mathrm{𝟏}_m.$$ (3.12) The Hermitian models are invariant under the unitary transformation: $`A_\mu UA_\mu U^{},\psi U\psi U^{}`$. As we shall see, the gauge symmetry can be embedded in the $`U(N)`$ symmetry. We expand $`U=exp(i\widehat{\lambda })`$ and parameterize $$\widehat{\lambda }=\underset{k}{}\stackrel{~}{\lambda }(k)exp(ik\widehat{x}).$$ (3.13) Under the infinitesimal gauge transformation, we find the fluctuations around the fixed background transform as $`\widehat{a}_\mu `$ $``$ $`\widehat{a}_\mu +i[\widehat{p}_\mu ,\widehat{\lambda }]i[\widehat{a}_\mu ,\widehat{\lambda }],`$ $`\widehat{\psi }`$ $``$ $`\widehat{\psi }i[\widehat{\psi },\widehat{\lambda }].`$ (3.14) We can map these transformations onto the gauge transformation in NCYM by our rule eq.(3.8): $`a_\alpha (x)a_\alpha (x)+{\displaystyle \frac{}{x^\alpha }}\lambda (x)ia_\alpha (x)\lambda (x)+i\lambda (x)a_\alpha (x),`$ $`\phi _\nu (x)\phi _\nu (x)i\phi _\nu (x)\lambda (x)+i\lambda (x)\phi _\nu (x),`$ $`\psi (x)\psi (x)i\psi (x)\lambda (x)+i\lambda (x)\psi (x).`$ (3.15) We have introduced another representation of matrices. For simplicity we consider the two dimensional case first: $$[\widehat{x},\widehat{y}]=iC.$$ (3.16) This commutation relation is realized by the guiding center coordinates of the two dimensional system of electrons in magnetic field. We recall that we have $`n`$ quanta with $`n`$ dimensional matrices. Each quantum occupies the space-time volume of $`2\pi C`$. We may consider a square von Neumann lattice with the lattice spacing $`l_{NC}`$ where $`l_{NC}^2=2\pi C`$. This spacing $`l_{NC}`$ gives the noncommutative scale. Let us denote the most localized state centered at the origin by $`|0`$. It is annihilated by the operator $`\widehat{x}^{}=\widehat{x}i\widehat{y}`$. We construct states localized around each lattice site by utilizing translation operators $`|x_i=exp(ix_i\widehat{p})|0`$. They are the coherent states on a von Neumann lattice $`𝐱_i=l_{NC}(n_i𝐞^x+m_i𝐞^y)`$ where $`n,m𝐙.`$ The generalizations to arbitrary even $`\stackrel{~}{d}`$ dimensions are straightforward. We evaluate the following matrix elements $$\rho _{ij}x_i|x_j=exp(\frac{i}{2}B_{\mu \nu }x_i^\mu x_j^\nu )exp(\frac{(x_ix_j)^2}{4C}).$$ (3.17) Although $`|x_i`$ are non-orthogonal, $`x_i|x_j`$ exponentially vanishes when $`(x_ix_j)^2`$ gets large. We also find $$x_i|exp(ik\widehat{x})|x_j=exp(ik\frac{(x_i+x_j)}{2}+\frac{i}{2}B_{\mu \nu }x_i^\mu x_j^\nu )exp(\frac{(x_ix_jd)^2}{4C}),$$ (3.18) where $`d^\mu =C^{\mu \nu }k_\nu `$. This matrix element sharply peaks at $`x_ix_j=d`$. It supports our interpretation that the eigenstate $`exp(ik\widehat{x})`$ with $`|k_\mu |>2\pi /l_{NC}`$ can be interpreted as string like extended objects whose length is $`|C^{\mu \nu }k_\nu |`$. When $`|k_\mu |<2\pi /l_{NC}`$, on the other hand, this matrix becomes close to diagonal whose matrix elements go like $$x_i|exp(ik\widehat{x})|x_jexp(ikx_i)x_i|x_j.$$ (3.19) It again supports our interpretation that $`exp(ik\widehat{x})`$ correspond to the ordinary plane waves when $`|k_\mu |<2\pi /l_{NC}`$. They are represented by the matrices which are close to diagonal. We may expand matrices $`\widehat{\varphi }`$ in the twisted reduced model by the following bi-local basis as follows: $$\widehat{\varphi }=\underset{i,j}{}\varphi (x_i,x_j)|x_ix_j|,$$ (3.20) where the Hermiticity of $`\widehat{\varphi }`$ implies $`\varphi ^{}(x_j,x_i)=\varphi (x_i,x_j)`$. The matrices $`\widehat{\varphi }`$ represent $`\widehat{a_\mu }`$ or $`\widehat{\psi }`$ in the super Yang-Mills case but the setting here is more generally applied to an arbitrary noncommutative field theory. The bi-local basis spans the whole $`n^2`$ degrees of freedom of matrices. <sup>7</sup><sup>7</sup>7Bi-local fields have also appeared in $`c=1`$ string theory . Here we work out the translation rule between the momentum eigenstate representation $`\widehat{\varphi }=_k\stackrel{~}{\varphi }(k)exp(ik\widehat{x})`$ and the bi-local field representation of eq.(3.20): $`\stackrel{~}{\varphi }(k)`$ $`=`$ $`{\displaystyle \frac{1}{n}}Tr(exp(ik\widehat{x})\widehat{\varphi })={\displaystyle \frac{1}{n}}{\displaystyle \underset{i,j}{}}x_i|exp(ik\widehat{x})|x_j\varphi (x_j,x_i)`$ $`=`$ $`{\displaystyle \frac{1}{n}}{\displaystyle \underset{x_c}{}}\varphi (x_c,d)exp(ikx_c),`$ $`\varphi (x_c,d)`$ $`=`$ $`{\displaystyle \underset{x_r}{}}exp({\displaystyle \frac{i}{2}}B_{\mu \nu }x_r^\mu x_c^\nu )exp({\displaystyle \frac{(x_rd)^2}{4C}})\varphi (x_j,x_i),`$ (3.21) where $`x_c=(x_i+x_j)/2`$ and $`x_r=x_ix_j`$. From eq.(3.21), we observe that the slowly varying field with the momentum smaller than $`2\pi /l_{NC}`$ consists of the almost diagonal components. Hence close to diagonal components of the bi-local field are identified with the ordinary slowly varying field $`\varphi (x_c)`$. On the other hand, rapidly oscillating fields are mapped to the off-diagonal open string states. A large momentum in the $`\nu `$-th direction $`|k_\nu |>2\pi /l_{NC}`$ corresponds to a large distance in the $`\mu `$-th direction $`|d^\mu |=|C^{\mu \nu }k_\nu |>l_{NC}`$. We can decompose $`d`$ as $`d=d_0+\delta d`$ where $`d_0`$ is a vector which connects two points on the von Neumann lattice and $`|\delta d^\mu |<l_{NC}`$. This decomposition is illustrated in Figure 2. Then the summation over $`x_r`$ in (3.21) is dominated at $`x_r=d_0`$. In this way the large momentum degrees of freedom are more naturally interpreted as extended open string-like fields. They are denoted by ‘open strings’ in this paper. In this representation, we make contact with the quenched reduced models in the large momentum region. Here we remark the important property concerning the infra-red and ultra-violet cut-offs of NCYM constructed with $`n`$ dimensional matrices. Since the unit lattice size of the von Neumann lattice is $`l_{NC}`$, the total lattice size is $`l_{NC}n^{1/\stackrel{~}{d}}`$. It implies that the maximum momentum is $`2\pi n^{1/\stackrel{~}{d}}/l_{NC}`$ by using the relation $`\widehat{p}_\mu =B_{\mu \nu }\widehat{x}^\nu `$. It in turn implies that the minimum momentum of the system is $`2\pi /n^{1/\stackrel{~}{d}}l_{NC}`$ since we have $`n^2`$ momentum modes. The matrix model construction of NCYM implies very natural infra-red and ultra-violet cut-offs which disappear in the large $`n`$ limit. ## 4 Estimations of the string scale In this section, we estimate the string scale $`\alpha _{eff}^{}`$ in IIB matrix model with noncommutative backgrounds. We have explained that the von Neumann lattice naturally appears in the preceding section. We argue that NCYM is superstring theory on the von Neumann lattice. We first give the arguments based on the tree level propagators. We then explain that our claim is supported by T-duality arguments. We give another evidence for it by investigating graviton exchange processes in the next section. As we have shown in the preceding section, the momentum which can be associated with the center of mass motion of an ‘open string’ is not full $`k_\mu `$ but rather $`k_\mu ^c=B_{\mu \nu }\delta d^\nu `$. There we have decomposed $`d^\mu =C^{\mu \nu }k_\nu `$ as $`d=d_0+\delta d`$ where $`d_0`$ is a vector which connects two points on the von Neumann lattice and $`|\delta d^\mu |<l_{NC}`$. We can indeed represent $`\stackrel{~}{\varphi }(k)`$ of eq.(3.21) as follows: $$\stackrel{~}{\varphi }(k^c,d)=\frac{1}{n}\underset{x_c}{}\varphi (x_c,d)exp(ik^cx_c).$$ (4.1) It is because $$exp(iB_{\mu \nu }d_0^\nu x_c^\mu )=1.$$ (4.2) We interpret $`\stackrel{~}{\varphi }(k^c,d)`$ as the creation-annihilation operator for the open string with momentum $`k^c`$ and length $`d`$. We consider the following tree level propagator: $$\underset{k_0}{}<\stackrel{~}{\varphi }(k)\stackrel{~}{\varphi }(k)>exp(ik_0\tau )exp(m\tau )$$ (4.3) where $`k_0`$ and $`\stackrel{}{k}`$ denote time-like and spatial momenta respectively in this section. The mass term is conventionally identified with the zero spacial momentum limit of the correlator eq.(4.3). In order to relate it to the mass of an ‘open string’, we consider a state with $`k_1<2\pi /l_{NC}<k_2,k_3`$. Such a state is extended in $`(x^2,x^3)`$ plane with the length $`l=C\sqrt{k_2^2+k_3^2}`$. As we have argued, the momentum which can be associated with the center of mass motion of an ‘open string’ is not full $`\stackrel{}{k}`$ but rather $`\stackrel{}{k}^c`$. We find $`mBl`$ from the zero momentum limit ($`\stackrel{}{k}^c0`$) of the correlator eq.(4.3). From these considerations, we propose to identify the mass of the state with the length $`l>>l_{NC}`$ as $`Bl`$. We recall that an open string with the length $`l`$ has the mass of $`l/2\pi \alpha ^{}`$. Therefore we find $`2\pi \alpha _{eff}^{}=C`$ with such an identification. We remark that our estimate is consistent with the string theory arguments. From eq.(2.7), we have indeed found that the $`\alpha _{eff}^{}c`$ in section 2. A hint for such an identification has been found in a finite temperature investigation in . Our estimate is also consistent with the space-time uncertainty principle of Yoneya. It is certainly true that we obtain NCYM in the low energy limit of string theory with $`b_{\mu \nu }`$ backgrounds. In such a limit, we retain only those degrees of freedom whose masses are smaller than the string scale. However in IIB matrix model with noncommutative backgrounds, we also find very high energy modes which may be interpreted as long ‘open strings’ due to the relation $`\widehat{x}^\mu =C^{\mu \nu }\widehat{p}_\nu `$. As can be seen in eq.(2.7) they are as massive as oscillator modes if their lengths exceed $`l_{NC}`$. We therefore emphasize here that our formulation is not a low energy limit of string theory. The graviton exchanges are observed only because we have very long ‘open strings’ in the matrix model. To put it differently, infrared singular behaviors are observed in noncommutative field theory only if we consider the ultra-violet cut-off which is much larger than the noncommutativity scale. It is the reason why closed strings do not decouple in such a formulation. In our picture, we interpret the bi-local fields as the zero modes of open strings. We classify the zero modes as ‘momentum modes’ and ‘winding modes’ as follows. We recall the von Neumann lattice $`|x_i>`$ which is constructed by the generators of the translation operators $`U_\mu =exp(il_{NC}e_\mu \widehat{p})`$. We can ‘compactify’ the theory by imposing the following conditions for fluctuations $$\widehat{\varphi }=U_\mu \widehat{\varphi }U_\mu ^{}.$$ (4.4) In this T-dual picture, the von Neumann lattice can be identified with the lattice spanned by the winding modes. We thus classify those modes as ‘winding modes’. We have explained that $`k_\mu ^c`$ can be interpreted as momentum modes which can be associated with the center of mass motion of ‘open strings’ on the von Neumann lattice. We thus classify them as ‘momentum modes’. In string theory, the introduction of constant $`b_{\mu \nu }`$ background is known to interpolate the Neumann and Dirichlet boundary conditions. In the large and small $`b_{\mu \nu }`$ limit, we find Dirichlet and Neumann boundary conditions respectively. We find only ‘winding’ and ‘momentum’ modes in these limits. If we expand IIB matrix model around the commutative backgrounds, we only find ‘winding’ modes. In string theory we also find only ‘winding’ modes if we consider strings which connect D instantons. The advantage of noncommutative backgrounds is that we find the both ‘momentum’ and ‘winding’ modes. Since we have found the both ‘momentum’ and ‘winding’ modes, it is no surprise that the theory possesses T duality. The remarkable property of NCYM is the existence of Morita equivalent pairs . We propose that two Morita equivalent theories can be related by the exchange of the ‘momentum’ and ‘winding’ modes. We have argued that the ‘winding’ modes of NCYM span the von Neumann lattice whose lattice unit size is $`l_{NC}`$. The total lattice size is $`l_{NC}n^{1/\stackrel{~}{d}}`$. We may reinterpret it as the maximum momentum $`2\pi n^{1/\stackrel{~}{d}}/l_{NC}`$ of the dual lattice. The dual lattice possesses the unit lattice size of $`l_{NC}/n^{1/\stackrel{~}{d}}`$. We consider a twisted large $`N`$ reduced model on such a lattice. In this way we find a pair of theories with the compactification radii $`R=l_{NC}n^{1/\stackrel{~}{d}}/2\pi `$ and $`R^{}=l_{NC}/n^{1/\stackrel{~}{d}}2\pi `$. They are related by the duality transformation $`R^{}=\alpha _{eff}^{}/R`$ with $`2\pi \alpha _{eff}^{}=C`$. Next we recall that the ‘momentum’ modes of NCYM are quantized in the unit of $`l_{NC}/n^{1/\stackrel{~}{d}}C`$. We can naturally reinterpret them as the ‘winding’ modes of the dual lattice. These winding modes can be obtained by introducing the unit magnetic flux in $`U(n)`$ gauge theory by imposing twisted boundary conditions. In this sense NCYM and twisted reduced models are Morita equivalent . We remark that the T-duality transformation we have discussed is expressed by the following open string metric transformation in string theory $$G_{\mu \nu }\mathrm{\Theta }^{\mu \rho }G_{\rho \sigma }(\mathrm{\Theta }^T)^{\sigma \nu }$$ (4.5) where $`\mathrm{\Theta }^{\mu \nu }=C^{\mu \nu }/(2\pi R^2)1/n^{2/\stackrel{~}{d}}`$. Our interpretation is that the two metrics which are related by the T-duality transformation in eq.(4.5) describe two tori we have just constructed. We conclude that our estimate of the inverse string tension $`2\pi \alpha _{eff}^{}=C`$ is also supported by the T-duality arguments. We argue that this is the exact result since it is obtained from the exact T duality property of the theory. ## 5 Graviton exchange processes In this section, we study gravitational interactions in IIB matrix model with noncommutative backgrounds. To be specific, we consider photon-photon scattering via exchange of a graviton. This process can be studied by considering block-block interactions. Namely we consider the backgrounds of the following type: $`A_\mu `$ $`=`$ $`\left(\begin{array}{cc}p_\mu +a_\mu & 0\\ 0& p_\mu +a_\mu ^{}\end{array}\right),`$ (5.3) where $`a_\mu `$ and $`a_\mu ^{}`$ denote the backgrounds which represent two colliding photons. The one-loop effective action of IIB matrix model is $$ReW=\frac{1}{2}𝒯r\mathrm{log}(P_\lambda ^2\delta _{\mu \nu }2iF_{\mu \nu })\frac{1}{4}𝒯r\mathrm{log}((P_\lambda ^2+\frac{i}{2}F_{\mu \nu }\mathrm{\Gamma }^{\mu \nu })(\frac{1+\mathrm{\Gamma }_{11}}{2}))𝒯r\mathrm{log}(P_\lambda ^2).$$ (5.4) Here $`P_\mu `$ and $`F_{\mu \nu }`$ are operators acting on the space of matrices as $`P_\mu X`$ $`=`$ $`[p_\mu ,X],`$ $`F_{\mu \nu }X`$ $`=`$ $`[f_{\mu \nu },X],`$ (5.5) where $`f_{\mu \nu }=i[p_\mu ,p_\nu ]`$. Now we expand the general expression of the one-loop effective action (5.4) with respect to the inverse powers of the relative distance between the two blocks. We quote the general expression in what follows: $`W`$ $`=`$ $`𝒯r\left({\displaystyle \frac{1}{P^2}}F_{\mu \nu }{\displaystyle \frac{1}{P^2}}F_{\nu \lambda }{\displaystyle \frac{1}{P^2}}F_{\lambda \rho }{\displaystyle \frac{1}{P^2}}F_{\rho \mu }\right)`$ (5.6) $`2𝒯r\left({\displaystyle \frac{1}{P^2}}F_{\mu \nu }{\displaystyle \frac{1}{P^2}}F_{\lambda \rho }{\displaystyle \frac{1}{P^2}}F_{\mu \rho }{\displaystyle \frac{1}{P^2}}F_{\lambda \nu }\right)`$ $`+{\displaystyle \frac{1}{2}}𝒯r\left({\displaystyle \frac{1}{P^2}}F_{\mu \nu }{\displaystyle \frac{1}{P^2}}F_{\mu \nu }{\displaystyle \frac{1}{P^2}}F_{\lambda \rho }{\displaystyle \frac{1}{P^2}}F_{\lambda \rho }\right)`$ $`+{\displaystyle \frac{1}{4}}𝒯r\left({\displaystyle \frac{1}{P^2}}F_{\mu \nu }{\displaystyle \frac{1}{P^2}}F_{\lambda \rho }{\displaystyle \frac{1}{P^2}}F_{\mu \nu }{\displaystyle \frac{1}{P^2}}F_{\lambda \rho }\right)+O((F_{\mu \nu })^5).`$ Since $`P_\mu `$ and $`F_{\mu \nu }`$ act on the $`(i,j)`$ blocks independently, the one-loop effective action $`W`$ is expressed as the sum of contributions of the $`(i,j)`$ blocks $`W^{(i,j)}`$. Therefore we may consider $`W^{(i,j)}`$ as the interaction between the $`i`$-th and $`j`$-th blocks. Using (5.6) we can easily evaluate $`W^{(i,j)}`$ to the leading order of $`1/\sqrt{(d^{(i)}d^{(j)})^2}`$ as $`W^{(i,j)}`$ $`=`$ $`{\displaystyle \frac{1}{r^8}}(𝒯r^{(i,j)}(F_{\mu \nu }F_{\nu \lambda }F_{\lambda \rho }F_{\rho \mu })2𝒯r^{(i,j)}(F_{\mu \nu }F_{\lambda \rho }F_{\mu \rho }F_{\lambda \nu })`$ (5.7) $`+{\displaystyle \frac{1}{2}}𝒯r^{(i,j)}(F_{\mu \nu }F_{\mu \nu }F_{\lambda \rho }F_{\lambda \rho })+{\displaystyle \frac{1}{4}}𝒯r^{(i,j)}(F_{\mu \nu }F_{\lambda \rho }F_{\mu \nu }F_{\lambda \rho })`$ $`=`$ $`{\displaystyle \frac{3}{2r^8}}(n_j\stackrel{~}{b}_8(f^{(i)})n_i\stackrel{~}{b}_8(f^{(j)})`$ $`8Tr(f_{\mu \nu }^{(i)}f_{\nu \sigma }^{(i)})Tr(f_{\mu \rho }^{(j)}f_{\rho \sigma }^{(j)})+Tr(f_{\mu \nu }^{(i)}f_{\mu \nu }^{(i)})Tr(f_{\rho \sigma }^{(j)}f_{\rho \sigma }^{(j)})),`$ where $$\stackrel{~}{b}_8(f)=\frac{2}{3}(Tr(f_{\mu \nu }f_{\nu \lambda }f_{\lambda \rho }f_{\rho \mu })+2Tr(f_{\mu \nu }f_{\lambda \rho }f_{\mu \rho }f_{\lambda \nu })\frac{1}{2}Tr(f_{\mu \nu }f_{\mu \nu }f_{\lambda \rho }f_{\lambda \rho })\frac{1}{4}Tr(f_{\mu \nu }f_{\lambda \rho }f_{\mu \nu }f_{\lambda \rho })).$$ (5.8) So the photon-photon scattering amplitude which corresponds to nonplanar diagrams in noncommutative gauge theory is $`{\displaystyle \frac{3}{2r^8}}(8Tr(f_{\mu \nu }^{(i)}f_{\nu \sigma }^{(i)})Tr(f_{\mu \rho }^{(j)}f_{\rho \sigma }^{(j)})+Tr(f_{\mu \nu }^{(i)}f_{\mu \nu }^{(i)})Tr(f_{\rho \sigma }^{(j)}f_{\rho \sigma }^{(j)}))`$ (5.9) $`=`$ $`{\displaystyle \frac{3}{2}}({\displaystyle \frac{1}{2\pi }})^{\stackrel{~}{d}}B^{\stackrel{~}{d}8}{\displaystyle d^{\stackrel{~}{d}}xd^{\stackrel{~}{d}}y\frac{1}{(xy)^8}}`$ $`\{8tr(f_{\mu \nu }(x)f_{\nu \sigma }(x))tr(f_{\mu \rho }(y)f_{\rho \sigma }(y))+tr(f_{\mu \nu }(x)f^{\mu \nu }(x))tr(f_{\rho \sigma }(y)f^{\rho \sigma }(y))\},`$ where we have used our mapping rule eq. (3.8). We consider the scattering of two incident photons with the wave functions $$e(p)_\mu exp(ipx),e(q)_\mu exp(iqx),$$ (5.10) where $`p^2=q^2=0`$ and $`pe(p)=qe(q)=0`$. In this case $`f_{\mu \nu }(p)`$ $`=`$ $`p_\mu e(p)_\nu p_\nu e(p)_\mu ,`$ $`f_{\mu \rho }(p)f^{\rho \nu }(q)`$ $`=`$ $`p_\mu qe(p)e(q)_\nu q_\nu pe(q)e(p)_\mu `$ (5.11) $`+p_\mu q_\nu e(p)e(q)+e(p)_\mu e(q)_\nu pq.`$ If we consider the forward scattering limit $`k=pq0`$: $$f_{\mu \rho }(p)f^{\rho \nu }(q)p_\mu p_\nu ,$$ (5.12) we find only graviton exchange in this limit. $$12(\frac{1}{2\pi })^{\stackrel{~}{d}}B^{\stackrel{~}{d}8}d^{\stackrel{~}{d}}xd^{\stackrel{~}{d}}y\frac{1}{(xy)^8}tr(f_{\mu \rho }(x)f^{\rho \nu }(x))tr(f_{\mu \sigma }(y)f^{\sigma \nu }(y)).$$ (5.13) This expression reminds us the one photon exchange amplitude between two conserved currents $`j`$ and $`\stackrel{~}{j}`$ in $`QED_4`$: $`{\displaystyle \frac{d^4k}{(2\pi )^4}j_\mu (k)\frac{1}{k^2+iϵ}\stackrel{~}{j}^\mu (k)}`$ (5.14) $`=`$ $`{\displaystyle \frac{i}{4\pi }}{\displaystyle d^4xd^4yj_\mu (x)\frac{1}{(xy)^2iϵ}\stackrel{~}{j}^\mu (y)}.`$ We decompose currents into positive and negative frequency parts $`j_\mu (x)`$ $`=`$ $`j_\mu ^+(x)+j_\mu ^{}(x),`$ $`j_\mu ^+(x)`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}𝑑\omega exp(i\omega x^0)j_\mu ^+(\omega ,\stackrel{}{x}),`$ $`j_\mu ^{}(x)`$ $`=`$ $`{\displaystyle _0^{\mathrm{}}}𝑑\omega exp(i\omega x^0)j_\mu ^{}(\omega ,\stackrel{}{x}).`$ (5.15) We rewrite eq.(5.14) as follows $`{\displaystyle \frac{1}{4}}{\displaystyle d^4x𝑑\stackrel{}{y}j_\mu ^{}(x^0,\stackrel{}{x})\frac{1}{|\stackrel{}{x}\stackrel{}{y}|}\stackrel{~}{j}^{+\mu }(x^0|\stackrel{}{x}\stackrel{}{y}|,\stackrel{}{y})}`$ $`+{\displaystyle \frac{1}{4}}{\displaystyle d^4y𝑑\stackrel{}{x}j_\mu ^+(y^0|\stackrel{}{x}\stackrel{}{y}|,\stackrel{}{x})\frac{1}{|\stackrel{}{x}\stackrel{}{y}|}\stackrel{~}{j}^\mu (y^0,\stackrel{}{y})}.`$ (5.16) In this way we find retarded Lienard-Wiechert type interactions. The point we would like to make here is that covariance implies causality. Since ten dimensional covariance is manifest in IIB matrix model, it naturally leads to ten dimensional causality in Minkowski space-time. On the other hand, ensuring ten dimensional causality is highly nontrivial in $`AdS`$/CFT correspondence. We recall the relevant part of the NCYM action as follows $`B^{\frac{\stackrel{~}{d}}{2}4}({\displaystyle \frac{1}{2\pi }})^{\frac{\stackrel{~}{d}}{2}}{\displaystyle \frac{1}{4g^2}}{\displaystyle d^{\stackrel{~}{d}}xtr(f_{\alpha \beta }f^{\alpha \beta })_{}}.`$ (5.17) The energy momentum tensor can be read off from it in the low energy limit as $$B^{\frac{\stackrel{~}{d}}{2}4}(\frac{1}{2\pi })^{\frac{\stackrel{~}{d}}{2}}\frac{1}{g^2}(f_{\mu \rho }f^{\rho \nu }\frac{1}{4}\delta _\mu ^\nu f_{\rho \sigma }f^{\sigma \rho }).$$ (5.18) So we can rewrite the gravitational interaction eq.(5.13) as follows $$12g^4d^{\stackrel{~}{d}}xd^{\stackrel{~}{d}}y\frac{1}{(xy)^8}T_{\mu \nu }(x)T^{\mu \nu }(y).$$ (5.19) We recall the $`\stackrel{~}{d}`$ dimensional propagators $$\frac{d^{\stackrel{~}{d}}p}{(2\pi )^{\stackrel{~}{d}}}\frac{1}{p^2}exp(ipx)=\frac{\mathrm{\Gamma }(\frac{d2}{2})}{4\pi ^{\frac{\stackrel{~}{d}}{2}}}\frac{1}{x^{\stackrel{~}{d}2}}.$$ (5.20) For $`\stackrel{~}{d}=10`$, we obtain $$G_{10}(x)=\frac{3}{2\pi ^5}\frac{1}{x^8}.$$ (5.21) The gravitational coupling $`\kappa `$ is found to be $$\kappa ^2=16\pi ^5g^4.$$ (5.22) We also read off the $`\stackrel{~}{d}`$ dimensional Yang-Mills coupling from eq.(5.17) as $$g_{YM}^2=C^{(\stackrel{~}{d}8)/2}g^2(2\pi )^{\stackrel{~}{d}/2}.$$ (5.23) Here we quote string theory predictions: $$\kappa ^2=\frac{1}{2}(2\pi )^7g_s^2\alpha ^4,g_{YM}^2=(2\pi )^{\stackrel{~}{d}3}g_s\alpha ^{(\stackrel{~}{d}4)/2}.$$ (5.24) Eq.(5.24) agrees with eq.(5.22) and eq.(5.23) with our identification $`2\pi \alpha _{eff}^{}=C`$. We also find that the IIB matrix model coupling can be expressed by $`\alpha _{eff}^{}`$ and $`g_s`$ as $$g^2=(2\pi )g_s\alpha _{eff}^{}{}_{}{}^{2}.$$ (5.25) which is consistent with our previous estimate through D-strings. What these investigations indicate is that NCYM is superstring theory with the above identified string scale and string coupling. We have argued that it is superstring theory on the von Neumann lattice. Since the lattice spacing $`l_{NC}`$ is not visible in the low energy limit, it may be expected that it behaves like ordinary superstring theory in the low energy limit. We find that the gravitational interaction eq.(5.19) exhibits the identical power law behavior $`1/(xy)^8`$ irrespective of the dimensionality of the backgrounds $`\stackrel{~}{d}`$. It appears as if these background represent D-branes in flat ten dimensional space-time. However we argue that such an interpretation is premature since we have only considered the one loop effect here. We argue that a more reliable picture is obtained through supergravity approach which allows us to investigate nonperturbative effects. As it is explained in section 2, we find Newton’s force law with these backgrounds. This is due to the existence of a massless bound state a la Randall and Sundrum. Such an effect is not visible in the perturbative calculations in this section. Therefore the supurgravity description of $`NCYM_4`$ suggests a nonperturbative compactification mechanism in IIB superstring and matrix model. In the matrix model construction, the longitudinal size of the system is bounded by $`l_{NC}n^{1/4}`$. It also implies that the transversal size $`r`$ is bounded by $`l_{NC}n^{1/4}`$ since it is identified with the maximum energy scale of the system (multiplied by $`l_{NC}^2`$). In eq.(2.8), the dilaton expectation value is $`O(1)`$ at the noncommutativity scale $`r^21`$. We then find $`g_{\mathrm{}}`$ is $`O(1/n)`$ since the dilaton decays like $`1/r^4`$ beyond the noncommutativity scale. We have fixed $`\lambda =g_{\mathrm{}}\alpha ^2b^2`$ to be $`O(1)`$ which implies that $`\alpha ^{}b\sqrt{n}`$. We conclude that $`l_{NC}l_s/n^{1/4}`$ and $`r`$ never exceeds $`O(l_s)`$ where $`l_s`$ is the string scale. Therefore there is simply no region with $`r>l_s`$ in the matrix model. We have taken the noncommutativity scale to be $`O(1)`$ and the cuff-off scale of $`r`$ becomes $`O(n^{1/4})`$. The cut-off can be removed in the large $`n`$ limit of the matrix model construction. In this way we can realize the entire space-time which is described by eq.(2.8). ## 6 Conclusions and Discussions In this paper we have argued that we can obtain Newton’s force law with $`NCYM_4`$. Since it contains four dimensional gauge theory and gravitation, it is a candidate of the unified theory. It can be regarded as a compactification of ten dimensional IIB superstring straight down to four dimensions. It is naturally obtained in IIB matrix model with noncommutative backgrounds. Therefore it provides a nonperturbative compactification mechanism of matrix models. We have identified the bi-local fields as the zero modes of open strings. They can be interpreted as the ‘momemtum’ and ‘winding’ modes on the von Neumann lattice. Our identification of the effective string scale with the noncommutativity scale is consistent with the exact T-duality which interchanges the ‘momentum’ and ‘winding’ modes. Although we have identified the zero modes of open strings, we have not constructed oscillator modes. We expect to find them in higher order diagrams. Let us consider a propagator (ribbon diagram). We associate it with a bi-local field since the double lines of the ribbon are mapped to two distinct space-time points. We need to draw many loops inside the ribbon at higher orders in perturbation theory. We can assign a space-time point to each loop. Our conjecture is that such an object can be interpreted as the propagator of oscillation modes. These arguments are illustrated in Figures 3 and 4. It is important to recall here that we identify the string coupling with the topological expansion parameter of the Feynman diagrams of NCYM. It is not equal to the NCYM coupling although the both are related in the low energy limit as suggested by supergravity solutions. The tree level string propagator is obtained by summing all planar diagrams. Our proposal that NCYM with maximal SUSY may be interpreted as superstring theory on the von Neumann lattice should be understood in this context. As we have pointed out, NCYM is obtained with a particular classical background in IIB matrix model. IIB matrix model is postulated as a nonperturbative formulation of superstring theory. In our proposal, the matrices $`A_\mu `$ are to be interpreted as space-time coordinates. If so, $`\stackrel{~}{d}`$ dimensional distributions of eigenvalues of matrices represent $`\stackrel{~}{d}`$ dimensional space-time. It is then expected that we find $`\stackrel{~}{d}`$ dimensional gauge theory and gravitation with such a background. In this paper we have argued that it is indeed the case with maximally supersymmetric backgrounds. From the findings in this paper, we draw the conclusion that NCYM provides a strong support for our basic premises of our IIB matrix model conjecture. Acknowledgments This work is supported in part by the Grant-in-Aid for Scientific Research from the Ministry of Education, Science and Culture of Japan.
warning/0004/hep-th0004172.html
ar5iv
text
# 1 Introduction ## 1 Introduction Gauge theories and string theories are very important fields of modern theoretical physics. Construction of a non-perturbative formulation of string theories is one of the main subjects which has not yet been completely understood. Although non-perturbative formulations of gauge theories were constructed as lattice gauge theories, we are still interested in analytical studies on non-perturbative effects of gauge theories. The relation between gauge theories and string theories has been studied for a long time, and both theories have developed while affecting each other deeply. Thus, studies on the connection between them seems to be important and is expected to be fruitful. With this in mind, we consider QCD strings in this article, from the viewpoint of Liouville theory. Investigation of non-perturbative effects of QCD (for example, confinement of quarks) has been one of the motivations for string theories. In the context of QCD strings, confinement of quarks is explained in terms of a linear potential produced by the “string” stretched between quark and antiquark. There are many reasons why we consider such a string theory, although this naive explanation has not been completed nor proved to this time.<sup>2</sup><sup>2</sup>2For a review of these topics, see Ref. . In the strong coupling expansion of lattice gauge theories, the quark-antiquark potential is given by the expectation value of the Wilson loop, each term of which corresponds to different configuration of the lattice surfaces. This reminds us of the sum of the “world-sheet surfaces” of an open string whose boundary is the Wilson loop. We can find another description of gauge theories in terms of strings in the large-$`N`$ limit. In large-$`N`$ gauge theories, Feynman diagrams are classified by the “Euler number,” and the physical observables are expressed as a series with respect to the “topology” of the graph. The dominant terms correspond to planar diagrams. This also reminds us of a “world-sheet” description. Furthermore, in the description of the dual Meisner effect, the fluxes of color electric charge are collimated, and form a “string” between quarks. We also find similar correspondence between gauge theories and string theories in the contexts of D-branes and AdS/CFT.<sup>3</sup><sup>3</sup>3We have found other connections between gauge theories and strings . In consideration of the above discussion, we study QCD strings with the following assumptions. (1) We assume that the color charge is attached to the ends of open strings, and we regard the world-sheet boundaries as Wilson loops. (2) The Wilson loops are not dynamical, since we attempt to describe pure YM theories in which quarks are not dynamical. (3) We assume that the strings are bosonic and exist in four-dimensional spacetime, because our purpose is to describe four-dimensional (large-$`N`$) non-SUSY gauge theories. ## 2 Review of Liouville theory It is most natural to assume that QCD strings are bosonic and four-dimensional objects. Therefore, we must to find some nontrivial way to construct a consistent noncritical string theory. Quantization of noncritical strings has been considered in Liouville theories . In particular, strings for $`d1`$ have been consistently quantized using DDK theory presented by Distler and Kawai and David . (Here $`d`$ denotes the spacetime dimension in which strings exist.) For this reason, we first review the work on the DDK theory.<sup>4</sup><sup>4</sup>4For a review of DDK, see Ref. . ### 2.1 DDK theory without boundaries Let us start with a $`d`$-dimensional ($`d26`$) bosonic string without boundaries for simplicity. We use the Euclidean signature both for the world-sheet metric and for the spacetime metric. The Polyakov action is $`S_M={\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle d^2\xi \sqrt{g}(g^{ab}_aX^\mu _bX_\mu )},`$ (2.1) where $`\mu `$ runs from $`1`$ to $`d`$. The partition function with respect to this action is diffeomorphism invariant, and we can fix the world-sheet metric as $`g_{ab}=\widehat{g}_{ab}e^{\varphi (\xi )}.`$ (2.2) However, the Weyl invariance of the partition function is broken for $`d26`$ at the quantum level, and the freedom of the Weyl transformation is no longer a gauge freedom. Therefore, we have to perform the path integral with respect to $`\varphi `$ rigorously. However, the measure $`[d\varphi ]_g`$ of the path integral with respect to $`\varphi `$ is given by the norm $`\delta \varphi _g={\displaystyle d^2\xi \sqrt{g}(\delta \varphi )^2}={\displaystyle d^2\xi \sqrt{\widehat{g}}e^{\varphi (\xi )}(\delta \varphi )^2},`$ (2.3) which depends on $`\varphi `$ itself in a complicated manner. Thus, to perform the path integral with this measure is difficult and seems to be almost impossible. In Louville theory based on the DDK argument, the measure of the path integral is redefined with respect to some fixed world-sheet metric $`\widehat{g}_{ab}`$ to avoid the above described difficulty. Namely, we use the measure $`[d\varphi ]_{\widehat{g}}`$ given by the norm $`\delta \varphi _{\widehat{g}}={\displaystyle d^2\xi \sqrt{\widehat{g}}(\delta \varphi )^2},`$ (2.4) which does not depend on $`\varphi `$. With this redefinition, we have to use a Jacobian $`J`$ to maintain consistency: $`[dX]_g[db]_g[dc]_g[d\varphi ]_g`$ $`=`$ $`[dX]_{\widehat{g}}[db]_{\widehat{g}}[dc]_{\widehat{g}}[d\varphi ]_{\widehat{g}}J,`$ (2.5) $`J`$ $`=`$ $`e^{S_L},`$ (2.6) where $`b`$ and $`c`$ are the ghost fields. We have the relations $`[dX]_g[db]_g[dc]_g=[dX]_{\widehat{g}}[db]_{\widehat{g}}[dc]_{\widehat{g}}e^{S_{\mathrm{Jac}}}`$ (2.7) $`S_{\mathrm{Jac}}={\displaystyle \frac{1}{2}}{\displaystyle \frac{26d}{48\pi }}{\displaystyle d^2\xi \sqrt{\widehat{g}}\{\widehat{g}^{ab}_a\varphi _b\varphi +2\widehat{R}\varphi \}},`$ (2.8) where $`\widehat{R}`$ is the world-sheet scalar curvature with respect to the metric $`\widehat{g}_{ab}`$. (Note that (2.7) does not contain $`[d\varphi ]`$, while (2.5) does contain $`[d\varphi ]`$.) Therefore, the Jacobian $`J`$ can be naturally assumed to be $$S_L=ud^2\xi \sqrt{\widehat{g}}\{\widehat{g}^{ab}_a\varphi _b\varphi +q\widehat{R}\varphi \},$$ (2.9) where $`u`$ and $`q`$ are some constants. Then, the total action, including the Jacobian term, is $`S`$ $`=`$ $`S_M+S_L+S_{\mathrm{ghost}}`$ (2.10) $`=`$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle d^2\xi \sqrt{\widehat{g}}\{\widehat{g}^{ab}_a\phi _b\phi +\widehat{g}^{ab}_aX^\mu _bX_\mu +\alpha ^{}Q\widehat{R}\phi \}}+S_{\mathrm{ghost}},`$ where we have redefined the factor $`q`$ and the field $`\varphi `$ as $`Q=\sqrt{{\displaystyle \frac{4\pi u}{\alpha ^{}}}}q,`$ (2.11) $`\phi (\xi )=\sqrt{4\pi \alpha ^{}u}\varphi (\xi )={\displaystyle \frac{q}{Q\alpha ^{}}}\varphi (\xi ).`$ (2.12) In the action (2.10), the field $`\phi `$ has a kinetic term and can be regarded as a new “coordinate” of the target space<sup>5</sup><sup>5</sup>5Note that $`\phi `$ has the dimension of length, though $`\varphi `$ is a dimensionless field.; we obtain one more spacetime dimension and a linear dilaton term through the process of quantization. The world-sheet metric $`g_{ab}`$ has been replaced with $`\widehat{g}_{ab}`$ in the process. Let us consider the partition function $`Z`$ with respect to the action (2.10), $`Z={\displaystyle [d\phi ][dX]e^S},`$ (2.13) where $`[d\phi ]`$ and $`[dX]`$ stand for $`[d\phi ]_{\widehat{g}}`$ and $`[dX]_{\widehat{g}}`$. (We omit $`\widehat{g}`$ from expressions of the measure in the following.) Here $`[dX]`$ represents the measure with respect to the matter and the ghost. It also represents the measure for the modular integration if $`N_g>0`$, where $`N_g`$ is the number of the genus of the world-sheet. We note that the partition function does not change under the simultaneous transformations that keep $`g_{ab}=\widehat{g}_{ab}e^{\frac{Q\alpha ^{}}{q}\phi }`$ invariant, $`\widehat{g}_{ab}\widehat{g}_{ab}e^{\delta (\xi )},`$ (2.14) $`\phi (\xi )\phi (\xi )^{}=\phi (\xi ){\displaystyle \frac{Q\alpha ^{}}{q}}\delta (\xi ),`$ (2.15) if the boundary condition of the integration in (2.13) allows the shift of $`\phi `$ given by (2.15). This is because the action in the initial formulation (2.1) depends only on $`g_{ab}`$ and $`X^\mu `$, which do not change under the above transformations. Thus, $`Z={\displaystyle [d\phi ][dX]e^{S[\widehat{g},X,\phi ]}}={\displaystyle [d\phi ^{}][dX]e^{S[\widehat{g}e^{\delta (\xi )},X,\phi ^{}]}}.`$ (2.16) After rewriting the dummy variable $`\phi ^{}`$ as $`\phi `$, we note the very important fact that the partition function is now represented in a Weyl invariant way with respect to the new metric $`\widehat{g}_{ab}`$. Now we are in a position to determine the value of $`Q`$. We found that the partition function is Weyl invariant. Therefore, the total central charge of the theory should be zero: $`C_\phi +C_M+C_{\mathrm{ghost}}=0,`$ (2.17) where $`C_\phi `$ is the central charge for $`S_L`$, $`C_M`$ is the central charge for $`S_M`$, which is equal to $`d`$, and $`C_{\mathrm{ghost}}`$ is the central charge for $`S_{\mathrm{ghost}}`$, which is calculated to be $`26`$. We can evaluate $`C_\phi `$ with a standard technique of conformal field theory as $`C_\phi =1+6\alpha ^{}Q^2.`$ (2.18) Thus (2.17) and (2.18) give us $`Q=\pm \sqrt{{\displaystyle \frac{25d}{6\alpha ^{}}}}.`$ (2.19) We point out that our action is exactly the same as that of the “linear dilaton string” in $`d+1`$ dimensional spacetime. In the linear dilaton string, $`\phi `$ is regarded as one of the spacetime coordinates, and $`Q`$ is given as a factor which appears in the dilaton term. However, from the viewpoint of Liouville theory, we do not regard $`\phi `$ as a real physical coordinate of spacetime. It is the parameter of the Weyl transformation and becomes the new spacetime coordinate through the quantization process. Thus, we call $`\phi `$ (or $`\varphi `$) the “Liouville mode” in this article. Now we have some problems. First, the Liouville mode $`\phi `$ does not have a stable vacuum with the action (2.10). In Refs. and , a cosmological constant is added to the action to obtain a stable vacuum. The Polyakov action with the (renormalized) cosmological constant $`\mu `$ is $`S_M={\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle d^2\xi \sqrt{g}(g^{ab}_aX^\mu _bX_\mu +\alpha ^{}\mu )}.`$ (2.20) It is natural to assume that the cosmological constant term is deformed by the quantum effect as $`{\displaystyle \frac{\mu }{4\pi }}{\displaystyle d^2\xi \sqrt{\widehat{g}}e^\varphi }`$ $``$ $`{\displaystyle \frac{\mu }{4\pi }}{\displaystyle d^2\xi \sqrt{\widehat{g}}e^{\gamma \varphi }}`$ (2.21) $`={\displaystyle \frac{\mu }{4\pi }}{\displaystyle d^2\xi \sqrt{\widehat{g}}e^{\alpha \phi }},`$ where $`\gamma `$ is a constant indicating an anomalous dimension, and we defined $`\alpha `$ as $`\alpha \frac{Q\alpha ^{}}{q}\gamma `$. To preserve the Weyl invariance of the theory, we choose $`\alpha `$ so that the cosmological constant term is Weyl invariant. This is realized if the conformal dimension of the operator $`e^{\alpha \phi }`$ is 2. Namely, we demand $`e^{\alpha \phi }`$ to be a (1,1)-primary operator. In complex coordinates, the holomorphic part of the energy-momentum tensor given by the action (2.10) is $`T_{ZZ}={\displaystyle \frac{1}{\alpha ^{}}}(\phi \phi Q^2\phi ),`$ (2.22) and the conformal weight $`\mathrm{\Delta }`$ of $`e^{\alpha \phi }`$ is given by $`\mathrm{\Delta }={\displaystyle \frac{\alpha ^{}}{2}}\alpha \left(Q{\displaystyle \frac{\alpha }{2}}\right),`$ (2.23) which should be 1. Thus we obtain two solutions: $`\alpha _\pm =Q\pm \sqrt{Q^2{\displaystyle \frac{4}{\alpha ^{}}}}=\sqrt{{\displaystyle \frac{25d}{6\alpha ^{}}}}\pm \sqrt{{\displaystyle \frac{1d}{6\alpha ^{}}}}.`$ (2.24) In the classical limit ($`d\mathrm{}`$), $`\alpha `$ should be zero, and we take the branch of $`\alpha _{}`$. Then we have a stable vacuum for a negative world-sheet curvature $`\widehat{R}`$. This is because the potential for $`\phi `$ is given by $`V(\phi )={\displaystyle \frac{1}{4\pi }}{\displaystyle d^2\xi (Q\widehat{R}\phi +\mu e^{\alpha \phi })},`$ (2.25) and has a minimum value if the factor of $`\phi `$ in the first term is negative. (We have assumed a constant-$`\phi `$ configuration as a classical solution here.) We have a second problem with the strings for $`d>1`$. In this region, $`\alpha `$ is a complex number. Thus we have to regard $`\mu e^{\alpha \phi }`$ as a tachyon vertex operator with momentum in the $`\phi `$ direction, rather than a cosmological constant term. Furthermore, the composite operator $`e^{\alpha \phi }`$ becomes non-normalizable . We also have to consider the condensation of the target-space tachyons, since we treat a noncritical bosonic string in which the tachyonic mode is not projected out. We point out that we do not have target-space tachyons if $`d1`$. This is because the Liouville theory is described as a $`d+1`$-dimensional string theory, and the world-sheet oscillation can be fixed completely by the gauge symmetry if $`d+12`$. Thus the tachyonic mode cannot appear. For this reason, we can construct a Weyl invariant string theory with a cosmological constant for the case $`d1`$. A consistent model for quantized noncritical strings for $`d>1`$ has not yet been constructed. ### 2.2 Liouville theory with boundaries For noncritical strings with a boundary, we assume $`S_L`$ to be $$S_L=\frac{1}{4\pi \alpha ^{}}_{}d^2\xi \sqrt{\widehat{g}}\{\widehat{g}^{ab}_a\phi _b\phi +\alpha ^{}Q\widehat{R}\phi \}+\frac{1}{2\pi \alpha ^{}}_{}𝑑s\sqrt{\widehat{g}_{ss}}\alpha ^{}Q\widehat{k}\phi ,$$ (2.26) where $`s`$ is a parameter of the boundary $``$ and $`ds\sqrt{\widehat{g}_{ss}}`$ denotes an invariant infinitesimal length on it . The quantity $`\widehat{k}`$ is the extrinsic curvature with respect to the metric $`\widehat{g}_{ab}`$. If the world-sheet has many boundaries, $``$ denotes all of them. The only difference between this and the boundaryless case is the existence of the boundary terms. We can also add a boundary cosmological constant term, $`\mu _b{\displaystyle _{}}𝑑s\sqrt{\widehat{g}_{ss}}e^{\frac{\alpha }{2}\phi }.`$ (2.27) For strings with boundaries, we also need to consider the boundary conditions. We can consider several types of boundary conditions for $`\phi `$.<sup>6</sup><sup>6</sup>6Boundary conditions and boundary states for linear dilaton theory have been considered by several authors. (For example, see Refs. ,, and references therein.) Now we set $`\widehat{k}=0`$ for simplicity. The reason why this choice is natural is discussed in §3.1. In this case, we have two choices for the boundary conditions. One of them is Neumann boundary conditions, and the other is Dirichlet boundary conditions. Neumann boundary conditions allow the ends of open strings to move freely. In other words, Neumann boundary conditions can be regarded as an equation of motion for the endpoint. Therefore, it does not restrict the discussion in §2.1, and the argument given for the boundaryless case still holds. On the other hand, we have to be careful if we consider Dirichlet boundary conditions. This is because the shift of $`\phi `$ (2.15) is not consistent with Dirichlet boundary conditions, and thus Weyl invariance is broken, in general. However, we show in §3.3 that we can use Dirichlet boundary conditions without breaking Weyl invariance in some special case. ## 3 Liouville theory as QCD string — <br>— A trial to go beyond the $`𝒅\mathbf{=}\mathrm{𝟏}`$ barrier ### 3.1 Liouville theory as a QCD string We presented our basic assumptions for noncritical QCD strings in §1. Here we make these more precise before proceeding with a further argument. The boundaries of the world-sheet correspond to non-dynamical rigid Wilson loops. In this sense, the boundary conditions for $`X^\mu `$ should be Dirichlet boundary conditions. A nontrivial problem is how to choose the boundary conditions for $`\phi `$. We choose the topology of the classical world-sheet to be a cylinder, and we set $`\widehat{k}=0`$ for simplicity. This is reasonable for the calculation of the static quark-antiquark potential. We usually consider a rectangular Wilson loop of infinite length along the time direction to calculate it. However, $`\widehat{k}`$ diverges at the corners of the rectangle, and this makes calculations difficult. To avoid this, we connect the shorter sides (the space-like sides) of the rectangle and thereby make it periodic, like a ring. This configuration consists of two parallel circular Wilson loops, and is like a cylinder. Then, the corners disappear, and the boundaries are straight from the two-dimensional viewpoint on the world-sheet. Then we can set $`\widehat{k}=0`$ naturally on the world-sheet. Of course, the color charges on the loops are set to make a color singlet. In the last step of the calculation, the periodicity of the loops is set to infinity, and thus the calculated value of the potential should attain the same value as that for an infinitely long rectangular Wilson loop. Next, we make our problems clear. To go beyond the $`d=1`$ barrier, we have mainly two problems to solve. First, we must stabilize the vacuum of the field $`\phi `$, or fix the zero mode of $`\phi `$ to stabilize the classical configuration of the string. Second, we have to treat the condensation of the target space tachyons if they exist. We propose some basic ideas to solve these problems in the following sections. We find that Dirichlet boundary conditions play important roles in the generalized Liouville theory. ### 3.2 Generalization of Liouville theory We saw that the cosmological constant term becomes tachyonic for $`d>1`$. Thus we cannot use it to stabilize $`\phi `$. One idea to stabilize $`\phi `$ without a cosmological constant is to add another Weyl invariant term to the action, which generates the minimum of $`V(\phi )`$. We point out that one of the simplest candidates for such a term is $`{\displaystyle \frac{\mu ^{}}{4\pi \alpha ^{}}}{\displaystyle d^2\xi e^{2Q\phi }\widehat{g}^{ab}_aX^\mu _bX_\mu },`$ (3.1) where $`\mu ^{}`$ in (3.1) is a constant. This is because one can verify that the operator $`e^{2Q\phi }\widehat{g}^{ab}_aX^\mu _bX_\mu `$ (3.2) has conformal dimension 2, namely $`\mathrm{\Delta }(e^{2Q\phi })=0`$ from (2.23). Then we obtain a new action with (3.1): $`S={\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle d^2\xi \sqrt{\widehat{g}}\{(1+\mu ^{}e^{2Q\phi })\widehat{g}^{ab}_aX^\mu _bX_\mu +\widehat{g}^{ab}_a\phi _b\phi +\alpha ^{}Q\widehat{R}\phi \}}+S_{\mathrm{ghost}}.`$ (3.3) However, one finds that the above action is not exactly Weyl invariant.<sup>7</sup><sup>7</sup>7Although the composite operator (3.2) is a (1,1)primary operator, multiple insertion of them creates another divergence, which causes breakdown of the scale invariance. Thus the insertion of the operator (3.2) into the action breaks Weyl invariance. In the case of a cosmological constant term, the multiple insertions do not give any divergence. Let us consider the most general action for Liouville theory. We naturally impose $`SO(d)`$ symmetry on $`d`$-dimensional spacetime, and we then obtain $`S={\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle d^2\xi \sqrt{\widehat{g}}\{a^2(\phi )(X^\mu )^2+(\phi )^2+\alpha ^{}R^{(2)}\mathrm{\Phi }(\phi )\}},`$ (3.4) where we have fixed the $`\phi \phi `$-component of the target-space metric to be 1, with appropriate redefinition (or target space general coordinate transformation) of $`\phi `$. Now the cosmological constant term is assumed to be zero. (We have suppressed the ghost term here. Also, note that we have dropped the boundary terms, since we set $`\widehat{k}=0`$.) The $`\beta `$-functions which should vanish for Weyl invariance are $`\beta _{MN}^G=\alpha ^{}(_{MN}+2_M_N\mathrm{\Phi })+O(\alpha ^2),`$ (3.5) $`\beta ^\mathrm{\Phi }=\alpha ^{}\left(Q^2{\displaystyle \frac{1}{2}}^2\mathrm{\Phi }+(\mathrm{\Phi })^2\right)+O(\alpha ^2),`$ (3.6) where $`M`$ and $`N`$ run from $`1`$ to $`d+1`$, and $`X^{d+1}=\phi `$. $`_{MN}`$ is the Ricci tensor of the $`d+1`$-dimensional target space. Collecting the above results, the equation of motion up to order $`\alpha ^2`$ become $`0=_{MN}+2_M_N\mathrm{\Phi },`$ (3.7) $`0=Q^2{\displaystyle \frac{1}{2}}^2\mathrm{\Phi }+(\mathrm{\Phi })^2,`$ (3.8) where $`Q`$ is given by (2.19). We stress that we can use these equations only in the region $`\alpha ^{}1`$, where the perturbative approximation for the $`\beta `$-functions is valid. The solution of these equations is given in Ref. as $`a(\phi )=a_0\sqrt{{\displaystyle \frac{1+\lambda e^{2Q\phi }}{1\lambda e^{2Q\phi }}}},`$ (3.9) where $`\lambda `$ and $`a_0`$ are constants, and $`\mathrm{\Phi }=Q\phi {\displaystyle \frac{3}{2}}\mathrm{log}(1\lambda e^{2Q\phi })+{\displaystyle \frac{1}{2}}\mathrm{log}(1+\lambda e^{2Q\phi })+\text{constant}.`$ (3.10) The allowed region for $`\phi `$ is given by $`|\lambda e^{2Q\phi }|<1`$, or $`\phi <{\displaystyle \frac{1}{2Q}}\mathrm{log}|\lambda |.`$ (3.11) We have to check the validity of this solution before further calculations. The target space scalar curvature for the solution (3.9) is given by $``$ $`=`$ $`3\left({\displaystyle \frac{a^{}}{a}}\right)^25{\displaystyle \frac{a^{\prime \prime }}{a}}`$ (3.15) $`=`$ $`{\displaystyle \frac{4Q^2b}{(1b^2)^2}}(5b^2+8b+5)\{\begin{array}{cc}\mathrm{}\hfill & (\lambda >0)\hfill \\ 0\hfill & (\lambda =0)\hfill \\ +\mathrm{}\hfill & (\lambda <0)\hfill \end{array}\text{as }|b|1,`$ where $`b=\lambda e^{2Q\phi }`$ and $`a^{}`$ denotes $`\frac{a}{\phi }`$. Therefore, the solution is valid only in the region satisfying $`|b|1`$, or equivalently $`\phi {\displaystyle \frac{1}{2Q}}\mathrm{log}|\lambda |,`$ (3.16) if $`\lambda 0`$. In the case $`\lambda =0`$, the solution is exactly the same as the “linear dilaton string” considered for $`d1`$. We investigate the $`\lambda 0`$ solution here. In the region satisfying (3.16), the solution is expanded as $`a^2(\phi )=a_0^2\{1+2\lambda e^{2Q\phi }\}+O(b^2),`$ (3.17) $`\mathrm{\Phi }(\phi )=Q\phi +2\lambda e^{2Q\phi }+O(b^2).`$ (3.18) We find that the above solution is consistent with the proposed action given in (3.3), if we set $`\mu ^{}=2\lambda `$ and rescale $`X`$ in (3.4) to $`\frac{X}{a_0}`$. Although the best method is to find a solution for the exact equation of motion, this seems to be very difficult. Thus we consider the physics only in the region defined by (3.16), where the perturbation with respect to $`\alpha ^{}`$ is valid, and we use the action $`S`$ $`=`$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle d^2\xi \sqrt{\widehat{g}}\{(1+2\lambda e^{2Q\phi })\widehat{g}^{ab}_aX^\mu _bX_\mu +\widehat{g}^{ab}_a\phi _b\phi +\alpha ^{}\widehat{R}(Q\phi +2\lambda e^{2Q\phi })\}}`$ (3.19) $`+S_{\mathrm{ghost}}`$ in this region. ### 3.3 Stabilization of the vacuum of $`𝝋`$ In this subsection, we attempt to stabilize the vacuum for $`\phi `$ in the action (3.19). We have two strategies. One is the standard stabilization using the minimum of the potential. However, we find that the potential in which we are interested has no minimum in some cases. The other one is to impose Dirichlet boundary conditions for $`\phi `$ and fix its zero mode. To use this method, we have to check the consistency of the Dirichlet boundary conditions and Weyl invariance. #### 3.3.1 Stabilization using the potential minimum In the case that Dirichlet boundary conditions are not imposed, the zero mode of $`\phi `$ must be stabilized at the minimum point of the potential. The vacua of $`X^\mu `$ are stable, and only the stabilization of $`\phi `$ is needed in our model. Thus, let us consider the effective action which we obtain after the path integration with respect to $`X^\mu `$, $$S(\phi )=\frac{1}{4\pi \alpha ^{}}d^2\xi \sqrt{\widehat{g}}\{(1+2\lambda e^{2Q\phi })\widehat{g}^{ab}_aX^\mu _bX_\mu +\widehat{g}^{ab}_a\phi _b\phi +\alpha ^{}\widehat{R}(Q\phi +2\lambda e^{2Q\phi })\},$$ (3.20) where $`𝒪`$ denotes the expectation value of the operator $`𝒪`$ obtained by the path integration with respect to $`X^\mu `$. The equation of motion with respect to the zero mode of $`\phi `$ is given by $`\delta S=\delta \phi _c{\displaystyle d^2\xi \sqrt{\widehat{g}}v^{}(\phi _c)}=0,`$ (3.21) where $`\phi _c`$ is the zero mode of $`\phi `$, and $`v(\phi _c)`$ is expressed by $`v(\phi _c)=\alpha ^{}\widehat{R}Q\phi +2\lambda e^{2Q\phi _c}(\alpha ^{}\widehat{R}+\widehat{g}^{ab}_aX^\mu _bX_\mu ).`$ (3.22) The Gauss-Bonnet theorem states that $`{\displaystyle _{}}d^2\xi \sqrt{\widehat{g}}\widehat{R}=4\pi \chi ,`$ (3.23) where $`\chi `$ is the Euler number of the world-sheet $``$. It is given by $`\chi =22N_gN_b`$, where $`N_g`$ is the number of the genus of $``$, and $`N_b`$ is the number of boundaries of $``$. Then the equation of motion for $`\phi _c`$ is $`4\pi \alpha ^{}Q\chi +4\lambda Qϵ^{2Q\phi _c}(4\pi \alpha ^{}\chi +A)=0,`$ (3.24) where $`A={\displaystyle _{}}d^2\xi \sqrt{\widehat{g}}\widehat{g}^{ab}_aX^\mu _bX_\mu >0.`$ (3.25) Since $`\lambda 0`$ and $`Q0`$, we obtain $`e^{2Q\phi _c}={\displaystyle \frac{4\pi \alpha ^{}\chi }{4\lambda (4\pi \alpha ^{}\chi +A)}}>0,`$ (3.26) and we immediately note that we have no solution for $`\chi =0`$. We also obtain $`v^{\prime \prime }(\phi _c)=8\pi \alpha ^{}Q^2\chi `$ (3.27) from (3.22) and (3.24), and we have no stable vacuum for $`\chi >0`$. Note that the leading string diagram in our model is a cylinder, for which we have $`\chi =0`$. In general, the leading term can be a disk ($`\chi =1`$) or a cylinder in the calculation of the correlation functions of Wilson loops. Thus, unfortunately, we do not have a stable vacuum suitable for these $`\chi 0`$ cases.<sup>8</sup><sup>8</sup>8We have the same problem for $`\chi 0`$ in DDK. However, in the calculation of string susceptibility, insertion of a $`\delta `$-function into the path integral, which keeps the world-sheet area constant, allows us to obtain the correct value for $`d1`$. #### 3.3.2 The Weyl invariant Dirichlet boundary conditions and stabilization of the vacuum of $`𝝋`$ We have seen in our model that we cannot make a consistent potential for $`\phi `$ with disk and cylinder diagrams. However, if we impose Dirichlet boundary conditions and fix the zero mode $`\phi _c`$, we are free of this difficulty. The problem in this case is the compatibility of the Dirichlet boundary conditions and the Weyl invariance of the world-sheet. In a linear dilaton string, Dirichlet boundary conditions break Weyl invariance . Therefore, we have to find a method to introduce Dirichlet boundary conditions into our theory without breaking Weyl invariance. A good place to start is to recall the origin of the Weyl-invariance breaking in a general dilatonic string with Dirichlet boundary conditions. In dilatonic string theory, we usually need a freedom of field redefinition to preserve Weyl invariance. For example, we have to make a constant shift (a field redefinition) of $`\phi `$, are in (2.15), to cancel the variation caused by the Weyl transformation (2.14) for $`d1`$. Dirichlet boundary conditions for $`\phi `$ forbids such a shift at the boundary, and breaks Weyl invariance. On the other hand, Dirichlet boundary conditions for $`X^\mu `$ do not break Weyl invariance, because no shift of $`X^\mu `$ is needed. However, we point out that we can employ Dirichlet boundary conditions for $`\phi `$ in a general dilatonic string if the criterion stated below is satisfied. In the theory with a dilaton, the required shift of the field is not a constant in general. For example, the shifts we need for the fields in (3.4) at the one-loop level are given as $`\delta \phi ={\displaystyle \frac{\alpha ^{}}{2}}_\phi \mathrm{\Phi }(\phi )\delta \sigma ,`$ (3.28) $`\delta X^\mu =0,`$ (3.29) where $`_\phi `$ stands for $`\frac{}{\phi }`$.<sup>9</sup><sup>9</sup>9The required field redefinition to preserve Weyl invariance appears in articles which discuss the $`\beta `$-functions for non-linear sigma models. For the models with boundaries, see Refs. and . The condition (3.28) for the shift is also found in Ref. . We can check that (3.28) gives the correct constant shift required in the linear dilaton case (2.15) if we set $`q=2`$.<sup>10</sup><sup>10</sup>10If $`q=2`$, we get $`u=\frac{1}{2}\frac{d25}{48\pi }`$. The denominator naturally coincides with the denominator of the factor in (2.8). In this linear dilaton case, the Weyl anomaly is obtained exactly at the one-loop level. We look deeper into the origin of the field redefinition in the Appendix. We note that we do not need field redefinition for $`\phi `$ if $`_\phi \mathrm{\Phi }=0`$. Therefore, we can use the Dirichlet boundary conditions $`\phi |_{}=\phi _0,`$ (3.30) where $`_\phi \mathrm{\Phi }(\phi )|_{\phi =\phi _0}=0,`$ (3.31) without breaking Weyl invariance at the one-loop level. The above stated criterion for consistent Dirichlet boundary conditions is one of the most important assertions of this article. As a next step, let us examine whether our model (3.19) has such a $`\phi _0`$ or not. Our dilaton term is $`\mathrm{\Phi }(\phi )=Q\phi +2\lambda e^{2Q\phi }+O(b^2).`$ (3.32) Thus the condition (3.31) is $`_\phi \mathrm{\Phi }=Q+4\lambda Qe^{2Q\phi }+O(b^2)=0,`$ (3.33) and we obtain $`\phi _0={\displaystyle \frac{1}{2Q}}\mathrm{log}(4\lambda )+O(b^2)`$ (3.34) for $`\lambda <0`$. The condition (3.33) is valid only in the region satisfying $`|b|=|\lambda e^{2Q\phi }|1`$, and now $`\lambda e^{2Q\phi _0}=1/4`$. Thus the above result suggests that we have an appropriate point $`\phi _0`$ for the Dirichlet boundary conditions (3.30) if $`\lambda <0`$. For this reason, we choose $`\lambda <0`$ and impose Dirichlet boundary conditions on the ends of the string to stabilize its configuration for arbitrary $`\chi `$. Note that the above argument naturally selects the branch of $`\lambda `$ uniquely. ### 3.4 Tachyon condensation: another role of Dirichlet boundary conditions in Liouville theory We found that we can use Dirichlet boundary condition to stabilize the string configuration, while preserving Weyl invariance for arbitrary $`\chi `$, at least up to $`O(\alpha ^2)`$. However, we have another big problem: If the target-space dimension ($`d+1`$ for our model) is greater than 2, we cannot fix all of the freedom of the world-sheet oscillation with gauge symmetry, and we have a physical oscillation. It is well known that we have a tachyonic state in the oscillation mode of the bosonic string in flat spacetime. In our model (3.19), the target space becomes asymptotically flat in the region $`\phi 0`$. Now we are considering the case $`d>1`$. Thus a tachyonic ground state appears in the region in which the spacetime is almost flat. If we have a tachyonic mode, we have to handle tachyon condensation. Tachyon condensation has been discussed by many authors (for example, see Ref. ), but this subject is difficult and has not yet been solved completely. We present some ideas to treat tachyon condensation here . We stress that Dirichlet boundary conditions play an important role in this subsection too. First, let us recall field condensation in usual quantum field theories. In field theory with field condensation (that is, in the theory with nonzero expectation value of the field ), we can calculate correct quantities if we know the correct expectation value of the field, even with the perturbation around an incorrect vacuum. For example, we can calculate the exact propagator with the perturbation around an incorrect vacuum by attaching tadpole diagrams to the tree propagator. Even though the mass squared is negative in a description around such a vacuum, tadpoles with appropriate weight create an additional shift of the mass squared, and make the total mass squared positive. In such a case, although the tachyonic mode exists in a perturbative theory around an incorrect vacuum, the theory is never wrong, and only the “vacuum” is wrong. In field theories, the true vacuum or exact expectation value of the field can be given by the Schwinger-Dyson equation. Thus we can get the correct weight of the tadpoles, we can calculate the correct propagator, and so on. We now make an analogy between field theories and tachyonic string theories. Although we have a tachyonic mode, we believe that the string theory is not fatally flawed, and the problem is that we do not know the true vacuum of it. The analogy with field theories tells us that we may be able to obtain correct quantities if we attach correct “tadpoles” to the world-sheet. Then, the question is what is the “tadpole” in string theories. We guess here that the tadpole in string theories is a macroscopic hole with Dirichlet boundary conditions in the world-sheet. We assume that the Dirichlet boundary condition for them is $`X^\mu (\xi _i)=a_i^\mu ,`$ (3.35) $`\phi (\xi _i)=\phi _0,`$ (3.36) where $`i`$ distinguishes each tadpole, $`a_i^\mu `$ is a constant, and $`\phi _0`$ is a constant which satisfies the condition (3.31). Of course, we have to consider the proper weights of the string wave functions on it. The above assumption results from the following considerations. The tachyonic tadpole is an off-shell state, because it does not carry momentum. We also know that off-shell states in string theory do not correspond to local emission vertexes. Thus we naturally assume that the tachyonic tadpole is a non-local macroscopic hole on the world-sheet.<sup>11</sup><sup>11</sup>11Some argument for the macroscopic hole as a tachyonic state is given in Ref. . We also have to preserve Weyl invariance, and it is natural to impose the above Dirichlet boundary conditions (3.35) and (3.36) on the edge of the hole. Neumann boundary conditions cannot be taken for a tadpole for the following reason. If we impose Neumann boundary conditions at a hole, the value of $`X^\mu `$ changes along the edge of the hole. This means that we observe a macroscopic hole even in $`d`$-dimensional spacetime. Although the Neumann boundary conditions for $`\phi `$ do not make a macroscopic hole in the visible $`d`$-dimensional spacetime, the hole can have momentum in the $`\phi `$ direction. This allows us to make an on-shell state, and the hole can break into on-shell open strings moving along the $`\phi `$ direction. These situations do not seem to be natural for our model. Contrastingly, the tadpole with the Dirichlet boundary conditions (3.35) and (3.36) does not leak any momentum from the world-sheet, and this gives a natural property for the tadpole. The macroscopic holes with Dirichlet boundary conditions on the world-sheet (or D-instantons in the target space) and the non-perturbative effects induced by them are discussed in Refs. and . However, we insist that the macroscopic holes discussed here play the role of “tadpoles” naturally even in Liouville theory. Unfortunately, we do not have the Schwinger-Dyson equation of string theory, and we do not know how to obtain the correct weight which should be attached to the tadpole. Thus, we cannot give a rigorous discussion to treat tachyon condensation, but we present a rough argument regarding tachyon condensation. To treat a macroscopic hole on the world-sheet is rather difficult, and we therefore approximate it as a point on the world-sheet which couples to the Dirichlet boundary conditions. In this case, the insertion of the tadpole is regarded as the insertion of $`h{\displaystyle _{}}d^2\xi _1\sqrt{\widehat{g}}\delta (X^M(\xi _1)a^M(\xi _1))`$ (3.37) into the world-sheet, where $`h`$ is the weight of the tadpole, and $`\xi _1`$ denotes the insertion point on the world-sheet. We have $`X^{d+1}=\phi `$ and $`a^{d+1}=\phi _0`$. In usual strings without a dilaton, the $`\delta `$-function in (3.37) becomes 1 after the integration over the moduli $`a^M`$, and the string propagator with the tadpoles can be estimated as propagator $`=`$ $`{\displaystyle \frac{1}{L_0+\overline{L_0}}}+{\displaystyle \frac{1}{L_0+\overline{L_0}}}h{\displaystyle \frac{1}{L_0+\overline{L_0}}}+{\displaystyle \frac{1}{L_0+\overline{L_0}}}h{\displaystyle \frac{1}{L_0+\overline{L_0}}}h{\displaystyle \frac{1}{L_0+\overline{L_0}}}+\mathrm{}`$ (3.38) $`=`$ $`{\displaystyle \frac{1}{L_0+\overline{L_0}h}},`$ where $`L_0`$ ($`\overline{L}_0`$) is the (anti)holomorphic part of the Hamiltonian of the corresponding conformal field theory. Thus, the insertion of the tadpoles (3.37) seems to make an additional shift to the energy of the tachyonic state. However, we cannot apply the above estimation directly to Liouville theory. Although we should integrate over $`a^\mu `$ to get Poincaré invariance in the $`d`$-dimensional spacetime, we never integrate over $`\phi _0`$ in our model (because it is fixed). Therefore, the expected non-perturbative effects induced by the tadpoles in Liouville theory seem to be different from those of non-dilatonic strings. In any case, we must develop a technique to estimate the effects of the insertion of Dirichlet boundaries. ## 4 Conclusion We attempted to quantize a noncritical (four dimensional) bosonic string as a natural candidate of a (large-$`N`$) pure QCD string. We considered the generalized Liouville action (3.4) as such a string. One of the main problems here is the stabilization of the Liouville mode $`\phi `$ while preserving Weyl invariance, and we found that we can stabilize it with the Dirichlet boundary conditions (3.30). The criterion for consistent Dirichlet boundary conditions at the one-loop level is given by (3.31), and the stabilized point is independent of the topology of the world-sheet. We found that the perturbative solution for the background in (3.4) seems to have a point which satisfies the condition (3.31). This argument led us to the unique selection of the branch of the solution. We also discussed tachyon condensation. Although the complete treatment of it is very difficult, we presented a simple strategy for it. The idea we presented is to attach tadpoles to the world-sheet, and we guessed that the tadpole in string theories might be represented as a macroscopic hole with Dirichlet boundary conditions. In non-dilatonic cases, a naive approximation for the propagator with tadpoles implies a shift of the energy of the tachyonic state. However, we guess that the non-perturbative effects in Liouville theory are different from those of non-dilatonic strings. Before we close this article, we stress again that Dirichlet boundary conditions have an important role in the generalized Liouville theory, and they can be imposed on the Liouville mode while preserving Weyl invariance if the appropriate condition mentioned above is satisfied. The investigation of Dirichlet strings in dilatonic backgrounds is very important, and it should yield necessary information about the construction of noncritical strings. Acknowledgments The author would like to thank Professor H.Kawai for valuable and inspiring discussion about the subjects presented here. ## Appendix A The Origin of the Field Redefinition and the Derivation of (3.28) Here we look deeper into the origin of the necessity of the field redefinition, at the one-loop level \[namely, up to $`O(\alpha ^2)`$ \]. Let us consider a general string action in $`d+1`$ dimensions, $`S`$ $`=`$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}d^2\xi \sqrt{\widehat{g}}\{G_{MN}(𝑿)\widehat{g}^{ab}_aX^M_bX^N+\alpha ^{}\widehat{R}\mathrm{\Phi }(𝑿)\}`$ (A.1) $`+{\displaystyle \frac{1}{2\pi \alpha ^{}}}{\displaystyle _{}}𝑑s\sqrt{\widehat{g}_{ss}}\alpha ^{}\widehat{k}\mathrm{\Phi }(𝑿(s)),`$ where we have included the boundary term, as it is needed in following calculation.<sup>12</sup><sup>12</sup>12For the non-linear sigma model with boundaries, see Refs. and . The renormalized action with dimensional regularization up to two loops is expressed as $`S`$ $`=`$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}d^{2+ϵ}\xi \sqrt{\widehat{g}}\left(G_{MN}(𝑿)+\alpha ^{}{\displaystyle \frac{1}{ϵ}}C_{MN}(𝑿)\right)\widehat{g}^{ab}_aX^M_bX^N`$ (A.2) $`+{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}d^{2+ϵ}\xi \sqrt{\widehat{g}}\alpha ^{}\widehat{R}\left(\mathrm{\Phi }(𝑿)+{\displaystyle \frac{1}{ϵ}}C_\mathrm{\Phi }(𝑿)\right)`$ $`+{\displaystyle \frac{1}{2\pi \alpha ^{}}}{\displaystyle _{}}d^{1+ϵ}s\sqrt{\widehat{g}_{ss}}\alpha ^{}\widehat{k}\left(\mathrm{\Phi }(𝑿(s))+{\displaystyle \frac{1}{ϵ}}\stackrel{~}{C}_\mathrm{\Phi }(𝑿(s))\right),`$ where the symmetric tensor $`\frac{1}{ϵ}C_{MN}`$ is the counterterm of the kinetic term, and the scalar $`\frac{1}{ϵ}C_\mathrm{\Phi }`$ ($`\frac{1}{ϵ}\stackrel{~}{C}_\mathrm{\Phi }`$) is the counterterm of the dilaton term in $``$ ($``$).<sup>13</sup><sup>13</sup>13We implicitly renormalized quadratically divergent terms into the cosmological constant term, and we set them to zero. If we perform the Weyl transformation $`\widehat{g}_{ab}\widehat{g}_{ab}e^{\delta \sigma }`$ and take the limit of $`ϵ0`$, the finite variation of the action at $`O(\delta \sigma )`$ is $`\delta _\sigma S`$ $`=`$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}d^2\xi \sqrt{\widehat{g}}\alpha ^{}\left\{{\displaystyle \frac{1}{2}}\delta \sigma C_{MN}(𝑿)\widehat{g}^{ab}_aX^M_bX^N+\widehat{R}{\displaystyle \frac{1}{2}}\delta \sigma C_\mathrm{\Phi }(𝑿)\right\}`$ (A.3) $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}d^2\xi \sqrt{\widehat{g}}\alpha ^{}\left(\mathrm{\Phi }(𝑿)+C_\mathrm{\Phi }(𝑿)\right)^2(\delta \sigma )`$ $`+{\displaystyle \frac{1}{2\pi \alpha ^{}}}{\displaystyle _{}}𝑑s\sqrt{\widehat{g}_{ss}}\alpha ^{}\widehat{k}{\displaystyle \frac{1}{2}}\delta \sigma \stackrel{~}{C}_\mathrm{\Phi }(𝑿(s))`$ $`+{\displaystyle \frac{1}{2\pi \alpha ^{}}}{\displaystyle _{}}𝑑s\sqrt{\widehat{g}_{ss}}\alpha ^{}{\displaystyle \frac{1}{2}}(\widehat{n}^a_a(\delta \sigma ))\left(\mathrm{\Phi }(𝑿(s))+\stackrel{~}{C}_\mathrm{\Phi }(𝑿(s))\right),`$ where $`\widehat{n}^a`$ is the unit outward normal vector. The term which contains $`^2(\delta \sigma )`$ in (A.3) is rewritten as $``$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}d^2\xi \sqrt{\widehat{g}}\widehat{g}^{ab}\alpha ^{}\{_a_bX^M_M(\mathrm{\Phi }(𝑿)+C_\mathrm{\Phi }(𝑿))`$ (A.4) $`+_aX^M_bX^N_M_N\mathrm{\Phi }(𝑿)\}\delta \sigma `$ $``$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}𝑑s\sqrt{\widehat{g}_{ss}}\alpha ^{}(\widehat{n}^a_a(\delta \sigma ))\left(\mathrm{\Phi }(𝑿(s))+C_\mathrm{\Phi }(𝑿(s))\right)`$ $`+`$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}𝑑s\sqrt{\widehat{g}_{ss}}\alpha ^{}\widehat{n}^a_aX^M_M\left(\mathrm{\Phi }(𝑿(s))+C_\mathrm{\Phi }(𝑿(s))\right)\delta \sigma ,`$ where $`M`$ and $`N`$ are the indices of the coordinates of spacetime, and $`_N`$ stands for $`\frac{}{X^N}`$. Therefore we get $`\delta _\sigma S`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _{}}d^2\xi \sqrt{\widehat{g}}\{({\displaystyle \frac{1}{2}}C_{MN}(𝑿)_M_N\mathrm{\Phi }(𝑿))\widehat{g}^{ab}_aX^M_bX^N`$ (A.5) $`+\widehat{R}{\displaystyle \frac{1}{2}}C_\mathrm{\Phi }(𝑿)\}\delta \sigma `$ $`+{\displaystyle \frac{1}{2\pi }}{\displaystyle _{}}𝑑s\sqrt{\widehat{g}_{ss}}\widehat{k}{\displaystyle \frac{1}{2}}\stackrel{~}{C}_\mathrm{\Phi }(𝑿(s))\delta \sigma `$ $`+{\displaystyle \frac{1}{4\pi }}{\displaystyle _{}}𝑑s\sqrt{\widehat{g}_{ss}}(\widehat{n}^a_a(\delta \sigma ))\left(\stackrel{~}{C}_\mathrm{\Phi }(𝑿(s))C_\mathrm{\Phi }(𝑿(s))\right)`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle _{}}d^2\xi \sqrt{\widehat{g}}\widehat{g}^{ab}_a_bX^M_M\left(\mathrm{\Phi }(𝑿)+C_\mathrm{\Phi }(𝑿)\right)\delta \sigma `$ $`+{\displaystyle \frac{1}{4\pi }}{\displaystyle _{}}𝑑s\sqrt{\widehat{g}_{ss}}\widehat{n}^a_aX^M_M\left(\mathrm{\Phi }(𝑿(s))+C_\mathrm{\Phi }(𝑿(s))\right)\delta \sigma .`$ The last three terms cannot be absorbed into any counterterm. The third term of (A.5) goes to zero if $`\stackrel{~}{C}_\mathrm{\Phi }=C_\mathrm{\Phi }`$. This is realized if all the $`\beta `$-functions, except for that of the dilaton, vanish . The problem is to determine how to deal with the last two terms. Fortunately, we can cancel them using field redefinition. If we perform a field redefinition $`X^MX^M+\delta X^M`$, the variation of the action is $`\delta _XS`$ $`=`$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}d^2\xi \sqrt{\widehat{g}}\mathrm{\hspace{0.33em}2}\widehat{g}^{ab}_aX^M_b\delta X_M`$ (A.6) $`+(\text{terms proportional to }\delta X_M)`$ $`=`$ $`{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}d^2\xi \sqrt{\widehat{g}}\mathrm{\hspace{0.33em}2}\widehat{g}^{ab}_a_bX^M\delta X_M`$ $`+{\displaystyle \frac{1}{4\pi \alpha ^{}}}{\displaystyle _{}}𝑑s\sqrt{\widehat{g}_{ss}}2\widehat{n}^a_aX^M\delta X_M`$ $`+(\text{terms proportional to }\delta X_M).`$ Therefore, if we set $`\delta X^M={\displaystyle \frac{\alpha ^{}}{2}}^M(\mathrm{\Phi }(𝑿)+C_\mathrm{\Phi }(𝑿))\delta \sigma ,`$ (A.7) we can cancel the last two terms of (A.5). The remaining terms proportional to $`\delta X_M`$ in (A.6) can be absorbed into the counterterms. Thus, after a proper field redefinition, $`\delta _\sigma S+\delta _XS`$ contains only terms proportional to the $`\beta `$-functions, and we can preserve Weyl invariance if we set each of the $`\beta `$-functions to zero. We know that the counterterm $`C_\mathrm{\Phi }`$ at the one-loop level corresponds to the central charge, and is a constant. Thus the required shift of $`X^M`$ is $`\delta X^M={\displaystyle \frac{\alpha ^{}}{2}}^M\mathrm{\Phi }(𝑿)\delta \sigma .`$ (A.8) Now we emphasize the very important fact that we do not need field redefinition at the special point $`𝑿_0`$ where $`^M\mathrm{\Phi }(𝑿_0)=0`$. In our model (3.19), $`G_{MN}`$, $`\mathrm{\Phi }`$ and the counter terms depend only on $`X^{d+1}=\phi `$, so that (A.8) can be written as $`\delta \phi ={\displaystyle \frac{\alpha ^{}}{2}}^\phi \mathrm{\Phi }(\phi )\delta \sigma `$ (A.9) $`\delta X^\mu =0,`$ (A.10) where $`\mu `$ runs from 1 to $`d`$ and $`^\phi =^{d+1}`$. Now $`G^{\phi \phi }=1`$, and $`^\phi \mathrm{\Phi }=_\phi \mathrm{\Phi }`$. Thus we obtain (3.28).
warning/0004/hep-th0004031.html
ar5iv
text
# ADM Worldvolume Geometry ## Abstract We describe the dynamics of a relativistic extended object in terms of the geometry of a configuration of constant time. This involves an adaptation of the ADM formulation of canonical general relativity. We apply the formalism to the hamiltonian formulation of a Dirac-Nambu-Goto \[DNG\] relativistic extended object in an arbitrary background spacetime. The mechanics of relativistic extended objects, or branes for short, has been the subject of various investigations in recent years. In particular, the usefulness of a geometrical approach which mantains general covariance with respect to worldvolume reparametrizations and rotations of worldvolume normal vector fields has become apparent (see e.g. ). However, this geometrical approach has been developed largely within the lagrangian formulation of theories of relativistic extended objects. The geometry of interest is then the worldvolume geometry. In this note, we report on an alternative approach where we focus on the geometry of the relativistic extended object itself, and its relation with the worldvolume geometry . One immediate application of this formalism is a geometrical Hamiltonian formulation of this type of theories, which mantains worldvolume covariance without resorting to gauge fixing from the outset. (For alternative treatments see e.g. ). We consider a relativistic extended object $`\mathrm{\Sigma }`$, of dimension $`d`$, embedded in an arbitrary fixed $`(N+1)`$-dimensional background spacetime $`\{M,g_{\mu \nu }\}`$. $`\mathrm{\Sigma }`$ is described locally by the spacelike embedding $`x^\mu =X^\mu (u^A)`$, where $`x^\mu `$ are local coordinates for the background spacetime, $`u^A`$ local coordinates for $`\mathrm{\Sigma }`$, and $`X^\mu `$ the embedding functions ($`\mu ,\nu ,\mathrm{}=0,1,\mathrm{},N`$, and $`A,B,\mathrm{}=1,\mathrm{},d`$). The tangent vectors to $`\mathrm{\Sigma }`$ are defined by $`ϵ_A:=(_AX^\mu )_\mu `$, so that the positive-definite metric induced on $`\mathrm{\Sigma }`$ is $$h_{AB}:=g(ϵ_A,ϵ_B).$$ (1) We construct out of the metric $`h_{AB}`$ the intrinsic geometry of $`\mathrm{\Sigma }`$. Note that it will be trivial in the special case of a relativistic string, since then $`\mathrm{\Sigma }`$ is one-dimensional. For the extrinsic geometry of $`\mathrm{\Sigma }`$, we introduce the normals $`\{m^I\}`$ , defined by $`g(ϵ_A,m^I)=0`$, and normalized with $`g(m^I,m^J)=\eta ^{IJ}`$, where $`\eta _{IJ}`$ is the Minkowski metric, with only one minus sign ($`I,J,\mathrm{}=0,1,\mathrm{},N+1d`$). The extrinsic curvature along the $`I`$-th normal is $$L_{AB}{}_{}{}^{I}:=g(m^I,D_Aϵ_B)=L_{BA}{}_{}{}^{I},$$ (2) where $`D_A=ϵ^\mu {}_{A}{}^{}D_{\mu }^{}`$ is the spacetime covariant derivative compatible with $`g_{\mu \nu }`$ along the tangent directions. In addition, the extrinsic geometry of $`\mathrm{\Sigma }`$ is determined by the extrinsic twist, $`\sigma _A^{IJ}`$, defined by $$\sigma _A{}_{}{}^{IJ}:=g(m^J,D_Am^I)=\sigma _A{}_{}{}^{JI}.$$ (3) This is a connection associated with the $`O(N+1d)`$ freedom in the definition of the normal fields. In the mathematical literature it is known as the normal form. When the appropriate Gauss-Codazzi-Mainardi integrability conditions hold, $`\{h_{AB},L_{AB}{}_{}{}^{I},\sigma _A{}_{}{}^{IJ}\}`$ define the geometry of $`\mathrm{\Sigma }`$. We consider now the time evolution of $`\mathrm{\Sigma }`$ in spacetime. We denote its trajectory, or worldvolume, by $`w`$. It is an oriented timelike surface in spacetime. Now the shape functions become time-dependent, $`X^\mu =X^\mu (t,u^A)`$, where $`t`$ is a coordinate that labels the leafs of the foliation of $`w`$ by $`\mathrm{\Sigma }`$s. From the point of view of an observer sitting on $`\mathrm{\Sigma }`$, this involves breaking the normal rotation symmetry of $`\mathrm{\Sigma }`$, $`O(N+1d)`$, down to $`O(Nd)`$, by choosing the unit (future-pointing) timelike normal to $`\mathrm{\Sigma }`$ into $`w`$, $`m^0=:\eta `$. We denote by $`m^i=:n^i`$ the remaining components of $`\{m^I\}`$ ($`i,j,\mathrm{}=1,2,\mathrm{},Nd`$). These are also normals to the worldvolume $`w`$. This symmetry reduction is the key ingredient in this geometrical construction. The time evolution of the embedding functions for $`\mathrm{\Sigma }`$ into the worldvolume can be written as, $$\dot{X}^\mu :=N\eta ^\mu +N^Aϵ^\mu {}_{A}{}^{},$$ (4) where, following standard usage in general relativity, $`N`$ is called the lapse function, and $`N^A`$ the shift vector. We emphasize that the content of this equation is simply that the time evolution of $`\mathrm{\Sigma }`$ is into the worldvolume $`w`$. Note that we can always chose a time evolution normal to $`\mathrm{\Sigma }`$, i.e. take the shift vector to vanish, $`N^A=0`$ (see e.g. ). At this point, we introduce the geometry of the worldvolume $`w`$. The worldvolume can be represented in parametric form by the embedding functions $`x^\mu =\chi ^\mu (\xi ^a)`$, where $`\xi ^a=\{t,u^A\}`$ are local coordinates for $`w`$, and $`\chi ^\mu `$ the embedding functions ($`a,b,\mathrm{}=0,1,\mathrm{},d`$). The tangent vectors to $`w`$, $`e_a:=e^\mu {}_{a}{}^{}_{\mu }^{}=(_a\chi ^\mu )_\mu `$, decompose in a part tangential to $`\mathrm{\Sigma }`$ and a part along the time evolution of $`\mathrm{\Sigma }`$, $$e^\mu {}_{a}{}^{}=\left(\begin{array}{c}\dot{X}^\mu \hfill \\ ϵ^\mu _A\hfill \end{array}\right).$$ (5) It follows that the lorentzian worldvolume induced metric, $`\gamma _{ab}:=g(e_a,e_b)`$ , decomposes according to the familiar ADM expression as, $$\gamma _{ab}=\left(\begin{array}{cc}N^2+N^AN^Bh_{AB}\hfill & h_{AB}N^B\hfill \\ h_{AB}N^B\hfill & h_{AB}\hfill \end{array}\right).$$ (6) Note that the worldvolume element is given by $`\sqrt{\gamma }=N\sqrt{h}`$. The various geometrical quantities that characterize the intrinsic geometry of $`\mathrm{\Sigma }`$, such as its Riemann curvature etc., can be decomposed by importing the appropriate expressions from the ADM treatment of spacetime in canonical general relativity. It is worth emphasizing the difference between, say, the lapse function in this context versus the lapse function in canonical general relativity. Whereas here it is a component of the velocity, in the latter case it is a component of the metric. The extrinsic curvature of the worldvolume $`w`$ along the $`i`$-th normal vector field $`\{n^i\}`$ is defined by $$K_{ab}{}_{}{}^{i}:=g(n^i,D_ae_b)=K_{ba}^i,$$ (7) where $`D_a=e^\mu {}_{a}{}^{}D_{\mu }^{}`$ is the gradient along the vectors tangential to $`w`$. It can be decomposed as $$K_{ab}{}_{}{}^{i}=\left(\begin{array}{cc}n_\mu {}_{}{}^{i}\ddot{X}_{}^{\mu }\hfill & H_A^i\hfill \\ H_A^i\hfill & L_{AB}^i\hfill \end{array}\right),$$ (8) where we have introduced the quantity $$H_A{}_{}{}^{i}:=N\sigma _A{}_{}{}^{0i}+N^BL_{AB}^i.$$ (9) We note that the time-time component of the extrinsic curvature is (minus) the projection into the normals of the acceleration of $`\mathrm{\Sigma }_t`$. For the degenerate case of a relativistic particle, this is all there is. The off-diagonal components involve the extrinsic twist of $`\mathrm{\Sigma }`$. This is a consequence of having broken the full normal rotation symmetry. The spatial components involve the extrinsic curvature of $`\mathrm{\Sigma }`$ along the normals $`\{n^i\}`$ The ADM decomposition of the mean extrinsic curvature of the worldvolume, $`K^i:=\gamma ^{ab}K_{ab}^i`$, is readily obtained as $`K^i`$ $`=`$ $`N^2[n_\mu {}_{}{}^{i}\ddot{X}_{}^{\mu }+2N^AH_A^i`$ (10) $``$ $`N^AN^BL_{AB}^i+N^2L^i],`$ where $`L^i=h^{AB}L_{AB}^i`$ is the mean extrinsic curvature of $`\mathrm{\Sigma }`$ along the normals to the worldvolume $`\{n^i\}`$. This is the generalization to an extended object of the acceleration for a relativistic particle. We can put the formalism we have developed to use in the analysis of the dynamics of a relativistic extended object. For example, for a simple DNG object, which extremizes the worldvolume, the equations of motion are given by the vanishing of the mean extrinsic curvature, $$K^i=0.$$ (11) If we specialize to normal time evolution, use of the expression (10) allows us to rewrite this equation in the form, $$n_\mu {}_{}{}^{i}\ddot{X}_{}^{\mu }+N^2L^i=0.$$ (12) This identifies a part of the extrinsic curvature of $`\mathrm{\Sigma }`$ as the driving force in its dynamics. We emphasize that these expressions hold in an arbitrary background spacetime. The natural application of this geometrical approach is the Hamiltonian formulation of a theory of relativistic extended objects. For simplicity, we consider here only the case of a DNG object. Although the benefits of a geometrical approach are already apparent in this case, the full power of the formalism is displayed when considering higher-order curvature dependent actions , or additional worldvolume fields as for example in the case of superconducting membranes . We write the DNG action as $`S=𝑑tL[X,\dot{X}]`$, where, with $`\mu `$ the brane tension, the lagrangian is $`L[X,\dot{X}]`$ $`=`$ $`\mu {\displaystyle _\mathrm{\Sigma }}d^du\sqrt{\gamma }`$ (13) $`=`$ $`\mu {\displaystyle _\mathrm{\Sigma }}d^duN\sqrt{h}.`$ Therefore, the lapse function plays the role of the lagrangian density for a DNG brane. The canonical momenta are given by $`P_\mu :=\delta L/\delta \dot{X}^\mu `$. Using the definition (4) for the lapse function, we find $$P_\mu =\mu \sqrt{h}\eta _\mu .$$ (14) The factor of $`\sqrt{h}`$ comes as no surprise – momenta are densities. The “shape” functions of $`\mathrm{\Sigma }`$ are the configuration variables, and their conjugate momenta are proportional to the unit normal of $`\mathrm{\Sigma }`$ into the worldvolume. The phase space for our extended object is then naturally associated with the geometry of $`\mathrm{\Sigma }`$. As expected from worldvolume reparametrization invariance, the hamiltonian vanishes, $`H[X,P]=d^du\left[P_\mu \dot{X}^\mu \right]L[X,\dot{X}]=0`$. However, according to the standard Dirac treatment of constrained systems, the hamiltonian is a linear combination of the phase space constraints that generate reparametrizations $`\{𝒞_0,𝒞_A\}`$, The explicit form of the constraints is easily obtained from the definition of the momenta as $`𝒞_A(X,P)`$ $`=`$ $`P_\mu ϵ^\mu {}_{A}{}^{}=0,`$ (15) $`𝒞_0(X,P)`$ $`=`$ $`P^2+\mu ^2h=0,`$ (16) and the hamiltonian is $$H[X,P]=_{\mathrm{\Sigma }_t}[\lambda 𝒞_0+\lambda ^A𝒞_A],$$ (17) where $`\lambda `$, $`\lambda ^A`$ are lagrange multipliers. Note that the lagrange multiplier $`\lambda `$ must be a scalar density of weight $`1`$ for the hamiltonian to be well-defined. The first constraint, or vector constraint, is universal for all reparametrizations invariant actions. Clearly it has no analogue in the case of a relativistic particle, when $`\mathrm{\Sigma }`$ degenerates to a point. It is easy to see that it is the generator of spatial diffeomorphisms. The second, or scalar constraint, is the generalization to extended objects of the familiar hamiltonian constraint for a parametrized massive relativistic particle. It generates the evolution of $`\mathrm{\Sigma }`$. For a relativistic string, it is possible to exploit the triviality of its one-dimensional $`\mathrm{\Sigma }`$, by choosing proper length as the parameter along the string, so that $`h=1`$, the potential is constant, and the scalar constraint takes the form $`P^2+\mu ^2=0`$. This simple observation explains the integrability of the string in a variety of background spacetimes (see e.g. ). (It also strongly suggests that the string may be integrable in any background spacetime.) Already for a membrane ($`d=2`$), however, such a priviledged parametrization for $`\mathrm{\Sigma }`$ is not available. The determinant of $`h_{AB}`$ contains four powers of (spatial derivatives of) $`X^\mu `$, and the theory is highly non-linear, even in a flat background. Hopes of integrability are evidently slim. Introducing the bracket $$\{X^\mu ,X^\nu \}=ϵ^{AB}ϵ^\mu {}_{A}{}^{}ϵ_{}^{\nu }{}_{B}{}^{},$$ (18) we can rewrite the scalar constraint for a relativistic membrane in the form $$P^2+\mu ^2\{X^\mu ,X^\nu \}^2=0.$$ (19) In this form we can make contact with current approaches to the canonical formulations of membrane dynamics (see ). The constraints are first class. The Poisson algebra of the constraint functions reproduce the familiar algebra of hypersurface deformation for $`\mathrm{\Sigma }`$, now seen as a hypersurface in the worldvolume $`w`$ . In general it is not a Lie algebra since the structure “constants” depend on the configuration variables via the (densitized) metric $`hh^{AB}`$ in the scalar-scalar Poisson bracket. This is familiar from the constraint algebra of general relativity. However, for a relativistic string, the triviality of the intrinsic geometry of $`\mathrm{\Sigma }`$ allows to set $`hh^{AB}=1`$, and we have a true Lie algebra, which can be shown to be isomorphic to the algebra of the conformal group in two dimensions . Let us consider the Hamilton equations. We have that $$\dot{X}^\mu =2\lambda P^\mu +\lambda ^Aϵ^\mu {}_{A}{}^{}.$$ (20) It reproduces the definition of the momentum (14), and it identifies $`\lambda ^A=N^A`$, $`\lambda =N/2\mu \sqrt{h}`$, assuming that $`P_\mu `$ is future-pointing. Using this information, the second Hamilton equation is $`\dot{P}_\mu `$ $`=`$ $`{\displaystyle \frac{1}{2}}h^1\dot{h}P_\mu N\mu \sqrt{h}L^in_{\mu i}`$ (21) $`+`$ $`\mu \sqrt{h}(𝒟_AN)ϵ_\mu {}_{}{}^{A}+𝒟_A(N^AP_\mu ),`$ where, for convenience, we have left $`\dot{h}`$ in its implicit form. Note that $`\dot{h}=2Nhk`$, where $`k=h^{AB}L_{AB}^0`$ is the mean extrinsic curvature of $`\mathrm{\Sigma }`$ embedded as a hypersurface in $`w`$. For the analysis of this equation it is convenient to take $`N^A=0`$, and the shift constant $`𝒟_AN=0`$. The projection along the momentum is a mere identity, $$\dot{P}_\mu P^\mu =\frac{1}{2}h^1\dot{h}P^2,$$ (22) as can be seen using the scalar constraint, and its time derivative. The normal projection $$\dot{P}_\mu n^\mu {}_{i}{}^{}=N\mu \sqrt{h}L_i,$$ (23) then reproduces the equations of motion in the form (11). This geometrical approach to the canonical formulation of relativistic extended objects provides valuable insight into their dynamics, and when it is relevant, into their canonical quantization. Furthermore, placing their description on the same footing as canonical general relativity has the potential to reap mutual benefits, both technically and conceptually.
warning/0004/cond-mat0004484.html
ar5iv
text
# 1 Introduction ## 1 Introduction We continue our study of the XX-model with boundaries defined by the Hamiltonian $$H=\frac{1}{2}\underset{j=1}{\overset{L1}{}}\left[\sigma _j^+\sigma _{j+1}^{}+\sigma _j^{}\sigma _{j+1}^+\right]+\frac{1}{\sqrt{8}}\left[\alpha _{}\sigma _1^{}+\alpha _+\sigma _1^++\alpha _z\sigma _1^z+\beta _+\sigma _L^++\beta _{}\sigma _L^{}+\beta _z\sigma _L^z\right]$$ (1) where $`\sigma _j^\pm =\frac{1}{2}(\sigma _j^x\pm \mathrm{}\sigma _j^y)`$. By $`\sigma ^x,\sigma ^y`$ and $`\sigma ^z`$ we denote the Pauli matrices. Since we restrict ourselves to hermitian boundary terms, the values of $`\alpha _z`$ and $`\beta _z`$ are real numbers and $`\alpha _+`$ and $`\beta _+`$ are the complex conjugates of $`\alpha _{}`$ respectively $`\beta _{}`$. In the first publication of this series we have shown how to diagonalize $`H`$ by introducing an auxiliary Hamiltonian $`H_{\mathrm{long}}={\displaystyle \frac{1}{2}}{\displaystyle \underset{j=1}{\overset{L1}{}}}\left[\sigma _j^+\sigma _{j+1}^{}+\sigma _j^{}\sigma _{j+1}^+\right]`$ $`+{\displaystyle \frac{1}{\sqrt{8}}}\left[\alpha _{}\sigma _0^x\sigma _1^{}+\alpha _+\sigma _0^x\sigma _1^++\alpha _z\sigma _1^z+\beta _+\sigma _L^+\sigma _{L+1}^x+\beta _{}\sigma _L^{}\sigma _{L+1}^x+\beta _z\sigma _L^z\right]`$ (2) (following an idea of ) which in turn may be diagonalized in terms of free fermions . Note that $`H_{\mathrm{long}}`$ commutes with $`\sigma _0^x`$ and $`\sigma _{L+1}^x`$. Hence the spectrum of $`H_{\mathrm{long}}`$ decomposes into four sectors $`(+,+),(+,),(,),(,+)`$ corresponding to the eigenvalues $`\pm 1`$ of $`\sigma _0^x`$ and $`\sigma _{L+1}^x`$. The spectrum and the eigenvectors of $`H`$ are obtained by projecting onto the $`(+,+)`$–sector. In this paper we are going to consider the profiles of the spin operators $`\sigma _j^z`$ and $`\sigma _j^x`$. We have already shown in how to compute the one point function of the $`\sigma _j^x`$-operator in terms of pfaffians. The $`\sigma _j^z`$-profiles are easier to compute. We will study the ground-state profiles as well as the profiles of certain excited states of $`H`$. The excited states of $`H`$ we are going to consider correspond to single fermion excitations with respect to the ground-state of $`H_{\mathrm{long}}`$, which are characterized by a non-vanishing excitation energy as $`L`$ goes to infinity and which correspond to boundary bound states. We are able to obtain exact results for the ground-state profiles only for certain values of the boundary parameters . In the case of the $`\sigma _j^z`$-profiles we will find exact expressions on the finite chain, in the continuum limit $`L1,z=j/L`$ fixed and in the limit $`L\mathrm{},j1`$. The $`\sigma _j^x`$-profiles are given in terms of pfaffians of $`2j\times 2j`$ matrices, that we have not been able to compute on the finite chain or in the continuum limit exactly, therefore our exact results for the $`\sigma _j^x`$-profiles are restricted to the limit $`L\mathrm{},j1`$. A combination of extensive numerical computations (for lattice lengths $`L600`$ ) and of our exact results has led us to conjectures for the profiles of $`\sigma _j^z`$ and $`\sigma _j^x`$ in the continuum limit for more general values of the boundary parameters. Note that once we know the $`\sigma _j^x`$-profile for one specific choice of the boundary terms, we know already the profile of $`\sigma _j^y`$ for the boundary terms with $`\sigma ^x`$ and $`\sigma ^y`$ interchanged in (1). We will discuss our results for the ground-state profiles in the context of conformal field theory. According to Burkhardt and Xue , in the continuum limit $`z=j/L,L1`$ the general form of the profile of any scalar scaling operator $`\varphi `$ with bulk-scaling dimension $`x_\varphi `$ is given by $$\varphi (j)_{ab}=[(L/\pi )\mathrm{sin}(\pi z)]^{x_\varphi }_{ab}(z).$$ (3) The functional form of $`_{ab}(z)`$ depends on the boundary conditions $`a`$ and $`b`$ at the left respectively the right boundary. We have shown in that the XX–model with hermitian boundaries corresponds to the free compactified boson on a cylinder with Neumann-Neumann, Dirichlet-Neumann or Dirichlet-Dirichlet boundary conditions. The bulk-scaling dimensions $`x_\varphi `$ obtained from the two-point functions for the periodic chain are $`1`$ respectively $`\frac{1}{4}`$ for $`\sigma _j^z`$ respectively $`\sigma _j^x`$ . Conformal field theory makes more restrictive predictions for the behaviour of the profiles of scaling operators near the boundary. From (3) we obtain that near the boundary the profile is given by $$\varphi (j)_{ab}=j^{x_\varphi }\psi _{ab}(z)z1.$$ (4) It was already argued in that if the scaling function $`\psi _{ab}(z)`$ is different from zero near the boundary, it has the asymptotic form $$\psi _{ab}(z)=A[1+_{ab}^\varphi z^2+\mathrm{}]z1.$$ (5) Furthermore it has been shown in that the amplitude $`_{ab}^\varphi `$ is related to the Casimir amplitude $`𝒜_{ab}`$ via $$_{ab}^\varphi /𝒜_{ab}=4\pi x_\varphi /c$$ (6) where $`c`$ denotes the central charge of the theory. The Casimir amplitude $`𝒜_{ab}`$ is defined by the large $`L`$ behaviour of the ground-state energy $`E_0`$, i.e. $$E_0=e_{\mathrm{}}L+f_{ab}+𝒜_{ab}L^1+\mathrm{}$$ (7) where by $`e_{\mathrm{}}`$ and $`f_{ab}`$ we denote the bulk respectively the surface free energy. The Casimir amplitudes for the boundary terms of the Hamiltonian (1) can be read off from the partition functions obtained in . As we are going to show, when dealing with boundaries, it is not a priori clear which combinations of spin operators correspond to the scaling operators of the continuum theory. Our conjectures for the continuum limit of the profiles will yield, that the profiles of $`\sigma _j^x`$ and of $`\sigma _j^z`$ indeed have the form of (3) (see however equation (35)). We will see that if the boundary conditions we are going to consider correspond to Dirichlet-Neumann of the free boson field, the profiles satisfy (5) and (6). We will also see that for the boundary terms corresponding to Dirichlet-Dirichlet and Neumann-Neumann boundary conditions the profiles show a more complicated behaviour near the boundary. In these cases the equations (5) and (6) do not hold in general for the profiles of $`\sigma _j^x`$ and $`\sigma _j^z`$. In the case of Dirichlet-Dirichlet boundary conditions we have to build an appropriate linear combination of $`\sigma _j^z`$ and $`\sigma _{j+1}^z`$ in order to obtain the correct behaviour of a scaling operator. However, in the generic case the linear combination we are going to obtain works only at one boundary at a time. The $`\sigma _j^x`$-profile vanishes exactly in this case. For the case of Neumann-Neumann boundary conditions we obtain the correct behaviour by considering a linear combination of $`\sigma _j^x`$ and $`\sigma _j^y`$. Again this works only at one boundary at a time in the generic case. Here the $`\sigma _j^z`$-profile vanishes if no diagonal boundary terms are present at the boundary (i.e. $`\alpha _z=\beta _z=0`$). However, we have also studied a type of boundaries corresponding to the Neumann-Neumann boundary conditions, where diagonal and non-diagonal boundaries are present at a boundary at the same time. In this case the profile of the $`\sigma _j^z`$–operator vanishes only in the leading order and is determined by a secondary field instead by a primary field. Our results in the limit $`L\mathrm{},j1`$ are in agreement with the results of Affleck . He has considered the semi-infinite XXZ spin-1/2 chain with a transverse magnetic field applied at the end of the chain (this corresponds to $`\alpha _+=\alpha _{}0`$, $`\alpha _z=\beta _z=\beta _+=\beta _{}=0`$ and $`L\mathrm{}`$ in our notation). Affleck has used the bosonization technique, whereas our results have been obtained on the lattice. Assuming Neumann boundary conditions for the boson field, he obtained the $`\sigma _j^x`$-profile for large values of $`j`$. At the free fermion point the result reads $$\sigma _j^x=(1)^jCj^{\frac{1}{4}}$$ (8) where the constant $`C`$ does not depend on the strength of the field and at the free fermion point its numerical value is given by $$C=\mathrm{0.912\hspace{0.17em}171\hspace{0.17em}210\hspace{0.17em}446}\mathrm{}.$$ (9) The case of a longitudinal magnetic field at the end of the chain has also been considered in (corresponding to $`\alpha _z0`$, $`\alpha _+=\alpha _{}=\beta _z=\beta _+=\beta _{}=0`$ and $`L\mathrm{}`$). Imposing Dirichlet boundary conditions for the boson field, the large $`j`$ behaviour of the $`\sigma _j^z`$-profile has been evaluated for small values of $`\alpha _z`$, yielding at the free fermion point $$\sigma _j^zB\alpha _z(1)^jj^1$$ (10) where the value of $`B`$ is undetermined. The case of diagonal and non-diagonal boundary terms at the boundary at the same time has not been considered. The boundary conditions where we have been able to compute the profiles exactly in the limit $`L\mathrm{},j1`$ can always be mapped onto the boundaries treated by Affleck. Our results obtained on the lattice agree with (8) and (10). For the $`\sigma _j^z`$-profiles we have not restricted ourselves to small values of $`\alpha _z`$ and we have derived the $`\alpha _z`$-dependence of $`B`$. As already mentioned above, we will also consider the magnetization profiles of certain eigenstates of $`H`$, which are obtained by the excitation of a massive fermion with respect to the ground-state of $`H_{\mathrm{long}}`$. The existence of such fermions has already been mentioned in for special values of the boundary parameters. Before considering the profiles, we are going to study the appearance of such fermions for general, hermitian boundary terms. We will see that massive fermions appear if the values of certain functions of the boundary parameters become greater than a certain threshold. The respective excitation masses will turn out to depend only on the boundary parameters of one boundary at a time. In the general case (non-diagonal together with diagonal boundary terms) we will find at most one mass per boundary. In the case of purely diagonal boundary terms the fermion masses correspond exactly to the non-vanishing energy gaps obtained in the Bethe Ansatz by comparing the energy of the reference state with the energy of the one-magnon excitations, which correspond to boundary 1-strings . We will see, that the magnetization profiles for the excited states under study here, show an exponential fall-off into the bulk for diagonal as well as for non-diagonal boundary conditions (the profiles for the non-diagonal boundary terms have been obtained numerically). Hence we are going to refer to these states as to boundary bound states. These profiles appear also for $`H`$ since the excited states we are going to study lie in the $`(+,+)`$-sector of the Hilbert space of $`H_{\mathrm{long}}`$ and do therefore correspond directly to eigenstates of $`H`$. This paper is organized as follows: We start with the exact results in section 2. These results are restricted to certain choices of boundary terms. In section 3 we will give our conjectures for the profiles in the continuum limit for more general types of boundaries. We are going to discuss these results in the context of conformal field theory in section 4. Boundary bound states are the subject of section 5. Our conclusions will be given in section 6. The details of our calculations will be given in the appendix. In Appendix A we shortly discuss the diagonalization of $`H_{\mathrm{long}}`$ and the projection onto the $`(+,+)`$-sector in order to provide the basic facts and notations being necessary to follow our calculations. Appendix B deals with the exact computations of the $`\sigma _j^z`$-profiles. How we arrived at our exact results for the $`\sigma _j^x`$-profiles is shown in Appendix C. In Appendix D we discuss the numerical verification of our conjectures of the ground-state profiles in the continuum limit. ## 2 Exact results For technical reasons explained in A, we have been able to obtain analytical results only for certain choices of boundary terms (see table 1 and table 2). We will begin with the results for the $`\sigma _j^z`$-profiles on the finite chain, before turning to the respective expressions in the continuum limit. Thereafter we will give our exact results in the limit $`L\mathrm{},j1`$ for the profiles of $`\sigma _j^x`$ and $`\sigma _j^z`$. ### 2.1 Exact results on the finite lattice ($`\sigma _j^z`$\- profiles) We have computed the exact profiles of $`\sigma _j^z`$ for values of the boundary parameters given in table 1 <sup>2</sup><sup>2</sup>2The cases denoted by a,b,c,d,e correspond to the cases $`2,4,11,14,16`$ in the notation of .. Note that for technical reasons (see B for details) our result for case $`e`$ is restricted to the choice $`\delta =0`$ in this section. Furthermore, observe that the ground-state is twofold degenerate for case $`d`$ with $`L`$ odd and for case $`e`$ with $`L`$ even, due to the existence of additional zero modes . In these cases, we have obtained the profiles for the two ground-states, which have a well-defined fermion number (see B). These profiles are indicated by a $`\pm `$-sign in the respective expressions. * Case a $$\sigma _j^z=\frac{1}{2L+2}\left((1)^L\mathrm{tan}\frac{\pi }{4L+4}+(1)^j\mathrm{cot}\frac{(2j1)\pi }{4L+4}\right).$$ (11) * Case b $$\sigma _j^z=\frac{(1)^j}{2L+1}\mathrm{cot}\frac{(2j1)\pi }{4L+2}.$$ (12) * Case c $$\sigma _j^z=\frac{1}{2L+1}\left((1)^L+(1)^j\left(\mathrm{sin}\frac{(2j1)\pi }{4L+2}\right)^1\right).$$ (13) * Case d, $`L`$ odd: The ground-state is twofold degenerate in this case. $$\sigma _j^z_\pm =\frac{1}{L}\left((1)^j\mathrm{cot}\frac{(2j1)\pi }{2L}\pm 1\right).$$ (14) * Case d, $`L`$ even $$\sigma _j^z=\frac{(1)^j}{L}\left(\mathrm{sin}\frac{(2j1)\pi }{2L}\right)^1.$$ (15) * Case e, $`L`$ odd, $`\delta =0`$ : The same as given for case d, $`L`$ even. * Case e, $`L`$ even, $`\delta =0`$ : The same as given for case d, $`L`$ odd. The ground-state is twofold degenerate in this case, too. ### 2.2 Exact results in the continuum limit The continuum limit defined by $`L1,z=j/L`$ fixed of the expressions (11)-(15) is readily obtained. Furthermore we have been able to compute the profile in the continuum limit for case e for general values of $`\delta `$ (see B). * Case a and case b $$\sigma _j^z=(1)^j\left(L\mathrm{sin}(\pi z)\right)^1\mathrm{cos}^2\left(\frac{\pi z}{2}\right)+\mathrm{O}(1/L^2).$$ (16) * Case c $$\sigma _j^z=(1)^j\left(L\mathrm{sin}(\pi z)\right)^1\mathrm{cos}\left(\frac{\pi z}{2}\right)+(1)^L/2L+\mathrm{O}(1/L^2).$$ (17) * Case d: The same expression as given for case e with $`\delta =0`$ and $`L`$ even and $`L`$ odd interchanged. * Case e, $`L`$ odd $$\sigma _j^z=(1)^j\mathrm{cos}\delta \left(L\mathrm{sin}(\pi z)\right)^1+\mathrm{O}(1/L^2).$$ (18) * Case e, $`L`$ even $$\sigma _j^z_\pm =(1)^j\mathrm{cos}(\delta \pi z)\left(L\mathrm{sin}(\pi z)\right)^1\pm 1/L+\mathrm{O}(1/L^2).$$ (19) ### 2.3 Exact results in the limit $`L\mathrm{},j1`$ We have obtained exact results for the profiles of $`\sigma _j^z`$ in the limit $`L\mathrm{},j1`$ for the five cases given in table 1. For the cases a,b,c,d the result may be read of from the expressions obtained on the finite lattice. This can not be done for case e with arbitrary values of $`\delta `$, since the respective result on the finite lattice is restricted to $`\delta =0`$. However, in the limit $`L\mathrm{},j1`$ the calculation for case e can be done for arbitrary values of $`\delta `$ (see B). Using $`\alpha _z`$ instead of the parameter $`\delta `$ (see table 1 for the relation between $`\delta `$ and $`\alpha _z`$) we have obtained $$\sigma _j^z=(1)^j\frac{\sqrt{8}\alpha _z}{(1+2\alpha _z^2)\pi }j^1+\mathrm{}.$$ (20) The profiles for the cases a,b,c,d are also given by this expression if one replaces $`\alpha _z`$ by $`\frac{1}{\sqrt{2}}`$. The result (20) has to be compared to (10), which yields $`B`$ as a function of $`\alpha _z`$, i.e. $$B=\frac{\sqrt{8}}{(1+2\alpha _z^2)\pi }.$$ (21) We have obtained exact results for the profiles of $`\sigma _j^x`$ in the limit $`L\mathrm{}`$ and $`j1`$ for the two cases given in table 2<sup>3</sup><sup>3</sup>3In the notation of these cases correspond to case 2 respectively case 9.. Our computations for case f (see C.2) yield $$\sigma _j^x=(1)^jAj^{\frac{1}{4}}+\mathrm{}$$ (22) where $`A`$ is given by $$A=\mathrm{}^{1/4}2^{7/12}C_G^3$$ (23) and $`C_\mathrm{G}`$ is Glaishers constant , which can be approximately given by $$C_\mathrm{G}=\mathrm{1.282\hspace{0.17em}427\hspace{0.17em}129\hspace{0.17em}100\hspace{0.17em}62}\mathrm{}.$$ (24) For case g we have obtained (see C.1) $$\sigma _j^x=(1)^jA\mathrm{cos}\phi j^{\frac{1}{4}}+\mathrm{}.$$ (25) Comparing the value of $`C`$ in (9) and the value of $`A`$ in (23) using (24) shows that our results (22) and (25) for $`\phi =0`$ are in agreement with (8). At this point we want to draw the readers attention to an interesting by-product we have obtained during our computations for case g. It is of major interest for our conjecture of the continuum limit of the $`\sigma _j^x`$ profiles in the case of Neumann-Neumann boundary conditions presented in the next section. We have obtained that the ground-state profile of $`\sigma _j^x`$ for arbitrary values of $`\phi `$ and $`\chi `$ (see table 2) is given by $$\sigma _j^x=\mathrm{cos}\left(\phi +\frac{\chi +2m\pi }{L+1}j\right)f(j,L)$$ (26) where $`f(j,L)`$ does not depend on $`\phi ,\chi `$ and may be computed numerically (see C.1). The value of $`m`$ depends on the value of $`\chi `$ and on the lattice length $`L`$. For odd $`L`$ we have $`m=0`$ for $`\pi <\chi <\pi `$. If $`\chi =\pi `$ the ground-state of $`H`$ is twofold degenerate. The value of $`m`$ is either $`m=0`$ or $`m=1`$. For even $`L`$ the value of $`m`$ is $`\frac{1}{2}`$ for $`0>\chi >\pi `$ and $`\frac{1}{2}`$ for $`0<\chi \pi `$. Here the ground-state is twofold degenerate for $`\chi =0`$ the value of $`m`$ being either $`\frac{1}{2}`$ or $`\frac{1}{2}`$. ## 3 Conjectures Our results given so far allow to conjecture the ground-state profiles in the continuum limit for more general values of the boundary parameters. We have checked our results by computing the profiles on the finite chain numerically for lattice lengths up to $`L=700`$. Thereafter we have used the finite size data to extrapolate the profiles for $`L\mathrm{}`$ and fixed values of $`z=j/L`$. We have performed these extrapolations for $`150`$ different values of the boundary parameters each with $`19`$ (L even) respectively $`20`$ (L odd) values of $`z`$. The accuracy of our extrapolations varies with the choice of the boundary parameters. In most of the cases the relative deviation from our conjectures presented in the following has been of the order of $`10^{12}10^7`$ (see D for details). We have considered the following types of boundary terms: * Dirichlet-Dirichlet: $$\alpha _\pm =\beta _\pm =0\alpha _z=\frac{1}{\sqrt{2}}\mathrm{tan}\left(\frac{\omega +\delta }{2}\right)\beta _z=\frac{1}{\sqrt{2}}\mathrm{tan}\left(\frac{\omega \delta }{2}\right)$$ (27) Note that for $`\omega =\frac{\pi }{2}`$ we recover the boundary terms of case e in table 1. * Dirichlet-Neumann: Here we have considered only a special type of boundary terms, i.e. $$\beta _+=\beta _{}0\beta _z=\alpha _\pm =0\alpha _z\text{ free}$$ (28) * Neumann-Neumann: We have studied two types of boundary terms, i.e. $$\alpha _z=\beta _z=0\alpha _\pm =R_\alpha \mathrm{}^{\pm \mathrm{}\phi }\beta _\pm =R_\beta \mathrm{}^{\pm \mathrm{}(\chi +\phi )}$$ (29) respectively $$\alpha _z=\beta _z\alpha _+=\alpha _{}=\beta _+=\beta _{}0.$$ (30) Note that the boundary terms (29) are identical to the boundaries of case g in table 2 if one sets $`R_\alpha =R_\beta =1`$, whereas no case in table 1 and table 2 corresponds to boundary terms which satisfy (30). In this case our result is based on purely numerical computations. Observe also, that this is the only type of boundary terms with diagonal and non-diagonal terms at one end of the chain, which is studied in this paper. ### 3.1 Dirichlet-Dirichlet boundary conditions For this type of boundary conditions the ground-state profile of $`\sigma _j^x`$ vanishes if the ground-state is non-degenerate, since the Hamiltonian commutes with $`\sigma _1^z\sigma _2^z\mathrm{}\sigma _L^z`$. Therefore we have restricted ourselves to the ground-state profiles of $`\sigma _j^z`$. Our results for the cases c,d,e from table 1 and our numerical investigations suggest the $`\sigma _j^z`$-profiles for general values of $`\alpha _z`$ and $`\beta _z`$, i.e. $$\sigma _j^z=(1)^j\frac{\mathrm{sin}(\omega +\delta 2(\omega +Q\pi )z)}{L\mathrm{sin}(\pi z)}+\frac{2(\omega +Q\pi )}{\pi L}.$$ (31) This equation is valid for $`L`$ even and $`L`$ odd. The value of $`Q`$ depends on the values of $`\omega `$ and the lattice length $`L`$. This dependence is given in table 3. The discontinuities are due to level crossings in the spectrum at the respective values of $`\omega `$. At these points ($`\omega =0`$ for $`L`$ odd and $`\omega =\pm \pi /2`$ for $`L`$ even) the ground-state is twofold degenerate. ### 3.2 Dirichlet-Neumann boundary conditions Our numerical computations suggest the following profile of the $`\sigma _j^x`$-operator: $$\sigma _j^x=\pi ^{1/4}A(1)^{L+1j}\left[\mathrm{sin}\left(\frac{\pi z}{2}\right)\right]^{\frac{1}{2}}\left(L\mathrm{sin}(\pi z)\right)^{\frac{1}{4}}$$ (32) where $`A`$ is given by (23). For two cases corresponding to this type of boundaries we have already considered the exact $`\sigma _j^z`$-profiles, namely for case a and b from table 1. The given expression (16) suggests universal behaviour with respect to the variation of the value of $`\beta _\pm `$. Our numerical calculations for various values of $`\beta _\pm `$ confirmed our guess. Furthermore they yield the dependence of the profile on the second parameter $`\alpha _z`$. We have obtained $$\sigma _j^z=(1)^j\frac{\sqrt{8}\alpha _z}{1+2\alpha _z^2}\mathrm{cos}^2\left(\frac{\pi z}{2}\right)\left(L\mathrm{sin}(\pi z)\right)^1.$$ (33) Note that the form of the profile does not depend on the value of $`\alpha _z`$. However, observe also that the amplitude does. For $`\alpha _z=0`$ the profile vanishes, since then the Hamiltonian $`H`$ commutes with $`\sigma _1^x\sigma _2^x\mathrm{}\sigma _L^x`$ (the ground-state of $`H`$ is non-degenerate ). ### 3.3 Neumann-Neumann boundary conditions Our exact result for case g from table 2, in particular equation (26), in combination with numerical computations yields the $`\sigma _j^x`$-profile for the first type of boundaries in question (see (29)): $$\sigma _j^x=\pi ^{1/4}A(1)^j\mathrm{cos}\left(\phi +(\chi +2m\pi )z\right)\left(L\mathrm{sin}(\pi z)\right)^{\frac{1}{4}}$$ (34) where the value of $`m`$ depends on the values of $`\chi `$ and the lattice length $`L`$ as already discussed at the end of section 2.3. The $`\sigma _j^z`$ profiles vanish, as long as the ground-state is non-degenerate (see B for details). We did not consider the $`\sigma _j^z`$ profiles where the ground-state is degenerate. For the second type of boundary terms (30) the $`\sigma _j^z`$ profile does not vanish. For odd values of $`L`$ our numerical investigations suggest $$\sigma _j^z=(1)^j\pi \sqrt{2}\frac{\alpha _z}{\alpha _+^2}\mathrm{cos}(\pi z)(L\mathrm{sin}(\pi z))^2.$$ (35) We have not been able to find an analytic expression for the profile for even values of $`L`$. For $`\alpha _+=0`$ the prefactor in (35) diverges. However, in this case the profile is already given by (31). For $`\alpha _z=0`$ the profile vanishes as in the case of the Dirichlet-Neumann boundary condition (28), since in this case the Hamiltonian $`H`$ has a higher symmetry. It commutes with $`\sigma _1^x\sigma _2^x\mathrm{}\sigma _L^x`$ (the ground-state of $`H`$ is also non-degenerate here ). Comparing (35) with (3) yields that the leading term of the $`\sigma _j^z`$-profile is zero as for the boundary terms (29), since according to the long range behaviour of the two-point function the bulk-scaling dimension $`x_\varphi `$ associated to the $`\sigma _j^z`$ operator is $`1`$. Since there exists no spinless primary field with bulk-scaling dimension $`2`$ we conclude that the profile (35) is determined by a secondary field (the profile of a primary field with spin vanishes ). We have checked the $`\sigma _j^x`$-profile only for odd $`L`$. It is not affected by the presence of the diagonal terms in (30) and is already given by (34) using $`\phi =\chi =m=0`$. ## 4 Comparison with the predictions of Conformal Field Theory In this section we are going to consider the behaviour of the profiles we have conjectured near the boundaries, i.e. we consider the limit $`z1`$ (respectively $`(1z)1`$) , in order to check for the validity of the equations (5) and (6). ### 4.1 Dirichlet-Dirichlet boundary conditions The behaviour of (31) near the left boundary is readily obtained: $$\sigma _j^z=(1)^j\frac{\mathrm{sin}(\omega +\delta )}{j\pi }\left(1\mathrm{cot}(\omega +\delta )2\omega ^{}z+\left(\frac{1}{6}\frac{2\omega ^2}{\pi ^2}\right)(\pi z)^2+\mathrm{}\right)+\frac{2\omega ^{}}{\pi L}.$$ (36) where $`\omega ^{}=\omega +Q\pi `$. The behaviour of the profile near the right boundary may be obtained by exchanging $`\delta `$ by $`\delta `$ and performing the transformation $`jL+1j`$. This can be seen from (27). Note that (36) does not agree with equation (5), since in (5) there appears no term which is linear in $`z`$. This problem is avoided by considering the profile of the following linear combination of $`\sigma _j^z`$ and $`\sigma _{j+1}^z`$ for odd values of $`j`$: $$\mathrm{tan}\left(\frac{\omega +\delta }{2}\right)\sigma _j^z\mathrm{cot}\left(\frac{\omega +\delta }{2}\right)\sigma _{j+1}^z=\frac{2}{j\pi }\left(1+\left(\frac{1}{6}\frac{2\omega ^2}{\pi ^2}\right)(\pi z)^2+\mathrm{}\right).$$ (37) The value of the Casimir amplitude $`𝒜_{ab}`$ can be obtained from the results in . It is given by $`𝒜_{ab}=\frac{\pi }{24}+\frac{\omega ^2}{2\pi ^2}`$. Hence (37) is in agreement with (6). Note that this operator does not yield the correct scaling behaviour near the right boundary, since the given linear combination of $`\sigma _j^z`$ and $`\sigma _{j+1}^z`$ depends on the value of $`\delta `$, which enters the left boundary term in a different way than the right boundary term (see (27)). Another linear combination can be chosen to obtain the correct behaviour near the right boundary. ### 4.2 Dirichlet-Neumann boundary conditions The Casimir amplitude for this type of boundary conditions is independent from the explicit values of the boundary parameters. Its value is given by $`𝒜_{ab}=\frac{\pi }{48}`$ . The profiles (32) and (33) vanish near the left respectively near the right boundary and therefore the predictions (5) and (6) are not valid. Near the right boundary we obtain from (32) $$\sigma _{L+1j}^x=(1)^jAj^{\frac{1}{4}}\left(1\frac{(\pi z)^2}{48}+\mathrm{}\right)$$ (38) whereas (33) yields near the left boundary $$\sigma _j^z=(1)^j\frac{\sqrt{8}\alpha _z}{1+2\alpha _z^2}(j\pi )^1\left(1\frac{(\pi z)^2}{12}+\mathrm{}\right).$$ (39) Both expressions are in agreement with the predictions (5) and (6). ### 4.3 Neumann-Neumann boundary conditions From (34) we obtain the profiles of $`\sigma _j^x`$ near the left boundary, i.e. $`\sigma _j^x=(1)^jAj^{\frac{1}{4}}\mathrm{cos}\phi `$ $`\times \left(1\mathrm{tan}(\phi )(\chi +2m\pi )z+\left({\displaystyle \frac{\pi ^2}{24}}{\displaystyle \frac{(\chi +2m\pi )^2}{2}}\right)z^2+\mathrm{}\right).`$ (40) The Casimir amplitude has been shown in to be $`𝒜_{ab}=\frac{\pi }{24}+2\pi \left(m+\frac{\chi }{2\pi }\right)^2`$. Note that the profile is in disagreement with the predicted behaviour for a true scaling operator given in (5). In order to obtain the correct behaviour near the boundary, we have to consider an appropriate linear combination of $`\sigma _j^x`$ and $`\sigma _j^y`$. At the left boundary we obtain $`\mathrm{cos}\phi \sigma _j^x\mathrm{sin}\phi \sigma _j^y=(1)^jAj^{\frac{1}{4}}\left(1+\left({\displaystyle \frac{\pi ^2}{24}}{\displaystyle \frac{(\chi +2m\pi )^2}{2}}\right)z^2+\mathrm{}\right)`$ (41) which is in agreement with (5) and (6). Note that building the given linear combination is equivalent to performing a rotation around the z-axis such that $`\phi =0`$ in (29). However, similar to the case of the $`\sigma _j^z`$ profile for the Dirichlet-Dirichlet boundary condition, this linear combination does not lead to the correct behaviour near the right boundary. To obtain the correct behaviour near the right boundary one has to perform a different rotation as long as $`\chi 0`$. It is interesting to consider the respective profile of the orthogonal linear combination of $`\sigma _j^x`$ and $`\sigma _j^y`$, which yields $`\mathrm{sin}\phi \sigma _j^x+\mathrm{cos}\phi \sigma _j^y=(1)^jAj^{\frac{1}{4}}\left((\chi +2m\pi )z+\mathrm{}\right).`$ (42) Note that this profile vanishes near the left boundary and therefore the predictions (5) and (6) do not apply. ## 5 Boundary bound states As already mentioned in the introduction, we have also studied the profiles of certain excited states of $`H`$. These states are obtained by the excitation of a massive fermion with respect to the ground-state of $`H_{\mathrm{long}}`$. We refer to these excitations as to boundary excitations, since we are going to see, that they give rise to a contribution to the profile which decays exponentially into the bulk. First, we are going to look for massive excitations for general, hermitian boundary terms in section 5.1 by studying the secular equation $`p(x^2)=0`$ which determines the fermion energies and which has been obtained in . In section 5.2 we will study the magnetization profiles near the boundaries of the states, where one massive fermion is excited, for two special types of boundary terms. In the case of diagonal boundary terms the energy gaps defined with respect to the ground-state energy are in one to one correspondence to the boundary 1-strings found as solutions of the Bethe Ansatz equations. The 1-strings define the energy gaps between the energy of the reference state and the eigenstates in the one magnon sector. At the free fermion point a boundary 1-string appears if $`2\alpha _z^2>1`$ or $`2\beta _z^2>1`$ . The energy gaps corresponding to the boundary 1-strings in the limit $`L\mathrm{}`$ are given by $$E_\alpha =\frac{1}{2}\left(\frac{1}{\sqrt{2}\alpha _z}+\sqrt{2}\alpha _z\right)E_\beta =\frac{1}{2}\left(\frac{1}{\sqrt{2}\beta _z}+\sqrt{2}\beta _z\right)$$ (43) where by $`E_\alpha `$ and $`E_\beta `$ we denote the energy gaps corresponding to the boundary 1-string which appears for $`2\alpha _z^2>1`$ respectively $`2\beta _z^2>1`$. ### 5.1 Massive excitations The fermion energies are obtained by $$2\mathrm{\Lambda }=\frac{1}{2}(x+1/x)$$ (44) where the values of $`x`$ are given by the roots of a polynomial $`p(x^2)`$ . Since the energies have to be real numbers, the roots of the polynomial lie either on the unit circle or on the real axis. We will focus on the roots which lie on the real axis for $`L\mathrm{}`$. Due to (44) these roots yield fermions with a finite energy in this limit. We have not been able to determine whether there exist massive fermions corresponding to roots lying on the unit circle. We are going to restrict ourselves to the search for zeros satisfying $`|x|<1`$, since if $`x`$ is a zero of $`p(x^2)`$ so is $`1/x`$. In the limit $`L\mathrm{}`$, the secular equation $`p(x^2)=0`$ reduces to $$\left(1+(12\alpha _z^22\alpha _{}\alpha _+)x^22\alpha _z^2x^4\right)\left(1+(12\beta _z^22\beta _{}\beta _+)x^22\beta _z^2x^4\right)=0.$$ (45) Solving for $`x^2`$ yields two solutions with $`|x^2|<1`$, i.e. $$x^2=\frac{1}{4\alpha _z^2}\left[12\alpha _{}\alpha _+2\alpha _z^2+\sqrt{8\alpha _z^2+(2\alpha _z^2+2\alpha _{}\alpha _+1)^2}\right]$$ (46) and $$x^2=\frac{1}{4\beta _z^2}\left[12\beta _{}\beta _+2\beta _z^2+\sqrt{8\beta _z^2+(2\beta _z^2+2\beta _{}\beta _+1)^2}\right].$$ (47) Note that these solutions depend only on the parameters of one boundary at a time. The fermion energies are obtained by taking the positive square root of (46) respectively (47) and using (44). Hence we conclude that the maximum number of massive excitations obtained by our ansatz is two (one per boundary). This result is interesting since it confirms an analogous result obtained by Ameduri et al in their study of the boundary sine-Gordon theory at the free fermion point. If the value of $`\alpha _z`$ or $`\beta _z`$ equals zero, the expression (46) respectively (47) has to be replaced by $$x^2=1/(2\alpha _{}\alpha _+1)\text{respectively}x^2=1/(2\beta _{}\beta _+1).$$ (48) If $`\alpha _{}=0`$ or $`\alpha _+=0`$ respectively $`\beta _{}=0`$ or $`\beta _+=0`$ the result simplifies to $$x^2=1/(2\alpha _z^2)\text{respectively}x^2=1/(2\beta _z^2).$$ (49) All solutions of (45) are only asymptotic solutions of the original polynomial if they satisfy the assumption $`|x|<1`$. In the case of diagonal boundary terms this condition coincides with the condition for the existence of boundary 1-strings in the Bethe Ansatz. The energies of the one-magnon excitations corresponding to the boundary 1-strings given in (43) correspond exactly to the fermion energies (44) obtained by the roots in (49). ### 5.2 Magnetization profiles near the boundary We have studied the magnetization profiles for certain states, which differ from the ground-state of $`H_{\mathrm{long}}`$ by a boundary excitation, for two types of boundaries. Note that the ground-state of $`H_{\mathrm{long}}`$ is at least twofold degenerate. We have chosen it such that the excited state in question lies in the $`(+,+)`$-sector. Therefore these states correspond also to eigenstates of $`H`$. First we have considered diagonal boundary terms $$\alpha _z=1/(2\beta _z)=\mathrm{}^\xi /\sqrt{2}\alpha _\pm =\beta _\pm =0$$ (50) for even and odd values of $`L`$. Without loss of generality we have restricted ourselves to $`\xi >0`$. In this case $`2\alpha _z^2`$ becomes larger than 1 and according to (49) and (44) there exists one boundary excitation with energy $`2\mathrm{\Lambda }=\mathrm{cosh}\xi `$. We are able to compute the $`\sigma _j^z`$-profile analytically in the limit $`L\mathrm{},j1`$ (see B), i.e. $$\xi \sigma _j^z\xi \frac{(1)^j}{\pi }\frac{1}{\mathrm{cosh}\xi }j^1+2(\mathrm{}^{2\xi }1)\mathrm{}^{2j\xi }$$ (51) where $`|\xi `$ denotes the state obtained by the excitation of the massive fermion. Observe that the correlation length $`1/(2\xi )`$ is related to the excitation mass $`2\mathrm{\Lambda }=\mathrm{cosh}\xi `$. The profile near the right boundary is $$\xi \sigma _{L+1j}^z\xi \frac{(1)^j}{\pi }\frac{1}{\mathrm{cosh}\xi }j^1$$ (52) as expected. The $`U(1)`$ charge, which is given by the projection of the total spin onto the z-axis, of the state $`\xi `$ is $`0`$ for $`L`$ even and $`\frac{1}{2}`$ for $`L`$ odd. Notice that the exponential part in (51) is not present if one considers the respective ground-state profile (see (20)). We conclude, that adding a massive excitation to the ground-state yields an additive contribution to the profile which decays exponentially into the bulk. The second type of boundaries we have considered are non-diagonal and given by $$\alpha _\pm =\sqrt{\mathrm{cosh}\xi _\alpha \mathrm{exp}\xi _\alpha }\beta _\pm =\sqrt{\mathrm{cosh}\xi _\beta \mathrm{exp}\xi _\beta }\alpha _z=\beta _z=0.$$ (53) We restricted ourselves to even values of $`L`$, since in this case for odd values of $`L`$ the excitation of a fermion with respect to the ground-state of $`H_{\mathrm{long}}`$ does not lead to a state which lies in the $`(+,+)`$-sector and hence does not correspond to an eigenstate of $`H`$ (see C.3 for details). Since the $`\sigma _j^z`$-profile vanishes exactly for the boundaries (53), we have studied the $`\sigma _j^x`$-profile in this case. In contrast to the $`\sigma _j^z`$-profiles given in (51) and (52) our results for the $`\sigma _j^x`$-profiles we are going to present are based on numerical computations for $`L=\mathrm{}`$ in the special case $`\xi _\beta =\xi _\alpha `$ and for $`L=800`$ in the general case. In we have shown that the $`\sigma _j^x`$-profiles for the boundaries given in (53) may be computed in terms of determinants of $`j\times j`$ matrices. In the case $`\xi _\beta =\xi _\alpha `$ it is possible to compute these matrices analytically in the limit $`L\mathrm{}`$ (see C.3). Thus the numerical study of the profile near the boundaries is easier than in the general case. Here one boundary excitation appears with energy $`2\mathrm{\Lambda }=\mathrm{cosh}\xi _\alpha `$ (see (48) and (44)). For $`\xi _\alpha >0`$ we have obtained by numerical extrapolation from data for $`j=1\mathrm{}100`$ $$\xi _\alpha \sigma _j^x\xi _\alpha (1)^{j+1}Aj^{\frac{1}{4}}\left(12\mathrm{}^{(1)^j\xi _\alpha }\mathrm{}^{2j\xi _\alpha }\right)$$ (54) and $$\xi _\alpha \sigma _{L+1j}^x\xi _\alpha (1)^jAj^{\frac{1}{4}}$$ (55) where $`A`$ is given by (23) and $`\xi _\alpha `$ denotes the state obtained by the excitation of the fermion with energy $`2\mathrm{\Lambda }=\mathrm{cosh}\xi _\alpha `$. Again the exponential term is absent in the respective expressions for the ground-state profile. Note again the relation between the correlation length and the excitation mass. Inspired by this result we have also studied the case $`\xi _\alpha \xi _\beta `$. If $`\xi _\alpha ,\xi _\beta >0`$ there exist two boundary excitations with energies $`2\mathrm{\Lambda }_\alpha =\mathrm{cosh}\xi _\alpha `$ and $`2\mathrm{\Lambda }_\beta =\mathrm{cosh}\xi _\beta `$ in the limit $`L\mathrm{}`$. We have considered the respective magnetization profiles on the finite chain numerically ($`L=800`$). Exemplarily the figures 1 and 2 show the profiles we have obtained numerically for the state $`\xi _\alpha `$ near the left and the right boundary, where we have chosen $`\xi _\alpha =0.075,\xi _\beta =0.05`$. Our results suggest that as long as $`\xi _\alpha \xi _\beta `$ the profile for the state $`\xi _\alpha `$ is still given by (54) and (55) (the deviation from the expression in (54) for large $`j`$ in figure 1 is expected due to the presence of the right boundary). The profile for the state $`\xi _\beta `$ (obtained by the excitation of the fermion with energy $`2\mathrm{\Lambda }_\beta =\mathrm{cosh}\xi _\beta `$) near the left boundary is just $$\xi _\beta \sigma _j^x\xi _\beta (1)^jAj^{\frac{1}{4}}$$ (56) whereas at the right boundary we have $$\xi _\beta \sigma _{L+j1}^x\xi _\beta (1)^{j+1}Aj^{\frac{1}{4}}\left(12\mathrm{}^{(1)^j\xi _\beta }\mathrm{}^{2j\xi _\beta }\right).$$ (57) However, if $`\xi _\alpha =\xi _\beta =\xi `$ our computations suggest a different behaviour, i.e. $$\xi _\alpha \sigma _j^x\xi _\alpha =\xi _\beta \sigma _j^x\xi _\beta (1)^jAj^{\frac{1}{4}}\mathrm{}^{(1)^j\xi }\mathrm{}^{2j\xi }$$ (58) near the left boundary and $$\xi _\alpha \sigma _{L+1j}^x\xi _\alpha =\xi _\beta \sigma _{L+1j}^x\xi _\beta (1)^jAj^{\frac{1}{4}}\mathrm{}^{(1)^j\xi }\mathrm{}^{2j\xi }$$ (59) near the right boundary. In contrast to the profiles we have obtained so far, these profiles are left-right symmetric. Furthermore the leading term proportional to $`j^{\frac{1}{4}}`$ does not appear, unlike in equations (54) and (57). Note that the Hamiltonian with boundary conditions (53) conserves parity if $`\xi _\alpha =\xi _\beta `$. Our numerical computations yield that the energies $`2\mathrm{\Lambda }_\alpha `$ and $`2\mathrm{\Lambda }_\beta `$ are non-degenerate on the finite chain. Only in the limit $`L\mathrm{}`$ the degeneracy appears. Hence the parity of the states $`\xi _\alpha `$ and $`\xi _\beta `$ is well-defined. This explains why the profiles reflect the symmetry of the Hamiltonian. ## 6 Conclusions In this paper we have studied the one point functions of the $`\sigma _j^z`$ and $`\sigma _j^x`$ operator for the XX-chain with hermitian boundaries, which in the continuum limit corresponds to a free boson field on a cylinder with Dirichlet-Dirichlet, Dirichlet-Neumann or Neumann-Neumann boundary conditions . No results for the magnetization profiles in the ground-state have previously been obtained on the lattice. The only known results have been found by means of the bosonization technique and were restricted to the limit $`L\mathrm{}`$ and $`j1`$ (see (8) and (10)). The expression (10) is valid only for small values of the longitudinal boundary field, i.e. for small values of $`\alpha _z`$ in our notation. We have obtained the exact ground-state profiles of the $`\sigma _j^z`$-operator on the finite chain for the boundary terms given in table 1. The respective profiles are given in (11)-(15). The continuum limit of the profiles is given in (16)-(19). The $`\sigma _j^z`$-profiles for the cases from table 1 in the limit $`L\mathrm{},j1`$ are given in (20). Exact results for the $`\sigma _j^x`$-profiles in this limit have been found for the cases in table 2 and are given in (22) respectively (25). Our results obtained on the lattice agree to the expressions (8) and (10) which have been found using the bosonization technique . Note that in contrast to (10) our result (20) is not restricted to small values of the boundary field. Supported by numerical computations, these exact results have led us to conjectures for the profiles in the continuum limit for more general values of the boundary parameters (see (27)-(30)). For the Dirichlet-Dirichlet case we have studied the most general boundary terms (27). The $`\sigma _j^z`$-profile for these boundary terms is given in (31) (the $`\sigma _j^x`$ profile vanishes exactly, if the ground-state is non-degenerate). We also considered a type of boundary terms, which correspond to the Dirichlet-Neumann boundary condition (see (28)). The $`\sigma _j^x`$ profile we conjectured is given in (32). The profile of the $`\sigma _j^z`$ operator is given in (33). The boundary terms given by (29) and (30) correspond to Neumann-Neumann boundary conditions on the boson field. The $`\sigma _j^x`$-profile for the first type of boundaries (29) is given in (34). Here the $`\sigma _j^z`$-profile vanishes exactly, if the ground-state is non-degenerate. For the second type of boundary terms (30) the profile of the $`\sigma _j^z`$ operator is given in (35). As in the case of the Dirichlet-Neumann boundary condition the $`\sigma _j^x`$ profile is not affected by the presence of the diagonal boundary terms. The expressions we have obtained for the profiles in the case of the Dirichlet-Dirichlet respectively the Neumann-Neumann boundary conditions do not coincide with the predictions (5) and (6) conformal field theory makes for the profiles of scaling operators near a boundary. Here one has to consider certain linear combinations of spin-operators to obtain the correct behaviour. These combinations and the respective behaviour near the boundary are given in (37) and (4.3). In this paper we have also discussed the appearance of massive excitations, i.e. of fermions with a non-vanishing energy as $`L`$ goes to infinity. Such a fermion appears if the reciprocal value of one of the expressions (46) or (47) becomes larger than 1. These expressions reduce to (48) respectively (49) in the case of purely non-diagonal respectively diagonal boundary terms. We have computed the profiles near the boundaries in the limit $`L\mathrm{},j1`$ for certain states with differ from the ground-state of $`H_{\mathrm{long}}`$ by a massive excitation for boundary parameters satisfying (50) or (53). For the first type of boundary terms there exists one massive excitation, whereas for the second type there might appear two of them. The profiles near the boundaries are given by (51),(52) respectively by (54)-(57). If the value of $`\xi _\alpha `$ equals $`\xi _\beta `$ (see (53)) the profiles are given by (58) and (59). The author is grateful to V.Rittenberg for many discussions and useful suggestions. ## Appendix A Diagonalization of $`H_{\mathrm{long}}`$ and the projection mechanism In this appendix we are going to recapitulate the method we used in to diagonalize $`H_{\mathrm{long}}`$ given in (2) and the projection mechanism, which yields the eigenvalues and eigenvectors of the Hamiltonian $`H`$ in (1). This is necessary in order to fix our notation we are going to use during our computations for the profiles of $`\sigma _j^z`$ and $`\sigma _j^x`$ which are presented in Appendix B respectively in Appendix C. In its diagonal form $`H_{\mathrm{long}}`$ reads $$H_{\mathrm{long}}=\underset{k=0}{\overset{L+1}{}}2\mathrm{\Lambda }_kb_ka_k\underset{k=0}{\overset{L+1}{}}\mathrm{\Lambda }_k$$ (60) where the $`b_k`$ and $`a_k`$ satisfy the anti-commutation relations of fermionic creation and annihilation operators. The factor $`2`$ in equation (60) is just an remnant of our notation introduced in . The values of the $`\mathrm{\Lambda }_k`$ are determined by the eigenvalues of a skew-symmetric $`(2L+4)\times (2L+4)`$ matrix $`M`$, i.e. $$M(\varphi _k^\pm )=\pm \mathrm{\Lambda }_k(\varphi _k^\pm ).$$ (61) The explicit form of $`M`$ is given in . The eigenvalues $`\mathrm{\Lambda }_k`$ can be obtained by the zeros $`x_k`$ of a polynomial $`p(x^2)`$, which has also been given in , i.e. $$\mathrm{\Lambda }_k=\frac{1}{4}(x_k+1/x_k).$$ (62) We have also shown in how to compute the eigenvectors $`(\varphi _k^\pm )`$ as a function of the $`x_k`$ and the boundary parameters $`\alpha _\pm ,\beta _\pm ,\alpha _z`$ and $`\beta _z`$. For certain choices of the boundary terms the $`x_k`$ and the $`(\varphi _k^\pm )`$ can be computed exactly on the finite chain, since the polynomial $`p(x^2)`$ factorizes into cyclotomic polynomials . For the boundary parameters given in table 1 and table 2 this is the case <sup>4</sup><sup>4</sup>4They correspond to the factorizable cases 2,4,9,11,14 and 16 in the notation of .. Otherwise the $`x_k`$ and the $`(\varphi _k^\pm )`$ may be computed solving (61) numerically. The eigenvectors $`(\varphi _k^\pm )`$ can be used to express the Clifford operators $`\tau _j^\pm `$, which are defined by $$\tau _j^{+,}=\left(\underset{i=0}{\overset{j1}{}}\sigma _i^z\right)\sigma _j^{x,y}$$ (63) in the terms of the creation and annihilation operators $`b_k`$ and $`a_k`$, i.e. $$\tau _j^\mu =\underset{k=0}{\overset{L+1}{}}(\varphi _k^{})_j^\mu a_k+(\varphi _k^+)_j^\mu b_k$$ (64) where the $`(\varphi _k^\pm )`$ satisfy the orthogonality relation $`_{j,\mu }(\varphi _l^+)_j^\mu (\varphi _k^{})_j^\mu =\delta _{lk}`$. Equation (64) allows to write the spin operators in terms of the $`b_k,a_k`$ and the eigenvectors $`(\varphi _k^\pm )`$. Therefore the knowledge of these vectors is a crucial point in view to the computation of the magnetization profiles. The vacuum state is defined by $$a_k|vac=0k.$$ (65) Note, that the vacuum state itself is not element of one of the four sectors $`(+,+),(+,),(,)`$ or $`(,+)`$. However, in view to applications to the Hamiltonian $`H`$ we are interested into the eigenstates of $`H_{\mathrm{long}}`$ which lie in the $`(+,+)`$-sector. Hence we have defined the states $$|v^\pm =\frac{1}{\sqrt{2}}(1\pm b_0)|vac$$ (66) where $`b_0`$ denotes the creation operator of the spurious zero mode which is always present in the spectrum of $`H_{\mathrm{long}}`$. The associated eigenvectors $`(\varphi _0^\pm )`$ are given by $$\varphi _0^+=(0,\frac{1}{2},0,\mathrm{},0,\frac{\mathrm{}}{2},0)\varphi _0^{}=(0,\frac{1}{2},0,\mathrm{},0,\frac{\mathrm{}}{2},0)$$ (67) where we used the notation $`(\varphi _k^\pm )=((\varphi _k^\pm )_0^{},(\varphi _k^\pm )_0^+,(\varphi _k^\pm )_1^{},(\varphi _k^\pm )_1^+,\mathrm{},(\varphi _k^\pm )_{L+1}^{},(\varphi _k^\pm )_{L+1}^+)`$. The states $`|v^\pm `$ satisfy $$\sigma _0^x|v^\pm =\pm |v^\pm \sigma _{L+1}^x=\pm \eta |v^\pm $$ (68) where $`\eta ^2=1`$. If $`\eta =1`$ the state $`|v^+`$ is element of the $`(+,+)`$-sector and so are all the states obtained by an even number of fermion excitations, where we disregard the spurious zero mode. If $`\eta =1`$ all the states obtained by the excitation of an odd number of fermions with respect to the state $`|v^{}`$ are elements of the $`(+,+)`$-sector and hence build up the spectrum of $`H`$, where we again have to disregard the spurious zero mode. In the case of hermitian boundaries the value of $`\eta `$ may be computed via $$\eta =(1)^{L+1}\frac{\alpha _{}\beta _++\alpha _+\beta _{}}{4^{L+1}_{k0}\mathrm{\Lambda }_k}.$$ (69) Note that (69) is of no use if there exist additional zero modes on top of the spurious zero mode. ## Appendix B Calculation of the $`\sigma _j^z`$ profiles In this appendix we are going to explain how to compute the $`\sigma _j^z`$-profiles and how we have derived the ground-state profiles for the boundary terms in table 1, i.e. equations (11)-(20). We will also derive the profiles (51) and (52) associated to the excitation of the massive fermion appearing for the boundaries (50). Note that these boundaries coincide with the boundary terms of case e in table 1 if one identifies $`\mathrm{}^\xi `$ with $`\mathrm{tan}(\frac{\pi }{4}+\frac{\delta }{2})`$. Using (64), (66) and (67) one can show that the $`\sigma _j^z`$ profile for $`1jL`$ and for any excited state $`\psi _r^\pm =_{n=1}^rb_{k_n}v^\pm `$ with $`k_n0`$ is given by $$\psi _r^\pm \sigma _j^z\psi _r^\pm =vac\sigma _j^zvac+\mathrm{}\underset{n=1}{\overset{r}{}}\left[(\varphi _{k_n}^{})_j^+(\varphi _{k_n}^+)_j^{}(\varphi _{k_n}^+)_j^+(\varphi _{k_n}^{})_j^{}\right]$$ (70) where $$vac\sigma _j^zvac=\mathrm{}vac\tau _j^+\tau _j^{}vac=\mathrm{}\underset{k=0}{\overset{L+1}{}}(\varphi _k^{})_j^+(\varphi _k^+)_j^{}.$$ (71) Before turning to the profiles for the boundaries in table 1, consider the case $`\alpha _z=\beta _z=0`$. According to it is always possible to chose the eigenvectors $`(\varphi _k^\pm )`$ such that the RHS of (70) and (71) vanishes for $`0<j<L+1`$ (see the equations (2.38) and (2.40) in ). In this case the profiles for all the states $`\psi _r^\pm `$ vanish identically. Thus the profile for the ground-state of $`H`$ vanishes also if it is non-degenerate. The partition functions obtained in show that for $`L1`$ this is the generic case. Therefore we considered only cases with at least one diagonal boundary term. If the ground-state of $`H`$ is degenerate one has also to consider linear combinations of the states $`\psi _r^\pm `$ which correspond to ground-states $`H`$. For the cases given in table 1 the polynomial $`p(x^2)`$ factorizes into cyclotomic polynomials. This enables us to compute the eigenvectors $`(\varphi _k^\pm )`$ exactly on the finite chain, as explained in . The explicit computations are cumbersome but straight forward and will therefore not be given. For all cases in table 1 there appears an additional zero mode on top of the spurious mode which does not give a contribution to the RHS of (70). Hence the profiles for the two states $`|v^+`$ and $`b_z|v^{}`$, where $`b_z`$ denotes the creation operator corresponding to the additional zero mode, are identical. One of these states is element of the $`(+,+)`$ sector and corresponds to the ground-state of $`H`$. However we do not have to know which one, since we are only interested into the profile (note that (69) is not applicable, due to the zero mode). Hence in the following we may assume the state $`|v^+`$ to be in the $`(+,+)`$-sector. For the cases a,b,c and d with $`L`$ even no complications arise. The ground-state of $`H`$ is non-degenerate and using the expressions for the $`(\varphi _k^\pm )`$ in (71) leads to a geometric series for each case. Performing the computations yields (11)-(13) and (15). For case d with $`L`$ odd the ground-state is two-fold degenerate, since here two zero modes appear on top of the spurious zero mode. The non-vanishing components of the respective eigenvectors are given by $$(\varphi _{z_1}^\pm )_0^{}=1(\varphi _{z_1}^\pm )_{L+1}^+=\pm \mathrm{}$$ (72) respectively $$(\varphi _{z_2}^+)_j^{}=\mathrm{}(\varphi _{z_2}^+)_j^+=\frac{(\mathrm{})^j}{\sqrt{L}}\{\begin{array}{cc}\mathrm{}\hfill & \text{for }j\text{ even}\hfill \\ 1\hfill & \text{for }j\text{ odd}\hfill \end{array}$$ (73) where $`0<jL`$ and $`(\varphi _k^{})_j^\pm =\pm (\varphi _k^+)_j^\pm `$. In this case we computed the profiles for the two states $`|v^+`$ and $`b_{z_2}b_{z_1}|v^+`$. This leads to (14) where one has to choose the $``$ sign for the profile of $`|v^+`$ respectively the $`+`$ sign for the profile of $`b_{z_2}b_{z_1}|v^+`$. However, the computations for case e respectively the boundary terms (50) are more involved. It is convenient to use the parameter $`s=\mathrm{}^{2\xi }`$ where $`\xi `$ has been introduced in (50) instead of $`\delta `$ being used in table 1. For odd values of $`L`$ we have obtained $$\sigma _j^z=\frac{\varsigma _j}{\sqrt{s}L}\frac{1s}{1s^L}s^{j1}.$$ (74) where $`\varsigma _j={\displaystyle \underset{k=1}{\overset{N}{}}}{\displaystyle \frac{x_k^{2j3}+x_k^{2j+3}x_k^{2j1}x_k^{2j+1}+s\left(x_k^{2j+1}+x_k^{2j1}x_k^{2j1}x_k^{2j+1}\right)}{1/s+sx_k^21/x_k^2}}`$ (75) with $`x_k=\mathrm{}^{\mathrm{}\frac{k\pi }{L}}`$ and $`N=(L1)/2`$. This expression simplifies to a geometric series for $`s=1`$. Straight forward computation yields (15) as for case d with $`L`$ even. We have not been able to compute this expression for $`s1`$. As for case d with $`L`$ odd, there appear two zero-modes on top of the spurious zero mode if we take $`L`$ even. The non-vanishing components of the corresponding eigenvectors for the first zero mode are again given by (72) whereas for the second zero mode they read $$(\varphi _{z_2}^+)_j^{}=\mathrm{}(\varphi _{z_2}^+)_j^+=\mathrm{}^j\left(\frac{2}{L(1+s)}\right)^{1/2}\{\begin{array}{cc}\hfill \mathrm{}& \text{for }j\text{ even}\hfill \\ \hfill \sqrt{s}& \text{for }j\text{ odd}\hfill \end{array}$$ (76) where $`0<jL`$ and $`(\varphi _k^{})_j^\pm =\pm (\varphi _k^+)_j^\pm `$. Again we computed the profiles for the two states $`|v^+`$ and $`b_{z_2}b_{z_1}|v^+`$. Labelling the profile for the state $`|v^+`$ by a $``$ sign and the profile for the state $`b_{z_2}b_{z_1}|v^+`$ by a $`+`$ sign, we have obtained $$\sigma _j^z_\pm =\frac{\varsigma _j}{\sqrt{s}L}\frac{1s}{1s^L}s^{j1}\pm \frac{2}{L(1+s)}\{\begin{array}{cc}1\hfill & \text{for }j\text{ even}\hfill \\ s\hfill & \text{for }j\text{ odd}\hfill \end{array}$$ (77) where $`\varsigma `$ is again given by (75) as for $`L`$ odd, but with $`N=L/21`$. Choosing $`s=1`$ allows for the computation of the sum in (75) and yields (14) as for case d with $`L`$ odd. Although our results on the finite chain are restricted to $`s=1`$, one can show that in the continuum limit ($`z=j/L`$ fixed, $`L\mathrm{}`$) $`\varsigma _j`$ is given by $$\varsigma _j=\frac{(1)^j}{1/s+1}\frac{1}{\mathrm{sin}(\pi z)}$$ (78) for $`L`$ odd respectively by $$\varsigma _j=\frac{(1)^j}{1/s+1}\mathrm{cot}(\pi z)$$ (79) for $`L`$ even. Plugging this into (77) and (74) respectively results in the equations (18) and (19). Note that the exponential contributions in (74) and (77) vanish in the continuum limit. The profile in the limit $`L\mathrm{},j`$ fixed, leading to (20), may also be computed. In this limit the quantity $`\varsigma _j/L`$ appearing in (74) and (77) can be represented as an integral, i.e. $$\varsigma _j^{\mathrm{}}=\underset{L\mathrm{}}{lim}\frac{\varsigma _j}{L}=\frac{2}{\pi }\underset{0}{\overset{\frac{\pi }{2}}{}}\frac{\mathrm{cos}((2j3)t)(s+1)\mathrm{cos}((2j1)t)+s\mathrm{cos}((2j+1)t)}{1/s+s2\mathrm{cos}(2t)}t.$$ (80) This integral can be calculated exactly in terms of hypergeometric functions. We obtained $`\varsigma _j^{\mathrm{}}={\displaystyle \frac{2}{\pi }}\left[{\displaystyle \frac{(1)^j}{2j1}}(1)^j({\displaystyle \frac{1}{s}}1){\displaystyle \frac{F(1,1/2+j;3/2+j;1/s)}{2j+1}}\right]\text{for }s>1`$ (81) $`{\displaystyle \frac{2}{\pi }}\left[{\displaystyle \frac{(1)^j}{j}}{\displaystyle \frac{s}{1+s}}\right]+\mathrm{O}(j^2)`$ (82) and $`\varsigma _j^{\mathrm{}}={\displaystyle \frac{2}{\pi }}\left[{\displaystyle \frac{(1)^j}{2j1}}(1)^j(1s){\displaystyle \frac{F(1,1/2j;3/2j;s)}{2j1}}\right]\text{for }0<s<1`$ (83) $`{\displaystyle \frac{2}{\pi }}\left[{\displaystyle \frac{(1)^j}{j}}{\displaystyle \frac{s}{1+s}}+{\displaystyle \frac{\pi (1s)}{2}}s^{j1/2}\right]+\mathrm{O}(j^2).`$ (84) For the asymptotic expansions of the hypergeometric functions see . Using these results in (77) and (74) gives for even and odd values of $`L`$ $$\sigma _j^z=\frac{(1)^j}{\pi \mathrm{cosh}\xi }j^1+\mathrm{}$$ (85) Using $`\mathrm{}^\xi =\sqrt{2}\alpha _z`$ yields (20). Note that the exponential contributions in (77) and (74) (which, in the limit in question, are present only for $`0<s<1`$) are cancelled by the exponential part in (84). We have also considered the profile for a state where the massive fermion is excited, i.e. $`\xi =b_\xi b_zv^+`$ for $`L`$ odd respectively $`\xi =b_\xi b_{z_1}v^+`$ for $`L`$ even (there are two zero modes for $`L`$ even, where the one labelled by $`z_1`$ gives no contribution to the profile (70)). Using (70),(85) and the explicit expressions of the eigenvectors $`(\varphi _\xi ^\pm )`$ corresponding to the massive fermion, i.e. $$(\varphi _\xi ^+)_j^{}=\mathrm{}(\varphi _\xi ^+)_j^+=\mathrm{}(1)^j\left(\frac{1\mathrm{}^{2\xi }}{1\mathrm{}^{2\xi L}}\right)^{1/2}\mathrm{}^{2\xi (j1)}.$$ (86) where $`0<j<L+1`$ and $`(\varphi _\xi ^{})_j^\pm =\pm (\varphi _\xi ^+)_j^\pm `$, yields (51) for $`\xi >0`$ $`(s<1)`$. Note that for $`\xi <0`$ $`(s>1)`$ and $`j`$ fixed the vector components (86) vanish as $`L`$ goes to infinity and the profile for the excited state $`\xi `$ equals the ground-state profile given in (85). By symmetry (see (50)) this yields the profile near the right boundary for the case $`\xi >0`$ $`(s<1)`$, i.e. equation (52). The $`U(1)`$ charge, i.e. the projection of the total spin onto the z-axis, of the state $`\xi `$ can be computed using the explicit expressions of the $`(\varphi _k^\pm )`$ and summing over $`j`$ in (70) and (71) before summing over $`k`$ (in this way the computation can be done exactly on the finite chain, independently of the value of $`s`$). This yields 0 for $`L`$ even respectively $`\frac{1}{2}`$ for $`L`$ odd as stated in section 5.2. ## Appendix C Calculation of the $`\sigma _j^x`$-profiles In this appendix we are going to derive our exact results for the $`\sigma _j^x`$ profiles for the two cases from table 2, i.e. equations (22),(25) and (26). Furthermore we are going to explain how we have obtained the profiles (54) and (55) corresponding to the boundary excitation for the boundary terms in (53) with $`\xi _\beta =\xi _\alpha `$. We have shown in how to compute the expectation values of the $`\sigma _j^x`$-operator in terms of pfaffians. Here we will only consider cases where diagonal boundary terms are absent (see table 2 and (53)). This enables us to represent $`\sigma _j^x`$ in terms of the determinant of a $`j\times j`$ matrix , i.e. $$v^+\sigma _j^xv^+=f_{0j}\left|\begin{array}{cccc}\tau _0^{}\tau _1^+& \tau _0^{}\tau _1^{}& \tau _0^{}\tau _3^+& \mathrm{}\\ \tau _2^+\tau _1^+& \tau _2^+\tau _1^{}& \tau _2^+\tau _3^+& \mathrm{}\\ \tau _2^{}\tau _1^+& \tau _2^{}\tau _1^{}& \tau _2^{}\tau _3^+& \mathrm{}\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\end{array}\right|$$ (87) where $`f_{0j}=\mathrm{}`$ for odd values of $`j`$ and $`f_{0j}=1`$ for even values of $`j`$. The expectation values of the basic contractions of pairs are given by $$\tau _i^\mu \tau _j^\nu =\underset{k=0}{\overset{L+1}{}}(\varphi _k^{})_i^\mu (\varphi _k^+)_j^\nu .$$ (88) The profiles for the excited states $$r=\underset{j=1}{\overset{r}{}}b_{k_j}v^{(1)^r}$$ (89) are also given by the RHS of (87) if one exchanges the $`\tau _i^\mu \tau _j^\nu `$ in (87) by $$\tau _i^\mu \tau _j^\nu _r=\tau _i^\mu \tau _j^\nu +\underset{l=1}{\overset{r}{}}\left[(\varphi _{k_l}^+)_i^\mu (\varphi _{k_l}^{})_j^\nu (\varphi _{k_l}^{})_i^\mu (\varphi _{k_l}^+)_j^\nu \right].$$ (90) We will start with our computations for case g from table 2 and will then turn to case f from table 2. Afterwards we will consider the boundary terms (53). Again it is possible to solve the eigenvalue equation (61) exactly . As for the case of the $`\sigma _j^z`$-profiles we will not give the explicit expressions for the eigenvectors $`(\varphi _k^\pm )`$. They have been obtained as explained in <sup>5</sup><sup>5</sup>5The boundary terms (53) correspond to case 10 in the notation of .. ### C.1 Case g We will first show, how to derive equation (26). Afterwards we will compute the function $`f(j,L)`$ in (26) in the limit $`L\mathrm{},j1`$. Using (26) this yields (25). It turns out, that for our purpose it is convenient to consider the Hamiltonian $`H_{\mathrm{long}}`$ in a different basis, i.e. we consider the Hamiltonian $$H_{\mathrm{long}}^{}=U_nH_{\mathrm{long}}U_n^1$$ (91) where $`U_n`$ is given by $$U_n=\left(\begin{array}{cc}1\hfill & 0\hfill \\ 0\hfill & 1\hfill \end{array}\right)\underset{j=1}{\overset{L}{}}\left(\begin{array}{cc}1\hfill & 0\hfill \\ 0\hfill & \mathrm{}^\mathrm{}\phi \mathrm{\Gamma }^j\hfill \end{array}\right)\left(\begin{array}{cc}1& 0\\ 0& (1)^n\end{array}\right)$$ (92) and $$\mathrm{\Gamma }=\mathrm{exp}\left(\mathrm{}\frac{\chi +n\pi }{L+1}\right).$$ (93) The $`\sigma _j^x`$ profile for any state $`g`$ in the original basis is related to the magnetization profiles for the state $`g^{}=Ug`$ via $$g\sigma _j^xg=\mathrm{cos}\left(\phi +\frac{\chi +n\pi }{L+1}j\right)g^{}\sigma _j^xg^{}+\mathrm{sin}\left(\phi +\frac{\chi +n\pi }{L+1}j\right)g^{}\sigma _j^yg^{}.$$ (94) The Hamiltonian in the new basis can still be written in terms of free fermions. Of course the fermion energies are not affected by the transformation. The $`L+1`$ non-zero fermion energies are given by $$2\mathrm{\Lambda }_k=\mathrm{sin}\left(\frac{2k+1}{L+1}\frac{\pi }{2}\frac{\chi }{L+1}\right)$$ (95) where $`1kL+1`$. Computing the eigenvectors $`(\varphi _k^\pm )^{}(\chi ,\phi )`$ in the new basis yields that they are independent from the values of $`\chi `$ and $`\phi `$ and can be given in terms of the vectors $`(\varphi _k^\pm )(0,0)`$ in the original basis, i.e. for $`n0`$ we have found $`(\varphi _k^\pm )^{}(\chi ,\phi )=(\varphi _{k+n}^\pm )(0,0)`$ $`\text{for }1kL+1n`$ (96) $`(\varphi _k^\pm )^{}(\chi ,\phi )=(\varphi _{k+n(L+1)}^{})(0,0)`$ $`\text{for }L+1n<kL+1`$ (97) whereas for $`n<0`$ $`(\varphi _k^\pm )^{}(\chi ,\phi )=(\varphi _{k|n|}^\pm )(0,0)`$ $`\text{for }|n|+1kL+1`$ (98) $`(\varphi _k^\pm )^{}(\chi ,\phi )=(\varphi _{L+1+j|n|}^{})(0,0)`$ $`\text{for }1k<|n|+1.`$ (99) Now define for $`n`$ fixed the state $$n=b_{L+2n}\mathrm{}b_{L+1}v^{(1)^n}\text{for }n>0$$ (100) respectively $$n=b_1\mathrm{}b_{|n|}v^{(1)^n}\text{for }n<0.$$ (101) For $`n=0`$ define $`0=v^+`$. The expectation values in the new basis may be computed in the same way than in the original basis by exchanging the $`(\varphi _k^\pm )`$ in (90) and (88) by the $`(\varphi _k^\pm )^{}`$. Using (96)-(99) yields that the $`\sigma _j^x`$-profiles for the states $`n`$ in the new basis are given by $$n^{}\sigma _j^xn^{}=v^+\sigma _j^xv^+.$$ (102) where the RHS has to be evaluated in the old basis and for the choice $`\chi =\phi =0`$. In the same way one can show that $$n^{}\sigma _j^yn^{}=v^+\sigma _j^yv^+$$ (103) where again the RHS has to be computed in the original basis for the choice $`\chi =\phi =0`$. Since for $`\chi =\phi =0`$ the Hamiltonian $`H_{\mathrm{long}}`$ commutes not only with $`\sigma _0^x`$ and $`\sigma _{L+1}^x`$, but also with $`\sigma _0^x\sigma _1^x\mathrm{}\sigma _{L+1}^x`$, the $`\sigma _j^y`$ profile in (103) vanishes. However, we have been interested in the ground-state profiles. Computing the value of $`\eta `$ using (69) with (95) yields $`\eta =(1)^{L+1}`$ and hence the state $`n`$ is element of the $`(+,(1)^{L+1+n})`$-sector (the excitation of a fermion changes the sector from $`(\pm ,ϵ)`$ to $`(,ϵ)`$ ). Which of states $`n`$ corresponds to the ground-state of $`H`$ may be determined by computing the associated energies using (95). For $`L`$ odd the ground-state of $`H`$ corresponds to the state $`0=v^+`$ if $`\chi \pi `$. For $`\chi =\pi `$ the ground-state of $`H`$ is twofold degenerate. We obtain two linear independent ground-states by choosing the states $`0`$ and $`2`$. For $`L`$ even and $`\chi >0`$ the ground-state of $`H`$ corresponds to $`1`$ whereas for $`\chi <0`$ the ground-state corresponds to $`1`$. For $`\chi =0`$ the ground-state is twofold degenerate and we obtain two linear independent ground-states choosing the states $`\pm 1`$. Finally, using (94) and (102) with $`f(j,L)=v^+\sigma _j^xv^+`$ (for the choice $`\chi =\phi =0`$) and $`m=n/2`$ yields (26) and the values of $`m`$ as explained in the text. In order to compute $`f(j,L)`$ we have calculated the expectation values of the basic contractions of pairs (88) using the vectors $`(\varphi _k^\pm )(0,0)`$. For $`i,j<L+1`$ this results in $$\tau _i^\pm \tau _j^\pm =0\tau _0^{}\tau _j^+=\sqrt{2}g(j)$$ (104) $$\tau _i^+\tau _j^{}=g(i+j)g(ij)\tau _i^{}\tau _j^+=g(i+j)+g(ij)$$ (105) where the values of $`i`$ are always even and the values of $`j`$ are always odd. Furthermore we have defined $$g(r)=\frac{\mathrm{}}{L+1}\frac{\mathrm{sin}(r\pi /2)}{\mathrm{sin}(r\pi /(2L+2))}.$$ (106) Using elementary operations on determinants in (87) one can show that $$f(j,L)=(1)^j\left(\frac{2}{\pi }\right)^j\sqrt{2}det\text{Q}_j,$$ (107) where the matrix elements of the $`j\times j`$-matrix $`\text{Q}_j`$ are given by $$(\text{Q}_j)_{kl}=\left(\frac{\pi }{2L+2}\right)^j\left(\mathrm{sin}\frac{(2k2l+1)\pi }{2L+2}\right)^1.$$ (108) The boundary behaviour of the profile (i.e. (25)) is obtained taking the limit $`L\mathrm{}`$ of $`\text{Q}_j`$, which yields $$\underset{L\mathrm{}}{lim}\text{Q}_j=\left(\begin{array}{ccccc}1& 1& \frac{1}{3}& \mathrm{}& \frac{1}{32j}\\ \frac{1}{3}& 1& 1& \mathrm{}& \frac{1}{52j}\\ \frac{1}{5}& \frac{1}{3}& 1& \mathrm{}& \frac{1}{72j}\\ \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}& \mathrm{}\\ \frac{1}{2j1}& \frac{1}{2j3}& \frac{1}{2j5}& \mathrm{}& 1\end{array}\right).$$ (109) The large $`j`$ behaviour of the determinant of this matrix is already known . With (107) we obtain for $`j1`$ $$f(j,\mathrm{})=(1)^jAj^{\frac{1}{4}}+\mathrm{}$$ (110) where $`A`$ is defined by (23). Using (26) this yields (25). The function $`f(j,L)`$ may also be calculated on the finite chain using (108) in (107) and computing the determinant numerically. ### C.2 Case f For this case there exists an additional zero mode in the spectrum. Hence the ground-state of $`H`$ corresponds to $`v^+`$ or to $`b_zv^{}`$, but we can not decide which one, since (69) can not be applied. However, the eigenvectors $`(\varphi _z^\pm )`$ corresponding to the additional zero mode give no contribution to (90) as long as $`i,j<L+1`$ . Hence the profiles for states $`v^+`$ and $`b_zv^{}`$ are identical for $`j<L+1`$. Defining $$f(r)=\mathrm{cos}\left(\frac{r}{2L+2}\pi \right)g(r)$$ (111) where $`g(r)`$ is already given in (106), we have obtained for $`L`$ odd: $$\tau _i^\pm \tau _j^\pm =0\tau _0^{}\tau _j^+=\sqrt{2}g(j)$$ (112) $$\tau _i^+\tau _j^{}=f(i+j)f(ij)\tau _i^{}\tau _j^+=g(i+j)+g(ij)$$ (113) where $`i`$ is always even and $`j`$ is always odd. For $`L`$ even we obtain $$\tau _i^\pm \tau _j^\pm =0\tau _0^{}\tau _j^+=\sqrt{2}f(j)$$ (114) $$\tau _i^+\tau _j^{}=g(i+j)g(ij)\tau _i^{}\tau _j^+=f(i+j)+f(ij).$$ (115) The expressions (112)-(115) have the same large $`L`$ limit as the respective expressions for case g in (104) and (105). Thus the behaviour of the ground-state profile near the left boundary is also given by the RHS of equation (110) which yields (22). ### C.3 Here we consider the $`\sigma _j^x`$-profile for the boundaries in (53) with $`\xi _\beta =\xi _\alpha `$. In this case there appears a massive fermion with energy $`2\mathrm{\Lambda }=2\mathrm{cosh}\xi _\alpha `$ (see section 5). Here we are interested into the effect of the excitation of this fermion onto the magnetization profile. For simplicity we use $`\xi \xi _\alpha `$ in the following. We have restricted ourselves to even values of $`L`$. Since $`\eta =(1)^{L+1}`$ the state $`|\xi =b_\xi v^{}`$ (where by $`b_\xi `$ we denote the creation operator associated to the massive fermion) lies in the $`(+,+)`$-sector of the Hilbert-space of $`H_{\mathrm{long}}`$ and hence corresponds to an eigenstate of $`H`$. For odd values of $`L`$ the $`(+,+)`$-sector is given by the excitations of an even number of fermions with respect to the ground-state $`|v^+`$ and hence the state $`|\xi `$ does not correspond to an eigenstate of $`H`$ for $`L`$ odd. A lengthy computation yields the expectation values of the basic contractions of pairs (88), i.e. $$\tau _i^+\tau _j^{}=g(i+j)g(ij)$$ (116) $$\tau _i^{}\tau _j^+=\mathrm{}\frac{\mathrm{\Xi }(i+j)+1}{L+1}+g(ij)+\frac{2\mathrm{}\mathrm{sinh}(2\xi )\mathrm{}^{(i+j)\xi }}{\mathrm{}^{2(L+1)\xi }1}$$ (117) $$\tau _0^{}\tau _j^+=\frac{\sqrt{2}\alpha _+}{\mathrm{}^{2\xi }+1}\left(\mathrm{}\frac{\mathrm{\Xi }(j)+1}{L+1}+g(j)+\frac{2\mathrm{}\mathrm{sinh}(2\xi )\mathrm{}^{j\xi }}{\mathrm{}^{2(L+1)\xi }1}\right)$$ (118) where $`g(r)`$ has been defined in (106) and $`\mathrm{\Xi }(q)={\displaystyle \underset{n=1}{\overset{\frac{L}{2}}{}}}\left[\mathrm{}^{2\xi }\mathrm{cos}\left({\displaystyle \frac{q2}{L+1}}n\pi \right)+\mathrm{}^{2\xi }\mathrm{cos}\left({\displaystyle \frac{q+2}{L+1}}n\pi \right)2\mathrm{cos}\left({\displaystyle \frac{q}{L+1}}n\pi \right)\right]`$ (119) $`\times \left[\mathrm{cosh}(2\xi )\mathrm{cos}\left({\displaystyle \frac{2n}{L+1}}\pi \right)\right]^1.`$ (120) For $`L\mathrm{}`$ we can rewrite $`\frac{\mathrm{\Xi }(q)}{L+1}`$ in terms of an integral which can be calculated using standard methods. In this limit we may write $$\tau _i^+\tau _j^{}=\frac{2\mathrm{}}{\pi }\left(\frac{\mathrm{sin}((i+j)\pi /2)}{i+j}\frac{\mathrm{sin}((ij)\pi /2)}{ij}\right)$$ (121) $`\tau _i^{}\tau _j^+=`$ $`{\displaystyle \frac{2\mathrm{}}{\pi }}[{\displaystyle \frac{\mathrm{sin}((ij)\pi /2)}{ij}}+{\displaystyle \frac{\mathrm{sin}((i+j)\pi /2)}{i+j}}\mathrm{}^{2\xi }`$ (122) $`2\mathrm{sinh}(2\xi )\mathrm{}^{(i+j)\xi }({\displaystyle \underset{m=0}{\overset{\frac{i+j1}{2}}{}}}{\displaystyle \frac{(1)^m\mathrm{}^{(2m+1)\xi }}{2m+1}}\mathrm{arctan}(\mathrm{}^\xi ))]`$ $`\tau _0^{}\tau _j^+=`$ $`{\displaystyle \frac{2\sqrt{2}\alpha _+\mathrm{}}{\mathrm{}^{2\xi }+1}}[{\displaystyle \frac{\mathrm{sin}(j\pi /2)}{j}}(1+\mathrm{}^{2\xi })`$ (123) $`2\mathrm{sinh}(2\xi )\mathrm{}^{j\xi }({\displaystyle \underset{m=0}{\overset{\frac{j1}{2}}{}}}{\displaystyle \frac{(1)^m\mathrm{}^{(2m+1)\xi }}{2m+1}}\mathrm{arctan}(\mathrm{}^\xi ))].`$ We used these expressions for the numerical calculation of the determinant in (87) for $`j=1,2\mathrm{}100`$. We have chosen 48 different values of $`\xi `$ in the range between $`4.0`$ and $`4.0`$. Numerical extrapolation yields $$v^+\sigma _j^xv^+(1)^jAj^{1/4}$$ (124) where the numerical value of $`A`$ is given by the RHS of (9) with an accuracy of the order of $`10^{13}`$. As already mentioned above here we are interested into the profile of the state where the massive fermion is excited with respect to $`|v^{}`$. The eigenvectors $`(\varphi _\xi ^\pm )`$ associated to the massive fermion are given by $$(\varphi _\xi ^+)_0^{}=\frac{\alpha _+}{\mathrm{cosh}\xi }/\rho (\varphi _\xi ^+)_0^+=0$$ (125) $$(\varphi _\xi ^+)_j^{}=\{\begin{array}{cc}\sqrt{2}\mathrm{exp}((1j)\xi )/\rho \hfill & \text{for }j\text{ even}\hfill \\ 0\hfill & \text{for }j\text{ odd}\hfill \end{array}$$ (126) $$(\varphi _\xi ^+)_j^+=\{\begin{array}{cc}0\hfill & \text{for }j\text{ even}\hfill \\ \mathrm{}\sqrt{2}\mathrm{exp}((1j)\xi )/\rho \hfill & \text{for }j\text{ odd}\hfill \end{array}$$ (127) $$(\varphi _\xi ^+)_{L+1}^{}=0,(\varphi _\xi ^+)_{L+1}^+=\frac{\mathrm{}\beta _+}{\mathrm{cosh}\xi }\mathrm{exp}((1L)\xi )/\rho $$ (128) where $$\rho =\sqrt{\frac{\mathrm{}^{2L\xi }\mathrm{}^{2\xi }}{\mathrm{sinh}(2\xi )}}$$ (129) and $`(\varphi _k^{})_j^\mu =(1)^j(\varphi _k^+)_j^\mu `$. In the case $`\xi <0`$ this mode does not contribute to (90) in the limit $`L\mathrm{}`$, since (129) diverges, i.e. $$\xi \sigma _j^x\xi (1)^jAj^{1/4}\text{for }\xi <0.$$ (130) This changes for $`\xi >0`$. We assumed the following behaviour of the profile: $$\xi \sigma _j^x\xi (1)^{j+1}A(1c\mathrm{}^{j\eta })j^{1/4}.$$ (131) Using (121)-(123) and (125)-(129) in (90) we computed the profile for $`j=1,2\mathrm{}100`$ for $`24`$ different values of $`\xi `$ in the range between $`0`$ and $`4`$. The values of $`c`$ and $`\eta `$ have been computed using numerical extrapolation by considering the quantity $$\mathrm{ln}\left(1+\frac{\xi \sigma _j^x\xi }{v^+\sigma _j^xv^+}\right)\mathrm{ln}cj\eta $$ (132) Our results have confirmed our assumption (131) and the numerical estimates from our extrapolations suggest $$c=2\mathrm{}^{(1)^j\eta }\eta =2\xi $$ (133) which yields (54). The largest deviation of the extrapolated values for $`c`$ and $`\eta `$ from (133) has been of the order of $`10^{11}`$. We omit any tables on this. Computing the profile for $`\xi >0`$ near the right boundary is equivalent to the computation of the profile near the left boundary for $`\xi <0`$ (see (53)) which is given by (130). This yields (55). ## Appendix D Numerical verification of the conjectures in section 3 We have verified our conjectures (31)-(35) for the continuum limit of the magnetization profiles for the boundary terms (27)-(30). In order to do so, we have used numerical diagonalization to obtain the vectors $`(\varphi _k^\pm )`$ (see (61)) numerically and used (70) and (87) to obtain the profiles of $`\sigma _j^z`$ respectively $`\sigma _j^x`$ on the finite chain. Notice that (87) is only valid if $`\alpha _z=\beta _z=0`$ and can not be applied to compute the $`\sigma _j^x`$ profile for the boundary terms (30). In this case we made use of the determinant representation (10.14) in of the pfaffian which determines the profiles. We computed the profiles on finite lattices with $`L=20,60,100,\mathrm{}`$ respectively $`L=21,63,105,\mathrm{}`$, where the maximum value of $`L`$ has been chosen in the range of $`500700`$. For each type of boundaries in (27)-(30) we have used $`3050`$ different values of the respective free parameters. The values obtained on the finite lattice have been extrapolated to $`L=\mathrm{}`$ for $`z=j/L`$ fixed using the BST algorithm, where $`z=\frac{n}{20}`$ for $`L`$ even respectively $`z=\frac{n}{21}`$ for $`L`$ odd with $`n`$. As already mentioned in the text, the relative deviation of our numerical results from the conjectures is typically of the order of $`10^{12}10^7`$. In general the accuracy of our computations depends on the choice of the parameters in (27)-(30) and we have searched for regions in the parameter space where the accuracy of our computations breaks down. We found that this is the case for small values of $`\alpha _\pm `$ and $`\beta _\pm `$ in (28) respectively in (29) and (30). As an example figure 3 shows the relative deviation of the extrapolated $`\sigma _j^x`$-profile for the Dirichlet-Neumann boundary condition (28) with $`\alpha _z=0`$ from the exact expression (32) for $`\beta _\pm =\sqrt{8},\beta _\pm =\sqrt{8}/20`$, $`\beta _\pm =\sqrt{8}/100`$ and for even values of $`L`$, where we have used finite size data from $`L=20`$ up to $`L=700`$. The deviations are always in the range of the accuracy of the extrapolations. Hence we avoid any error bars. The behaviour illustrated in figure 3 is not unexpected, since the vanishing of a non-diagonal boundary term implies a sudden change of the respective boundary condition on the free boson from Neumann to Dirichlet. For the boundaries in question this change implies that the $`\sigma _j^x`$ profile vanishes exactly even on the finite chain (the Hamiltonian $`H`$ commutes with $`\sigma _1^z\sigma _2^z\mathrm{}\sigma _L^z`$ for $`\beta _\pm =0`$ and the ground-state of $`H`$ is non-degenerate). Since the profiles on the finite chain are smooth functions of all parameters appearing in the Hamiltonian, the deviation of the finite size data from the profile (32) we conjectured for the case $`\beta _\pm 0`$ in the continuum limit (which does not depend on the value of $`\beta _\pm `$) increases as $`\beta _\pm `$ goes to zero. Assuming that our conjecture is correct, it is not surprising that the accuracy of our extrapolations decreases simultaneously. ## References
warning/0004/hep-th0004032.html
ar5iv
text
# Monopole Chern-Simons Term: Charge-Monopole System as a Particle with Spin ## 1 Introduction The Dirac charge-monopole system was one of the first models with which the importance of topology for physics was realized. The model was investigated in various aspects classically and quantum mechanically -, and the Dirac’s quantization of the charge-monopole constant was understood naturally in terms of the underlying topological fibre bundle structure ,-. In addition, an interesting analysis of the Dirac quantization condition relies on 3-cocycles . Another very seminal interplay of physics and topology is related to the Chern-Simons (CS) “effects” -, whose one of the basic physical aspects established by Deser, Jackiw and Templeton consists in quantization of coupling in field theory analogous to Dirac quantization in quantum mechanics. The spin particle nature of the charge-monopole system was observed by Jackiw, Rebbi, Hasenfrantz and ’t Hooft in the context of the “spin from isospin” field theoretical mechanism , but the first results on the nature of this system anticipating its interpretation as a particle with spin were obtained by Poincaré (see ref. ). In this paper we exploit the CS nature of the charge-monopole coupling term to interpret the charge-monopole system as a particle with spin from the opposite side, not appealing to its field theoretical origin but treating it classically and quantum mechanically as a system with finite number of degrees of freedom. More specifically, we observe that the charge-monopole coupling term having a nature of (0+1)-dimensional CS term manifests its topological nature in two ways. First, the CS term changes the plane dynamical geometry of a free particle for the cone dynamical geometry of a charged particle without distorting the free (geodesic) character of the motion. Second, in the limit of zero charge’s mass the CS term describes a spin system. This observation allows us to interpret the charge-monopole system alternatively: (i) as a free particle of fixed spin with translational and spin degrees of freedom interacting via the helicity constraint; in such interpretation, noncommuting gauge-covariant momenta (velocities) are the analogs of the Foldy-Wouthuysen coordinates of the Dirac particle; (ii) as a symmetric spinning top with dynamical moment of inertia and “isospin” U(1) gauge symmetry; (iii) as a system with higher derivatives; this interpretation is the charge-monopole analog of another known classical equivalence between the relativistic scalar massive particle in a background of the constant homogeneous electromagnetic field and higher derivative model of relativistic particle with torsion underlying (2+1)-dimensional anyons -; (iv) as an analog of the massless particle with nonzero spin, that naturally leads to the twistor formulation for the charge-monopole system. We also show that the reparametrization and scale invariant CS term supplied with the kinetic term of the same invariance gives rise to the alternative description for the spin. The partially gauge fixed version of such spin model corresponds to the charge-monopole system in a spherical geometry. The obtained results are used to discuss the relationship between the charge-monopole system and (2+1)-dimensional anyon. The paper is organized as follows. Section 2 is devoted to discussion of the classical theory of the charge-monopole system in the context of its similarity to the 3D free particle. Here we observe the geodesic character of the charge’s motion and demonstrate that being reduced to the fixed level of the angular momentum integral, it is described by the Lagrangian corresponding to the 2D non-relativistic particle in the planar gravitational field of a point-like source carrying simultaneously a nontrivial magnetic flux. This, in particular, explains why the charge-monopole system and 2D charged particle moving in the field of the point magnetic vortex have the same “hidden” or “dynamical” SO(2,1) symmetry revealed by Jackiw . We obtain the general solution to the canonical equations of motion, compare the charge-monopole system with the 3D free particle from the point of view of integrals of motion and their Lie-Poisson algebras, and, finally, interpret the charge-monopole system as a reduced E(3) system. In section 3 we discuss the classical and quantum theory of the spin represented in the form of the (0+1)-dimensional topological theory given by the CS charge-monopole action corresponding to the limit of zero charge’s mass. In section 4 we construct the description of the charge-monopole system as a particle with spin. We start with the free particle system with spin, whose value is fixed by the charge-monopole constant. Then we switch on interaction between translational and spin degrees of freedom by introducing the helicity constraint which freezes spin degrees of freedom and provides finally the physical equivalence of the extended model to the initial charge-monopole system. In section 5 the system is interpreted as a spinning symmetric top. In this picture the initial U(1) electromagnetic gauge symmetry is changed for the U(1) gauge symmetry generated by the “isospin”-fixing constraint, which makes the rotations about the top’s symmetry axis to be unobservable. The spinning top picture is used in section 6 to get the higher derivative form for the CS term, which, in turn, is employed in section 7 for constructing the twistor formulation for the charge-monopole system proceeding from its analogy to the 4D massless particle with spin. In section 8 we discuss the charge-monopole system in a spherical geometry and find the alternative formulation for the spin. Here we also observe that the SO(2,1) symmetry of the charge-monopole system can be treated formally as a relic of the reparametrization invariance surviving the Lagrangian gauge fixing procedure applied to the Euclidean relativistic version of the model. Section 9 contains discussion of the relationship between the charge-monopole system and (2+1)-dimensional anyon, and in the last section we present some concluding remarks. ## 2 Charge-monopole dynamics and 3D free particle ### 2.1 Lagrangian formalism A non-relativistic particle of unit mass and electric charge $`e`$ in the field of magnetic monopole of charge $`g`$ is described by the Lagrangian $$L=\frac{1}{2}\dot{𝐫}{}_{}{}^{2}+e𝐀\dot{𝐫},$$ (2.1) with a U(1) gauge potential $`𝐀(𝐫)`$ defined by the relations $$_iA_j_jA_i=F_{ij}=ϵ_{ijk}B_k,B_i=g\frac{r_i}{r^3},r=\sqrt{𝐫^2}.$$ (2.2) The case of arbitrary mass $`m`$ can be obtained from (2.2) via the transformation of time and charge: $`tm^{1/2}t`$, $`em^{1/2}e`$. In definition (2.2) it is assumed that the point $`r=0`$ is excluded, i.e. the configuration space of the system, $`=\text{IR }^3\{0\}`$, is diffeomorphic to $`(0,\mathrm{})\times S^2`$ and inherits a nontrivial topology of a two-sphere $`S^2`$. It is well known that the electromagnetic potential (2.2) gives a connection of the monopole U(1) fibre bundle being a nontrivial Hopf bundle over $`S^2`$, and the problems with Dirac strings of singularity can be escaped by covering the configuration space $``$ with two charts . These topological complications, however, do not play any role for us under treating the classical dynamics of the system, but will reveal themselves at the quantum level. Equations of motion following from Lagrangian (2.1) result in the Lorentz force law, $$\ddot{𝐫}=\frac{\nu }{r^3}𝐫\times \dot{𝐫},\nu =eg,$$ (2.3) which implies that instead of the orbital angular momentum vector $`𝐋=𝐫\times \dot{𝐫},`$ the vector $$𝐉=𝐋\nu 𝐧,𝐧=𝐫r^1,$$ (2.4) is the conserved angular momentum of the system. The nontrivial term $`\nu 𝐧`$ can be understood as the electromagnetic angular momentum produced by both the electric charge and magnetic monopole (see ref. ). Note that $`J=\sqrt{𝐉^2}`$ is restricted from below by the modulus of the charge-monopole coupling constant $`\nu `$, $`J|\nu |`$. Due to the relation $`\mathrm{𝐉𝐧}=\nu `$, the trajectory of the particle lies on the cone. The cone’s axis is given by the vector $`𝐉`$ and its half-angle is $$\mathrm{cos}\gamma =\nu J^1.$$ (2.5) Since the force $`𝐟=\nu r^3𝐫\times \dot{𝐫}`$ is orthogonal to $`𝐫`$ and to the velocity $`\dot{𝐫}`$, it is perpendicular to the cone. Therefore, the particle performs a free motion on the cone. This can also be observed directly as follows. Taking into account the conservation of the angular momentum $`𝐉`$ and that in spherical coordinates the charge-monopole coupling term takes a form $`L_{int}=\nu \mathrm{cos}\vartheta \dot{\phi }`$ (see Eq. (3.9)), we can choose the system of coordinates with axis $`\vartheta =0`$ directed along the vector $`𝐉`$. Then in accordance with Eq. (2.5), $`\mathrm{cos}\vartheta =\nu J^1`$, $`\dot{\vartheta }=0`$, and Lagrangian (2.1) is reduced to $`L=\frac{1}{2}(\dot{r}{}_{}{}^{2}+(1\nu ^2J^2)r^2\dot{\phi }{}_{}{}^{2})\nu ^2J^1\dot{\phi }`$. After transformation $$tt^{}=\alpha t,\alpha =\sqrt{1\nu ^2J^2},$$ (2.6) we obtain finally the following form for the charge-monopole Lagrangian: $$L=\frac{1}{2}(\alpha ^2\dot{r}{}_{}{}^{2}+r^2\dot{\phi })\alpha \nu ^2J^1\dot{\phi }.$$ (2.7) The last term is the total time derivative of the topologically nontrivial angular variable. It is the reduced form of the charge-monopole interaction term having a nature of $`(0+1)`$-dimensional CS term -. As we shall see, quantum mechanically this will give rise to the Dirac quantization condition for $`\nu `$, but classically the total derivative can be omitted without changing the equations of motion. The topologically nontrivial term corresponds exactly to the 2D term describing the interaction of the charge $`e`$ with a (singular) point vortex carrying the magnetic flux $`\mathrm{\Phi }=2\pi \alpha \nu ^2J^1e^1.`$ Without the last term, Eq. (2.7) is a Lagrangian of a free particle of unit mass on the cone given by the relations $`x=r\mathrm{cos}\phi ,`$ $`y=r\mathrm{sin}\phi ,`$ $`z=r\sqrt{\alpha ^21},`$ $`r>0,`$ $`0\phi 2\pi ,`$ with $`0<\alpha <1`$, that confirms our statement on a free (geodesic) motion of the charge over the cone (2.5). Together with the pointed out nature of the total derivative term, this, as we shall see, explains why the charge-monopole system and 2D charged particle moving in the field of the point magnetic vortex have the same dynamical SO(2,1) symmetry . Since the conical metric $`ds^2=\alpha ^2(dr)^2+r^2(d\phi )^2`$ corresponds to the metric produced by the point mass , we can say that the classical motion of the charge in the field of magnetic monopole (reduced to the fixed value of the integral $`𝐉`$) is equivalent to the classical motion of a particle in a planar gravitational field of a point massive source carrying simultaneously magnetic flux $`\mathrm{\Phi }=2\pi \alpha \nu ^2J^1e^1.`$ From the form of transformation (2.6) and Lagrangian (2.7) it is clear that the case $`J=|\nu |`$ is singular and should be treated as a limit case, i.e. we have to assume that $`J>|\nu |`$. We shall discuss this peculiarity of the charge-monopole system in different aspects in what follows. In the case of a 3D free particle ($`\nu =0`$, $`=\text{IR }^3`$), the motion is characterized by the coordinate $`𝐫`$ and by the conserved linear momentum $`𝐩`$. Alternatively, with the appropriate choice of the origin of the system of coordinates, the particle’s motion can be characterized by the unit vector $`𝐧`$ and by the conserved orbital angular momentum $`𝐋`$ supplemented with the canonically conjugate scalars $`r`$ and $`p_r=\mathrm{𝐩𝐫}r^1`$. Since $`\mathrm{𝐋𝐧}=0`$, for a given $`𝐋`$ the particle’s trajectory is in the plane orthogonal to the orbital angular momentum. So, we conclude that classically the topological nature of $`(0+1)`$-dimensional charge-monopole CS term is manifested in changing the global structure of the dynamics without distorting its local free (geodesic) character: the “plane dynamical geometry” of the free particle ($`e=0`$) is changed for the free “cone dynamical geometry” of the charged particle. We shall discuss the relation between the two systems in the context of (dynamical) integrals of motion in subsection 2.3. ### 2.2 Canonical formalism: solutions to the equations of motion To solve the equations of motion in general form and analyze in more detail the system’s dynamics, we turn to the canonical formalism. The Hamiltonian corresponding to Lagrangian (2.2) is $`H=\frac{1}{2}𝐏^2,`$ where $`𝐏=𝐩e𝐀`$ is a classical analog of the gauge-covariant derivative defined via the momentum $`𝐩`$ canonically conjugate to $`𝐫`$. This gives rise to the Poisson brackets $$\{P_i,P_j\}=\frac{\nu }{r^3}ϵ_{ijk}r_k,\{r_i,P_j\}=\delta _{ij},\{r_i,r_j\}=0,$$ (2.8) and to the equations of motion $$\dot{𝐫}=𝐏,\dot{𝐏}=\frac{\nu }{r^3}𝐋,$$ (2.9) with $`𝐋=𝐫\times 𝐏`$. From Eq. (2.9) we find that the vectors $`𝐧`$ and $`𝐋`$ evolve according to the equations $`\dot{𝐋}=\nu r^2𝐋\times 𝐧,`$ $`\dot{𝐧}=r^2𝐋\times 𝐧,`$ and, as a consequence, precess about the integral $`𝐉=𝐋\nu 𝐧`$ with the same frequency: $`\dot{𝐋}=r^2𝐉\times 𝐋,`$ $`\dot{𝐧}=r^2𝐉\times 𝐧.`$ The vectors $`𝐉`$ and $`𝐧`$ obeying the relations $`𝐧^2=1`$ and $`\mathrm{𝐉𝐧}=\nu `$, together with the two scalars $`𝐫^2`$ and $`\mathrm{𝐏𝐫}`$ form the complete set of observables in terms of which $`𝐫`$ and $`𝐏`$ can be completely “restored”. The equations of motion for the scalar observables have a simple form $`(𝐫{}_{}{}^{2})\dot{}=2\mathrm{𝐏𝐫},`$ $`(\mathrm{𝐏𝐫})\dot{}=𝐏^2=2H,`$ being exactly the same as in the case of a free particle. Their integration gives $$(\mathrm{𝐏𝐫})(t)=(\mathrm{𝐏𝐫})(t_0)+𝐏^2(tt_0),𝐫^2(t)=𝐏^2(tt_0)^2+2(\mathrm{𝐏𝐫})(t_0)(tt_0)+𝐫^2(t_0).$$ (2.10) The minimal charge-monopole distance corresponds to the moment of time for which $`\mathrm{𝐏𝐫}=0`$, and is given by the relation $`𝐫_{\mathrm{min}}^2=𝐋^2𝐏^2`$. To complete the integration of equations of motion, we solve the equation $`\dot{𝐧}=r^2𝐉\times 𝐧`$ and get $`𝐧(t)=\nu J^1𝐣+𝐧_{}(t),𝐧_{}(t)=\left(𝐧(t_0)+𝐣\nu J^1\right)\mathrm{cos}\tau (t)+𝐣\times 𝐧(t_0)\mathrm{sin}\tau (t),`$ (2.11) $`\tau =JL^1\mathrm{tan}^1\left(\mathrm{𝐏𝐫}L^1\right),L=\sqrt{𝐉^2\nu ^2},`$ (2.12) where $`𝐣=𝐉J^1`$, and evolution of $`\mathrm{𝐏𝐫}`$ is given by Eq. (2.10). Zero value of the angular function $`\tau (t)`$ corresponds to the point of perihelion ($`r=r_{min}`$) with respect to which the trajectory is symmetric. The vector $`𝐧_{}`$ is the projection of $`𝐧`$ to the plane orthogonal to the angular momentum $`𝐉`$, and Eqs. (2.11), (2.12) give the classical scattering angle of the particle’s motion projected into the plane perpendicular to $`𝐉`$ as a function of it: $$\phi _{scat}^{}=\tau (+\mathrm{})\tau (\mathrm{})=\pi JL^1.$$ (2.13) Eq. (2.11) together with Eq. (2.10) and relation $`𝐏=\dot{𝐫}`$ give a complete solution to the equations of motion (2.9). In correspondence with Eqs. (2.5), (2.13), in the limit $`J\mathrm{}`$ the cone over which the particle moves is close to the plane: $`\gamma \pi /2`$ and $`\phi _{scat}^{}\pi `$. In another limit, $`J|\nu |`$, the cone is degenerated into a half-line, $`\gamma 0(\pi )`$ for $`\nu <0(>0)`$, whereas the number of full rotations $`𝒩=[\phi _{scat}^{}/2\pi ]`$ is infinite, where $`[.]`$ is the integer part. Therefore, the case $`J=|\nu |`$ ($`𝐋=0`$) with $`𝐏0`$, which corresponds to the motion of the charge with constant velocity along a straight half-line defined by $`𝐧(t_0)`$ in the direction to or from the monopole is not a proper limit: such a trajectory corresponds to the half of the trajectory with $`J\mathrm{}`$ ($`L\mathrm{}`$) but not to the limit $`J|\nu |`$. This supports our conclusion of the previous subsection on a necessity to treat the values of the angular momentum to be confined to the domain $`|\nu |<J<\mathrm{}`$. ### 2.3 Integrals of motion and their algebra Let us compare the charge-monopole system with a 3D free particle from the point of view of integrals of motion and corresponding Lie-Poisson algebras formed by them. For the $`3D`$ free particle ($`\nu =0`$), the first integrals of motion are linear, $`𝐩`$, and angular, $`𝐋=𝐫\times 𝐩`$, momenta forming the set of 5 ($`\mathrm{𝐩𝐋}=0`$) algebraically independent conserved quantities not depending explicitly on time. With respect to the Poisson brackets, they form the algebra of Euclidean group $`E(3)`$. Algebraically, the Hamiltonian is a square of the vector $`𝐩`$, $`H=\frac{1}{2}𝐩^2`$, but it generates the symmetry of time translations being independent from the symmetries of space translations and rotations generated by $`𝐩`$ and $`𝐋`$. This is not the only independent symmetry which can be generated via constructing algebraically dependent quantities from $`𝐋`$ and $`𝐩`$. Indeed, the vector $`𝐐=𝐩\times 𝐋`$ is the analog of the Laplace-Runge-Lenz vector of the Coulomb-Kepler system, which can be treated as a generator of the corresponding canonical symmetry transformations. The integrals $`𝐋`$ and $`𝐐`$, $`\mathrm{𝐋𝐐}=0`$, form together with the Hamiltonian the nonlinear algebra: $$\{L_i,V_j\}=ϵ_{ijk}V_k,\{Q_i,Q_j\}=2Hϵ_{ijk}L_k,\{H,V_i\}=0,V_i=L_i,Q_i.$$ The renormalized (at $`𝐩0`$) vector $`𝐑𝐐/\sqrt{𝐩^2}`$ and the angular momentum vector $`𝐋`$ form the Lorentz algebra $`so(3,1)`$. One can construct another vector, $`𝐊𝐋\times 𝐑L^1`$, $`L=\sqrt{𝐋^2}`$, which together with $`𝐑`$ and $`𝐋`$ provides us with the complete set of the three orthogonal vectors, $`\mathrm{𝐑𝐋}=\mathrm{𝐊𝐋}=\mathrm{𝐑𝐊}=0`$, of the same norm, $`𝐑^2=𝐊^2=𝐋^2`$. They form the following nonlinear algebra: $$\{L_i,V_j\}=ϵ_{ijk}V_k,\{R_i,R_j\}=\{K_i,K_j\}=ϵ_{ijk}L_k,\{R_i,K_j\}=\delta _{ij}L,$$ (2.14) where $`V_i=L_i,R_i,K_i`$. One notes that the scalar $`L`$ rotates the set of vectors $`𝐑`$ and $`𝐊`$: $$\{L,R_i\}=K_i,\{L,K_i\}=R_i.$$ (2.15) Therefore, the vector integral $`𝐊`$ also forms the $`so(3,1)`$ algebra with the orbital angular momentum and is (non-canonically) conjugate to $`𝐑`$ (see the last Poisson bracket relation in Eq. (2.14)). The rotated about $`𝐋`$ vectors $`𝐑^{}=𝐑\mathrm{cos}\phi +𝐊\mathrm{sin}\phi `$ and $`𝐊^{}=𝐊\mathrm{cos}\phi 𝐑\mathrm{sin}\phi `$, $`\phi =const`$, possess exactly the same set of properties as $`𝐑`$ and $`𝐊`$. There are also the so called dynamical integrals of motion depending explicitly on time and they appear as follows. Solving the equations of motion $`\dot{𝐫}=𝐩`$, $`\dot{𝐩}=0`$, one gets $`𝐫=𝐩(tt_0)+𝐗,`$ $`𝐗𝐫(t_0).`$ One can treat $`𝐗=𝐫𝐩(tt_0)`$ as a vector integral of motion dependent explicitly on time: $`\frac{d}{dt}𝐗=\{𝐗,H\}+𝐗/t=0`$. It generates the transformations $`x_ix_i\epsilon _i(tt_0)`$ corresponding to the Galilei boosts. The integrals of motion $`𝐩`$ and $`𝐗`$ satisfy the Heisenberg algebra $`\{X_i,X_j\}=\{p_i,p_j\}=0`$, $`\{X_i,p_j\}=1\delta _{ij}`$. Algebraically, the vector $`𝐗`$ is equivalent to the vector integral not containing the explicit dependence on time, $`𝐗_{}=𝐫𝐩(\mathrm{𝐩𝐫})𝐩^2=𝐐𝐩^2,`$ $`𝐗_{}𝐩=0,`$ and to the scalar $`D=\mathrm{𝐗𝐩}=\mathrm{𝐫𝐩}𝐩^2(tt_0)`$ being a dynamical integral of motion generating the time dilations . One notes that the angular momentum $`𝐋`$ can also be treated as an integral algebraically dependent on $`𝐗`$ and $`𝐩`$: $`𝐋=𝐗\times 𝐩=𝐗_{}\times 𝐩`$. On the other hand, from the equations of motion it follows that $`\frac{d}{dt}𝐫^2=2\mathrm{𝐩𝐫},`$ $`\frac{d}{dt}(\mathrm{𝐩𝐫})=𝐩^2.`$ Integration of the second equation gives rise to the dynamical integral $`D`$ and the subsequent integration of the first equation gives, similar to Eq. (2.10), $`𝐫^2(t)=𝐩^2(tt_0)^2+2D(tt_0)+𝐫^2(t_0)`$. The last relation can be rewritten in the form of the dynamical integral of motion $`=𝐫^2(t_0)=𝐫^2𝐩^2(tt_0)^22D(tt_0),`$ which generates the time special conformal transformations . It can be represented equivalently as $$=(𝐋^2+D^2)𝐩^2.$$ (2.16) The integrals $`D`$, $`2H`$ and $``$ form the same algebra as the scalars $`\mathrm{𝐩𝐫}`$, $`𝐩^2`$ and $`𝐫^2`$ (the latter set is reduced to these integrals at the initial moment $`t=t_0`$), which is the $`so(2,1)sl(2,R)`$ algebra -: $$\{𝒥_0,𝒥_1\}=𝒥_2,\{𝒥_0,𝒥_2\}=𝒥_1,\{𝒥_1,𝒥_2\}=𝒥_0,$$ (2.17) where $`𝒥_0=\frac{1}{4}(2H+)`$, $`𝒥_1=\frac{1}{4}(2H)`$, $`𝒥_2=\frac{1}{2}D`$, with the Casimir central element $`𝒞=𝒥_0^2+𝒥_1^2+𝒥_2^2=\frac{1}{4}𝐋^20`$. The dynamical integral $`D`$ together with $`H`$ commutes in the sense of Poisson brackets with the vector integrals $`𝐋`$, $`𝐑`$ and $`𝐊`$, whereas the integral $``$ due to Eqs. (2.15), (2.16) has nontrivial Poisson bracket relations with $`𝐑`$ and $`𝐊`$. Let us turn to the charge-monopole system, where instead of the orbital angular momentum, the vector $`𝐉=𝐋\nu 𝐧`$ is conserved. Since the equations of motion for the scalar variables $`𝐫^2`$ and $`\mathrm{𝐏𝐫}`$ look exactly as the corresponding equations for the free particle (with the change of $`𝐩`$ for $`𝐏`$), the charge-monopole system has the set of the scalar dynamical integrals of motion of the same form , $$D=\mathrm{𝐏𝐫}𝐏^2(tt_0),=𝐫^2𝐏^2(tt_0)^22D(tt_0).$$ (2.18) Though $`P_i`$ are characterized by the nontrivial Poisson brackets (2.8), the scalar dynamical integrals (2.18) and the Hamiltonian generate the same $`so(2,1)`$ algebra (2.17) as in a free case with the Casimir central element $$𝒞=𝒥_0^2+𝒥_1^2+𝒥_2^2=\frac{1}{4}(𝐉^2\nu ^2)<0.$$ (2.19) The free particle’s vector integrals of motion $`𝐑`$ and $`𝐊`$ also have analogs in the charge-monopole system. These are given by the vectors $$𝐑𝐍\frac{J^2}{\sqrt{J^2\nu ^2}},𝐊𝐉\times 𝐍\frac{J}{\sqrt{J^2\nu ^2}},$$ (2.20) where the vector integral $`𝐍=\left(𝐧+\nu J^1𝐣\right)\mathrm{cos}\tau 𝐣\times 𝐧\mathrm{sin}\tau `$ can be identified with the vector $`𝐧_{}(t_0)`$ (see Eq. (2.11)). Vector $`𝐍`$ satisfies the relations $`\mathrm{𝐍𝐉}=0`$, $`𝐍^2=1\nu ^2J^2`$, and $`\{N_i,N_j\}=\nu ^2J^4ϵ_{ijk}J_k,`$ $`\{N_i,𝐫^2\}=\{N_i,\mathrm{𝐏𝐫}\}=\{N_i,𝐏^2\}=0.`$ From these relations we find that $`\mathrm{𝐑𝐉}=\mathrm{𝐊𝐉}=\mathrm{𝐑𝐊}=0`$, $`𝐑^2=𝐊^2=𝐉^2`$, and that the Poisson bracket algebra of the complete set of orthogonal vectors $`𝐉`$, $`𝐑`$ and $`𝐊`$ is given by Eq. (2.14) with $`𝐋`$ changed for $`𝐉`$. This set of vector integrals is in involution with the dynamical scalar integral of motion $`D`$ and with $`H`$, but the vectors $`𝐑`$ and $`𝐊`$ have nontrivial Poisson bracket relations with the dynamical integral $`=(D^2+𝐉^2\nu ^2)/2H`$ (see Eq. (2.15)). Therefore, the dynamical geometry similarity of the charge-monopole system to the 3D free particle discussed in subsection 2.1 also reveals itself in existence of similar sets of integrals of motion (depending and not depending explicitly on time), which form between themselves the same (nonlinear) Lie-Poisson algebras. ### 2.4 Charge-monopole as a reduced E(3) system Like a 3D free particle, the charge-monopole system may be treated as a reduced E(3) system. To get such an interpretation, let us pass over from the Hamiltonian variables $`𝐫`$ and $`𝐏`$ to the set of variables $`𝐧`$, $`𝐉`$, $`r`$ and $`P_r=\mathrm{𝐏𝐫}r^1`$. They have the following Poisson brackets: $`\{r,P_r\}=1,\{r,𝐧\}=\{r,𝐉\}=\{P_r,𝐧\}=\{P_r,𝐉\}=0,`$ (2.21) $`\{J_i,J_j\}=ϵ_{ijk}J_k,\{J_i,n_j\}=ϵ_{ijk}n_k,\{n_i,n_j\}=0.`$ (2.22) Poisson brackets (2.22) correspond to the algebra of generators of the Euclidean group E(3) with $`J_i`$ being a set of generators of rotations and $`n_i`$ identified as generators of translations. The quantities $`𝐧^2`$ and $`\mathrm{𝐉𝐧}`$ lying in the center of $`e(3)`$ algebra, $`\{𝐧^2,n_i\}=\{𝐧^2,J_i\}=\{\mathrm{𝐧𝐉},n_i\}=\{\mathrm{𝐧𝐉},J_i\}=0`$, are fixed in the present case by the relations $$𝐧^2=1,\mathrm{𝐧𝐉}=\nu .$$ (2.23) In terms of the introduced variables, the Hamiltonian of the system takes the form $$H=\frac{1}{2}P_r^2+\frac{(𝐉\times 𝐧)^2}{2r^2}.$$ (2.24) Therefore, the charge-monopole system can be treated as the E(3) system reduced by the conditions (2.23) fixing the Casimir elements and supplemented by the independent canonically conjugate variables $`r`$ and $`P_r`$. It is the second relation from Eq. (2.23) that encodes the topological difference between the charge-monopole and the 3D free particle cases: for $`\nu 0`$, the space given by the spin vector $`𝐉`$ is homeomorphic to $`\text{IR }^3\{0\}`$, ($`J>|\nu |`$), whereas for $`\nu =0`$ the corresponding space $`\text{IR }^3`$ is topologically trivial. In the next section we shall discuss the physical consequences of the nontrivial topological structure of the charge-monopole system. In accordance with Eqs. (2.23), one could treat the vector $`𝐋=𝐉+\nu 𝐧`$, $`\mathrm{𝐋𝐧}=0`$, as an orbital angular momentum. However, the Poisson bracket relations $`\{L_i,L_j\}=ϵ_{ijk}(L_k+\nu n_k)`$ following from (2.22) and restriction $`𝐋^2>0`$ corresponding to $`J>|\nu |`$ prevent such interpretation. Nevertheless, as we shall see, it is possible to treat $`𝐋`$ as an orbital angular momentum in extended physically equivalent formulation of the model. ## 3 Monopole Chern-Simons term and spin This section contains mainly the known results which are necessary for the self-contained presentation of the subsequent analysis. The integrand in action corresponding to the the charge-monopole interaction term, $`\theta =e\dot{𝐫}𝐀(𝐫)dt`$, can be treated as a differential one-form, $`\theta =e𝐀(𝐫)d𝐫.`$ Then the relations (2.2) defining the monopole vector potential are equivalent to the relation $$d\theta =\frac{\nu }{2r^3}ϵ_{ijk}r_idr_jdr_k.$$ (3.1) The right-hand side of Eq. (3.1) is the gauge-invariant curvature two-form, $$d\theta =e,=\frac{1}{2}F_{ij}dr_idr_j$$ (3.2) and, consequently, the gauge-non-invariant one-form $`\theta `$ has a sense of $`(0+1)`$-dimensional CS term -. The two-form (3.1) can be represented equivalently as $$d\theta =\frac{\nu }{2}ϵ_{ijk}n_idn_jdn_k,n_i=r_ir^1.$$ (3.3) Via the (local) parametrization by the spherical angles, $`𝐧=𝐧(\vartheta ,\phi )`$, this can be treated as the differential area of a two-sphere multiplied by the charge-monopole coupling constant: $`d\theta =\nu d(\mathrm{cos}\vartheta )d\phi `$. If the vector $`𝐧(t)`$ describes some closed curve $`\mathrm{\Gamma }`$ on the sphere (i.e. if $`𝐧(t_1)=𝐧(t_2)`$)<sup>1</sup><sup>1</sup>1 According to Eqs. (2.11), (2.12) and relation $`(\mathrm{𝐏𝐫})\dot{}=𝐏^2=const`$, such a closed curve is smooth., the Stokes theorem gives $`_\mathrm{\Gamma }\theta =_{S_+}𝑑\theta =_S_{}𝑑\theta ,`$ where $`\mathrm{\Gamma }=S_+=S_{}`$, $`S_+S_{}=S^2`$. Within the path-integral quantization scheme, the alternative representations for the same charge-monopole interaction term in the action can differ only in $`2\pi n,`$ $`n\text{ZZ}`$, $$_{S_+}𝑑\theta \left(_S_{}𝑑\theta \right)=\nu _{S^2}d\mathrm{cos}\vartheta d\phi =4\pi \nu =2\pi n,$$ (3.4) and we arrive at the Dirac quantization condition for the charge-monopole coupling constant: $`2\nu =n\text{ZZ}`$. Defining the dependent variables $`s_i=\nu n_i,`$ one can represent (3.3) equivalently as $$\omega _{spin}=d\theta =\frac{1}{2𝐬^2}ϵ_{ijk}s_ids_jds_k,s_is_i=\nu ^2.$$ (3.5) The two-form (3.5) is closed and nondegenerate, and can be treated as a symplectic form corresponding to the symplectic potential $`\theta `$. If we drop out the kinetic term in the charge-monopole action (that corresponds to taking the charge’s zero mass limit, $`m0`$, ), we get the CS action $$S=_{t_1}^{t_2}\theta ,$$ (3.6) describing the spin system. Indeed, for any function $`f`$ on the sphere $`S^2`$, the symplectic form (3.5) defines the Hamiltonian vector field $`X_f`$ on the cotangent bundle $`T^{}S^2`$: $`i_{X_f}\omega (Y)=\omega (X_f,Y)=df.`$ This gives $`X_f=_af\omega ^{ab}\frac{}{x^b}`$, and defines the corresponding Poisson brackets, $`\{f,g\}=\omega (X_f,X_g).`$ Here $`x^a`$, $`a=1,2`$, are the local coordinates on $`S^2`$ and $`\omega ^{ab}`$ are the elements of the matrix inverse to the symplectic matrix $`\omega _{ab}`$, $`\omega _{spin}=\frac{1}{2}\omega _{ab}dx^adx^b`$. Taking into account Eq. (3.5), we get $`\{s_i,s_k\}=ϵ_{ijk}s_k,`$ $`s_is_i=\nu ^2.`$ These relations define the classical spin system with fixed spin modulus. Geometric quantization applied to such a system (for the details see ref. ) leads to the same Dirac quantization of the parameter $`\nu `$, $`|\nu |=j`$, $`j=1/2,1,3/2,\mathrm{}`$, and results in $`(2j+1)`$-dimensional representation of $`su(2)`$, $$s_1=\frac{1z^2}{2}\frac{d}{dz}+jz,s_2=i\frac{1+z^2}{2}\frac{d}{dz}ijz,s_3=z\frac{d}{dz}j,$$ (3.7) realized on the space of holomorphic functions with the basis $`\psi _j^kz^{j+k}`$, $`k=j,j+1,\mathrm{},+j`$, $`s_3\psi _j^k=k\psi _j^k`$, and scalar product $$(\psi _1,\psi _2)=\frac{2j+1}{\pi }\frac{\overline{\psi _1(z)}\psi _2(z)}{(1+|z|^2)^{2j+2}}d^2z.$$ The classical relation $`𝐬^2=\nu ^2`$ is changed for the quantum relation $`𝐬^2=j(j+1)`$. Here the complex variable $`z`$ is related to the spherical angles via the stereographic projection $`z=\mathrm{tan}\frac{\vartheta }{2}e^{i\phi }`$ from the north pole, or via $`z=\mathrm{cot}\frac{\theta }{2}e^{i\phi }`$ for the projection from the south pole. In both cases the symplectic two-form is represented as $$\omega _{spin}=2i\nu \frac{d\overline{z}dz}{(1+\overline{z}z)^2}.$$ (3.8) Geometrically, the obtained spin system is a Kähler manifold with Kähler potential $`𝒦=2i\nu \mathrm{ln}(1+\overline{z}z)`$: $`\omega _{spin}=\frac{^2}{z\overline{z}}𝒦d\overline{z}dz`$, $`\overline{z}=z^{}`$. Locally, in spherical coordinates the spin Lagrangian is given by $$L_{spin}=\nu \mathrm{cos}\vartheta \dot{\phi },$$ (3.9) and in terms of global complex variable the Lagrangian (3.9) takes the form $$L_{spin}=i\nu \frac{\overline{z}\dot{z}\dot{\overline{z}}z}{1+\overline{z}z}.$$ (3.10) The appearence of the two stereographic projections for the spin system (3.6) reflects the above mentioned necessity to work in two charts in the case of the initial ($`m0`$) charge-monopole system to escape the problems with Dirac string singularities. In terms of globally defined independent variables $`z`$, $`\overline{z}`$ no gauge invariance left in the spin system given by the Lagrangian (3.10) but it is hidden in a fibre bundle structure reflected, in particular, in the presence of two charts . We conclude that the charge-monopole interaction term has a nature of (0+1)-dimensional Abelian CS term, that leads to the Dirac quantization of the charge-monopole coupling constant, $`|\nu |=j`$. In the limit of zero charge’s mass the total charge-monopole Lagrangian is reduced to the Lagrangian (3.10) in terms of independent variables $`z`$ and $`\overline{z}`$, which describes the spin-$`j`$ system. In what follows, we consider other possibilities to describe spin system proceeding from its nature associated with the monopole CS term. ## 4 Charge-monopole as a particle with spin The spin nature of the charge-monopole CS term and observed free character of the charge’s dynamics allow us to get the alternative description for the charge-monopole system as a free particle of fixed spin with translational and spin degrees of freedom interacting via the helicity constraint. To find such a description, we forget for the moment that the spin system has been obtained via the identification $`s_i=\nu n_i,`$ (in the limit $`m0`$), and simply start with its Hamiltonian description given by the symplectic form (3.5) and corresponding brackets $`\{s_i,s_k\}=ϵ_{ijk}s_k.`$ The canonical Hamiltonian of the spin theory given by the first order action (3.6) or by the Lagrangian (3.10) is equal to zero. Let us extend such a pure spin system by adding to it independent translational degrees of freedom described by the particle’s coordinates $`r_i`$ and canonically conjugate momenta $`p_i`$, i.e. we suppose that the corresponding symplectic two-form is $$\omega =dp_idr_i+\omega _{spin}.$$ (4.1) Moreover, let the dynamics of the system is given by the Hamiltonian $`H=\frac{1}{2}𝐩^2.`$ Then this Hamiltonian and relations (3.5), (4.1) specify the nonrelativistic free particle with internal degrees of freedom describing spin of fixed value. In such a system we have $`6+2`$ independent phase space variables instead of 6 variables in the initial charge-monopole system. We can reduce such an extended system to the initial system by introducing into it one first class constraint. Having in mind the identification $`s_i=\nu n_i,`$ for the initial system (2.1), let us postulate the helicity constraint $$\chi \mathrm{𝐬𝐫}+\nu r0.$$ (4.2) This constraint can be interpreted as a constraint introducing interaction between translational and spin degrees of freedom. But the constraint (4.2) is not conserved by the Hamiltonian, $`\frac{1}{2}\{𝐩^2,\chi \}0`$, and for consistency of such a theory we have to modify the latter. For the purpose, let us find the complete set of gauge-invariant variables commuting in the sense of Poisson brackets with constraint (4.2). Since we have $`6+2`$ phase space variables and one constraint, there are 6 independent gauge-invariant variables. They are $`r_i`$ and $$\mathrm{\Pi }_ip_i\frac{1}{r^2}ϵ_{ijk}r_js_k,$$ (4.3) $`\{r_i,\chi \}=0`$, $`\{\mathrm{\Pi }_i,\chi \}0`$. Therefore, it is natural to change $`H=\frac{1}{2}𝐩^2`$ for its gauge-invariant analog, $`H=\frac{1}{2}𝚷^2,`$ and to take the sum of it and constraint (4.2) multiplied by an arbitrary function $`\lambda =\lambda (t)`$, $$H=\frac{1}{2}𝚷^2+\lambda (\mathrm{𝐬𝐫}+\nu r),$$ (4.4) as a total Hamiltonian . The Poisson brackets for the gauge-invariant variables are $$\{\mathrm{\Pi }_i,\mathrm{\Pi }_j\}=\frac{\mathrm{𝐫𝐬}}{r^4}ϵ_{ijk}r_k,\{\mathrm{\Pi }_i,r_j\}=\delta _{ij},\{r_i,r_j\}=0.$$ (4.5) Taking into account constraint (4.2), the Poisson brackets between $`\mathrm{\Pi }_i`$ and $`\mathrm{\Pi }_j`$ coincide with the Poisson brackets (2.8) between $`P_i`$ and $`P_j`$. Identifying $`\mathrm{\Pi }_i`$ with $`P_i`$, the dynamics generated by the Hamiltonian (4.4) for the gauge-invariant variables $`r_i`$, $`\mathrm{\Pi }_i`$ is exactly the same as the dynamics in the initial charge-monopole system (2.1), and the physical content of the extended system (4.4) is the same as that of the initial system (2.1). In particular, the vector $$𝐉=𝐫\times 𝐩+𝐬$$ (4.6) is the integral of motion, whose components satisfy the Poisson bracket relations $`\{J_i,J_j\}=ϵ_{ijk}J_k,`$ and generate rotations. It can be represented equivalently as $`𝐉=𝐫\times 𝚷+(\mathrm{𝐧𝐬})𝐧,`$ and on the constraint surface (4.2) takes the form $`𝐉𝐫\times 𝚷\nu 𝐧,`$ which coincides with (2.4) with identified $`\mathrm{\Pi }_i`$ and $`P_i`$. With the help of relation (4.6), the Hamiltonian (4.4) can be written down equivalently (cf. with Eq. (2.24)), $$H=\frac{1}{2}p_r^2+\frac{(𝐉\times 𝐧)^2}{2r^2}+\lambda (\mathrm{𝐉𝐧}+\nu ),$$ (4.7) where $`p_r=\mathrm{𝐩𝐫}r^1`$ is the momentum canonically conjugate to the radial variable $`r`$. Since $`\mathrm{𝐋𝐧}=0`$, $`𝐋=𝐫\times 𝐩`$, the difference of the present Hamiltonian interpretation of the charge-monopole system as a particle with spin from the reduced E(3) system from section 2.4 is that here the helicity is fixed weakly by the constraint (4.2), whereas there it was fixed strongly by the second relation from Eq. (2.23). As a consequence, here the components of the angular momentum vector $`𝐋`$ form the $`so(3)`$ algebra, $`\{L_i,L_j\}=ϵ_{ijk}L_k`$, but they, unlike the total angular momentum vector $`𝐉`$, are not physical variables: $`\{L_i,\chi \}0`$. In the present interpretation, we have additional spin variables $`s_i`$ restricted by the condition $`𝐬^2=\nu ^2`$, but the only physical observable constructed from them is the combination $`\mathrm{𝐬𝐧}`$, which is fixed by the helicity constraint. The important comment is in order here. Strictly speaking, in correspondence with the discussion above on the necessity of restriction $`J^2>\nu ^2`$, we have to suppose that the relation $`𝐬^2=\nu ^2`$ has to be treated only in the sense of the limit $`𝐬^2=\nu ^2+\epsilon ^2`$, $`\epsilon 0`$. This will not only correspond to the specified restriction on $`J^2`$ in accordance with relations (4.2) and (4.6), but is necessary for the consistent treatment of the extended model as a constrained system. Indeed, in correspondence with general theory of gauge systems , only in this case the constraint (4.2) can be treated as a good constraint condition, which in the two-dimensional spin phase subspace given by $`s_i`$, $`\{s_i,s_j\}=ϵ_{ijk}s_k`$, $`𝐬^2=\nu ^2+\epsilon ^2`$, specifies one-dimensional physical subspace (on which it will act transitively). The Lagrangian corresponding to the described extended system is $$L=\frac{1}{2}\dot{𝐫}{}_{}{}^{2}\frac{1}{r^2}(𝐫\times \dot{𝐫})𝐬\lambda (\mathrm{𝐫𝐬}+\nu r)+L_{spin},$$ (4.8) with $`L_{spin}`$ chosen in the form (3.10) and $`\lambda `$ treated as a Lagrange multiplier. We conclude that the system (4.8) describing the nonrelativistic particle of spin $`\nu `$ with interacting translational and spin degrees of freedom (see the second $`\mathrm{𝐋𝐬}`$-coupling term and the third constraint term in Lagrangian) is classically equivalent to the charge-monopole system (2.1). The quantum theory of the charge-monopole system in such interpretation is obvious. As we have seen, the quantization of the spin variables results in the Dirac condition, $`\nu =ϵj`$, $`ϵ=+`$ or $``$, and gives rise to the corresponding $`(2j+1)`$-dimensional representation of the $`su(2)`$ with spin operators (3.7) acting on the space of holomorphic functions. Choosing the representation diagonal in $`r_i`$ and realizing $`p_j`$ as differential operators, $`p_j=i/r_j`$, one can work on the space of functions of the form $`\mathrm{\Psi }^j(𝐫,z)=_{k=j}^j\psi _k(𝐫)z^{j+k}`$. The quantum analog of the classical constraint takes the form of the quantum condition separating the physical subspace, $`(\mathrm{𝐬𝐧}+ϵj)\mathrm{\Psi }_{phys}^j(𝐫,z)=0.`$ It is necessary to note that quantum mechanically the above mentioned necessity of the regularization $`𝐬^2=\nu ^2+\epsilon ^2`$ is taken into account automatically. Indeed, since the eigenvalues of the operators $`𝐬^2`$ and $`(\mathrm{𝐬𝐧})^2`$ are separated in a necessary way, $`𝐬^2=j(j+1)`$, $`(\mathrm{𝐬𝐧})^2=j^2`$, one can say that the quantization “cures” the classical system. ## 5 Charge-monopole system as a symmetric top In this section we show that the charge-monopole system can also be interpreted as a reduced symmetric spinning top with dynamical tensor of inertia and “isospin” U(1) gauge symmetry. To this end, we return to the spin symplectic form (3.3), and supplement the unit vector $`𝐧𝐞_3`$ with the two vectors $`𝐞_1`$ and $`𝐞_2`$ forming together the oriented orthonormal set of vectors, $$𝐞_a𝐞_b=\delta _{ab},𝐞_1\times 𝐞_2=𝐞_3.$$ (5.1) As a consequence of basic relations (5.1), the vectors $`𝐞_a`$, $`a=1,2,3,`$ satisfy also the completeness relation $$e_a^ie_a^j=\delta ^{ij}.$$ (5.2) Using Eqs. (5.1), (5.2), one can represent the two-form (3.3) as $$d\theta =\nu d𝐞_1d𝐞_2,$$ (5.3) from which we get another representation for the one-form, $$\theta =\frac{\nu }{2}(𝐞_1d𝐞_2𝐞_2d𝐞_1).$$ (5.4) Taking into account the relation $`\dot{𝐞}_3^2=(\dot{𝐞}_1𝐞_3)^2+(\dot{𝐞}_2𝐞_3)^2`$, we get the alternative Lagrangian for the charge-monopole system, $`L`$ $`=`$ $`{\displaystyle \frac{1}{2}}\dot{r}{}_{}{}^{2}+{\displaystyle \frac{1}{2}}r^2((\dot{𝐞}_1𝐞_1\times 𝐞_2)^2+(\dot{𝐞}_2𝐞_1\times 𝐞_2)^2)+{\displaystyle \frac{\nu }{2}}(𝐞_1\dot{𝐞}_2𝐞_2\dot{𝐞}_1)`$ (5.5) $`{\displaystyle \frac{\lambda _1}{2}}(𝐞_1^21){\displaystyle \frac{\lambda _2}{2}}(𝐞_2^21)\lambda _{12}𝐞_1𝐞_2,`$ with $`\lambda _1`$, $`\lambda _2`$ and $`\lambda _{12}`$ being Lagrange multipliers, variation over which gives the Lagrangian constraints $`𝐞_1^21=0`$, $`𝐞_2^21=0`$ and $`𝐞_1𝐞_2=0`$. With these constraints, the equations of motion for $`𝐫=r𝐞_3`$, $`𝐞_3=𝐞_1\times 𝐞_2`$, following from (5.5) coincide with the charge-monopole Lagrange equations (2.3). To prove the complete equivalence of the model (5.5) to the initial system (2.1), we pass over to the Hamiltonian formalism. The canonical Hamiltonian corresponding to Lagrangian (5.5) is $$H_c=\frac{1}{2}p_r^2+\frac{1}{2r^2}(𝐉\times 𝐞_3)^2+\frac{\lambda _1}{2}(𝐞_1^21)+\frac{\lambda _2}{2}(𝐞_2^21)+\lambda _{12}𝐞_1𝐞_2,$$ (5.6) where $`𝐞_3=𝐞_1\times 𝐞_2`$ and the total angular momentum is $`𝐉=𝐞_1\times 𝐩_1+𝐞_2\times 𝐩_2`$ with $`𝐩_{1,2}`$ being the momenta canonically conjugate to $`𝐞_{1,2}`$. Note that in terms of the velocity phase space variables the total angular momentum is reduced to $`𝐉=r^2𝐞_3\times \dot{𝐞}_3\nu 𝐞_3`$ in correspondence with Eqs. (2.4). The application of Dirac-Bergmann theory to the system (5.5) results in the following complete set of constraints: $`\pi _10,\pi _20,\pi _{12}0,`$ (5.7) $`𝐞_1^210,𝐩_1𝐞_10,𝐞_2^210,𝐩_2𝐞_20,𝐞_1𝐞_20,𝐩_1𝐞_2+𝐩_2𝐞_10,`$ (5.8) $`𝐩_1𝐞_2𝐩_2𝐞_1+\nu 0,`$ (5.9) with the momenta $`\pi _1`$, $`\pi _2`$ and $`\pi _{12}`$ canonically conjugate to Lagrange multipliers. Constraints (5.7) mean that the Lagrange multipliers are pure gauge variables and can be completely excluded, e.g., by supplementing (5.7) with the gauge conditions $`\lambda _10`$, $`\lambda _20`$, $`\lambda _{12}0`$. The constraints (5.8) form the subset of second class constraints, whereas (5.9) is the first class constraint. Reduction of the symplectic two-form of the system, $`\omega =d𝐩_1d𝐞_1+d𝐩_2d𝐞_2+dp_rdr`$, to the surface of second class constraints (5.8) is given by $$\omega =\frac{1}{2}d(𝐉\times 𝐞_a)d𝐞_a+dp_rdr,$$ (5.10) where the summation over $`a=1,2,3`$ is assumed. The two-form (5.10) is the symplectic form on the reduced phase space which is described by the basis of vectors $`𝐞_a`$, $`a=1,2,3,`$ subject to conditions (5.1) as strong relations, by the angular momentum vector $`𝐉`$ and by the canonically conjugate radial variables $`r`$ and $`p_r`$. From (5.10) we obtain the following Poisson-Dirac brackets on the reduced phase space: $$\{e_a^i,e_b^j\}=0,\{J^i,e_a^j\}=ϵ^{ijk}e_a^k,\{J^i,J^j\}=ϵ^{ijk}J^k,\{r,p_r\}=1,$$ (5.11) and all other brackets for radial variables $`r`$ and $`p_r`$ to be equal to zero. The remaining constraint (5.9) takes the form $$\chi =\mathrm{𝐉𝐞}_3+\nu 0,$$ (5.12) and the total Hamiltonian of the system is reduced to the canonical Hamiltonian extended by the first class constraint multiplied by an arbitrary function $`\lambda (t)`$: $$H=\frac{1}{2}p_r^2+\frac{1}{2r^2}(𝐉\times 𝐞_3)^2+\lambda (\mathrm{𝐉𝐞}_3+\nu ).$$ (5.13) To understand the physical sense of the obtained system, let us define the scalar quantities $`I_a\mathrm{𝐉𝐞}_a`$, $`I_aI_a=𝐉^2`$. They satisfy the following Poisson bracket relations: $$\{I_a,I_b\}=ϵ_{abc}I_c,\{I_a,e_b^i\}=ϵ_{abc}e_c^i,\{I_a,J^i\}=0.$$ (5.14) Since $`I_a`$ commute with $`J_i`$ and satisfy $`su(2)`$ algebra, one can treat them as components of the isospin vector. On the other hand, due to the basic relations (5.1), the quantities $`e_a^i`$ can be treated as the elements of the group $`SO(3)`$. Then the quantities $`J^i`$ have a sense of the basis of the left-invariant vector fields on this group whereas the quantities $`I_a`$ can be identified with the basis of the right-invariant vector fields . Due to the relations (5.14), the first class constraint (5.12) generates gauge transformations of the system which have a sense of SO(2)$``$U(1) isospin rotations generated by $`I_3`$. Under such transformations the vectors $`𝐉`$ and $`𝐞_3`$ are invariant. The electromagnetic U(1) gauge invariance of the curvature form (3.2) corresponding to the CS charge-monopole coupling term is changed here for the described SO(2)$``$U(1) gauge invariance of the two-form (5.3) corresponding to the CS form (5.4). Since $`(𝐉\times 𝐞_\mathrm{𝟑})^2=I_1^2+I_2^2`$, the system given by the Hamiltonian (5.13) and brackets (5.11) can be identified as a symmetric spinning top with the symmetry axis given by $`𝐞_3`$ and dynamical moment of inertia $`_1=_2=r^2`$. The isospin U(1) gauge symmetry generated by the constraint (5.12) means that the rotation about the symmetry axis $`𝐞_3`$ is of pure gauge nature and, so, is not observable. The formal difference of the symmetric spinning top system from the reduced E(3) system discussed in section 2.4 is that here the condition $`\mathrm{𝐉𝐧}=\mathrm{𝐉𝐞}_3=\nu `$ appears in the form of first class constraint (weak equality) unlike the strong equality in the case of the reduced E(3) system. The physical content of both systems is exactly the same. Indeed, besides the radial variables $`r`$ and $`p_r`$, the only observable quantities (commuting in the sense of Dirac-Poisson brackets with the constraint (5.12)) are the vector of angular momentum $`𝐉`$ and the unit vector $`𝐞_3𝐧`$. So, the E(3) system can also be treated as a symmetric spinning top reduced by the action of the first class constraint (5.12). The system is quantized as follows. In correspondence with the classical relations (5.1), (5.2) and the sense of the complete orthonormal set of vectors $`𝐞_a`$, they can be parametrized by the three Euler angles $`\alpha ,\beta ,\gamma `$, and we can choose a representation diagonal in these angle variables and in the radial variable $`r`$. Then the components of the operator $`𝐉`$ are realized in the form of linear differential operators of angular variables . An arbitrary state can be decomposed over the complete basis of Wigner functions, $$\mathrm{\Psi }(r,\alpha ,\beta ,\gamma )=\underset{j,s,k}{}\psi _{j,s,k}(r)D_{s,k}^j(\alpha ,\beta ,\gamma ),$$ where $`j`$ takes either integer, $`j=0,1,2,\mathrm{}`$, or half-integer, $`j=1/2,3/2,\mathrm{}`$, values, and $`s,k=j,j+1,\mathrm{},+j`$, $`𝐉^2D_{s,k}^j=j(j+1)`$, $`J_3D_{s,k}^j=sD_{s,k}^j`$, $`I_3D_{s,k}^j=kD_{s,k}^j`$. The quantum analog of the first class constraint (5.12) is transformed into the equation separating the physical states: $$(I_3\nu )\mathrm{\Psi }_{phys}=0.$$ (5.15) This equation has nontrivial solutions only for $`\nu =n/2`$, $`n\text{ZZ}`$, i.e. we arrive once again at the Dirac quantization condition, and the corresponding solutions of Eq. (5.15) have the form $$\mathrm{\Psi }_{phys}(r,\alpha ,\beta ,\gamma )=\underset{j,s}{}{}_{}{}^{}\psi _{s,\frac{n}{2}}^{j}(r)D_{s,\frac{n}{2}}^j(\alpha ,\beta ,\gamma ),$$ (5.16) where prime means that in the sum $`j`$ takes the values $`j=|n/2|,|n/2|+1,\mathrm{}`$. Once again, let us note that here the quantization separates in the necessary way the value of the quantized parameter $`|\nu |=|\frac{n}{2}|`$ and possible eigenvalues of the angular momentum operator in the physical subspace: $`𝐉^2=j(j+1)>\nu ^2`$. It is interesting to note that in the present “spinning top” picture for the charge-monopole system the total angular momentum is represented in terms of the isospin angular momentum as $`𝐉=𝐞_aI_a`$, whereas from the point of view of the field theoretical mechanism , only spin part of the total angular momentum vector of the charge-monopole system is created from the isospin degrees of freedom. The seeming contradiction is explained by the constraint (5.15) prescribing the total angular momentum to take the values starting from the minimal value $`|n/2|=|\nu |`$, which can be interpreted as the “internal” spin of the charge-monopole system, and in this sense here spin is also created by isospin. ## 6 Higher-derivative form of CS term The charge-monopole system can also be described by the Lagrangian with CS term represented in a higher-derivative form. To show this, we write down the CS form (5.4) as $`\theta =\nu 𝐞_2d𝐞_1`$. Identifying the vectors $`𝐞_3`$ and $`𝐞_1`$ with the unit vectors $`𝐧=𝐫r^1`$ and $`\dot{𝐧}|\dot{𝐧}|^1`$, respectively, and taking into account that $`𝐞_2=𝐞_3\times 𝐞_1`$, the charge-monopole coupling term can be written down in the equivalent higher-derivative form $$L_{int}=\frac{\nu }{\dot{𝐧}^2}𝐧(\dot{𝐧}\times \ddot{𝐧}).$$ (6.1) Simple geometrical consideration shows that the higher derivative term $`(𝐧\times \dot{𝐧})\ddot{𝐧}/\dot{𝐧}^2`$ has a sense of angular velocity of rotation of the vector $`\dot{𝐧}`$ about the vector $`𝐧`$. The total charge-monopole Lagrangian is given by $$L=\frac{1}{2}\dot{𝐫}{}_{}{}^{2}\nu \frac{r}{(𝐫\times \dot{𝐫})^2}(𝐫\times \dot{𝐫})\ddot{𝐫}.$$ (6.2) Naively, the equations of motion following from Lagrangian (6.2), $$\frac{L}{𝐫}\frac{d}{dt}\frac{L}{\dot{𝐫}}+\frac{d^2}{dt^2}\frac{L}{\ddot{𝐫}}=0,$$ (6.3) are the third order differential equations and their equivalence to the second-order equations (2.3) is not obvious. Let us prove the equivalence of equations of motion (6.3) to Eq. (2.3). First we note that the higher-derivative term (6.1) satisfies the following relations: $`𝐫L_{int}/𝐫=0,\dot{𝐫}L_{int}/\dot{𝐫}=L_{int},\ddot{𝐫}L_{int}/\ddot{𝐫}=+L_{int},`$ $`\dot{𝐫}L_{int}/𝐫=(𝐫\dot{𝐫})r^2L_{int},𝐫L_{int}/\dot{𝐫}=𝐫L_{int}/\ddot{𝐫}=\dot{𝐫}L_{int}/\ddot{𝐫}=0.`$ (6.4) Multiplying the equations (6.3) subsequently by the vectors $`𝐫`$, $`\dot{𝐫}`$ and $`𝐫\times \dot{𝐫}`$, and using Eqs. (6.4), we arrive at the three equations $`\ddot{𝐫}𝐫=0,`$ $`\ddot{𝐫}\dot{𝐫}=0,`$ $`\ddot{𝐫}(𝐫\times \dot{𝐫})=\frac{\nu }{r^3}(𝐫\times \dot{𝐫})^2.`$ Since $`𝐫`$, $`\dot{𝐫}𝐫(𝐫\dot{𝐫})r^2`$ and $`𝐫\times \dot{𝐫}`$ form the complete basis of orthogonal vectors, we conclude that the equations of motion (6.3) are equivalent to the obtained system of three scalar equations, which, in turn, is equivalent to one vector equation (2.3). Therefore, the coupling of the charge to the magnetic monopole can be alternatively described by the scale invariant higher derivative term (6.1). It is worth noting that analogously to the present system, the relativistic massive particle in (2+1) dimensions coupled to the external constant homogeneous electromagnetic field turns out to be classically equivalent to the higher derivative model of relativistic particle with torsion given by the action $`S_{tor}=(m+\alpha \kappa )𝑑s,`$ where $`ds^2=dx_\mu dx^\mu `$, $`\kappa `$ is a scale invariant torsion of the particle’s world trajectory, $`\kappa =ϵ^{\mu \nu \lambda }x_\mu ^{}x_\nu ^{\prime \prime }x_\lambda ^{\prime \prime \prime }/(x^{\prime \prime })^2`$, $`x_\mu ^{}=dx_\mu /ds`$, and $`\alpha `$ is a parameter. Such a model underlies (2+1)-dimensional anyons , and in the last section we shall discuss the relationship between the charge-monopole system and anyons. ## 7 Twistor description of the charge-monopole system The helicity constraint appearing in the charge-monopole system interpreted as a particle with spin is analogous to the helicity-fixing constraint in the case of massless particle with spin. The latter system admits, in particular, the twistor description -. Using this observation, one can get the twistor formulation for the charge-monopole system. In the twistor approach for the massless particle, the corresponding energy-momentum vector is treated as a “composite” object constructed from the twistor (even spinor) variables, and the helicity fixing constraint generates the U(1) transformations for twistors. By analogy, let us introduce mutually conjugate even complex variables $`z_a`$ and $`\overline{z}_a=z_a^{}`$, $`a=1,2`$, forming two conjugate spinors. With them, we represent the charge coordinate vector $`𝐫`$ as a composite vector, $$\phi _ir_iz\sigma _i\overline{z}0,$$ (7.1) where $`\sigma _i`$ is the set of Pauli matrices. Introducing the momenta $`𝒫_a`$, $`\overline{𝒫}_a`$ canonically conjugate to the spinor variables, $`\{z_a,𝒫_b\}=\delta _{ab}`$, $`\{\overline{z}_a,\overline{𝒫}_b\}=\delta _{ab}`$, we construct the generators of rotations (the total angular momentum vector), $$J_i=(𝐫\times 𝐩)_i+\frac{i}{2}\left(z\sigma _i𝒫\overline{𝒫}\sigma _i\overline{z}\right),$$ (7.2) forming the su(2) algebra: $`\{J_i,J_j\}=ϵ_{ijk}J_k`$. Here $`𝐩`$ is the vector canonically conjugate to $`𝐫`$. Following the twistor approach, we also introduce the constraint $$\chi \frac{i}{2}(\overline{z}\overline{𝒫}z𝒫)\nu 0,$$ (7.3) which is in involution with the constraints (7.1) and generates the U(1) gauge transformations for the spinor variables, $`z_az_a^{}=e^{i\gamma }z_a`$, $`\overline{z}_a\overline{z}{}_{a}{}^{}=e^{i\gamma }\overline{z}_a`$, $`𝒫_a𝒫_a^{}=e^{i\gamma }𝒫_a`$, $`\overline{𝒫}_a\overline{𝒫}{}_{a}{}^{}=e^{i\gamma }\overline{𝒫}_a`$, with $`\gamma =\gamma (t)`$ being a parameter of transformation. Taking into account the identity $$\sigma _{ab}^i\sigma _{cd}^i=\delta _{ab}\delta _{cd}2ϵ_{ac}ϵ_{bd},$$ (7.4) with $`ϵ_{ab}=ϵ_{ba}`$, $`ϵ_{12}=1`$, we get the relation $`r\overline{z}z`$ as the consequence of the constraint (7.1). This relation and Eq. (7.2) allow us to represent the constraint (7.3) in the equivalent form of the helicity-fixing constraint, $$\stackrel{~}{\chi }\mathrm{𝐉𝐫}+\nu r0.$$ (7.5) Since the total phase space described by $`r_i`$, $`p_i`$, $`z_a`$, $`\overline{z}_a`$, $`𝒫_a`$ and $`\overline{𝒫}_a`$ is $`14`$-dimensional, and we have 4 first class constraints (7.1) and (7.3), there are only 6 independent physical phase space degrees of freedom like in the charge-monopole system. They are given by the observables having zero Poisson brackets with all the constraints. Such independent variables are $`r_i`$ and $$\mathrm{\Pi }_ip_i+\frac{1}{2\overline{z}z}(z\sigma _i𝒫+\overline{𝒫}\sigma _i\overline{z}),$$ (7.6) for which we have the Poisson bracket relations $`\{r_i,r_j\}=0`$, $`\{r_i,\mathrm{\Pi }_j\}=\delta _{ij}`$ and $$\{\mathrm{\Pi }_i,\mathrm{\Pi }_j\}=(\overline{z}z)^2ϵ_{ijk}(𝐫\times 𝚷𝐉)_k\frac{\nu }{r^3}ϵ_{ijk}r_k.$$ (7.7) The weak equality means the equality on the surface of constraints (7.1) and (7.3). Physical variables $`\mathrm{\Pi }_i`$ correspond here to the charge-monopole variables $`P_i`$. The Hamiltonian of the charge-monopole system in the twistor formulation can be taken as the linear combination of $`𝚷^2`$ and of the first class constraints, $$H=\frac{1}{2}𝚷^2+\rho _i\phi _i+\lambda \chi ,$$ (7.8) with $`\rho _i=\rho _i(t)`$, $`\lambda =\lambda (t)`$ being arbitrary functions (Lagrange multipliers). Direct calculation shows that the equations of motion generated by this Hamiltonian are reduced to $`\ddot{𝐫}\nu r^3𝐫\times \dot{𝐫}`$, i.e. the Hamiltonian (7.8) and constraints (7.1), (7.3) give the alternative twistor description for the charge-monopole system. The quantum theory of the system in the twistor approach is the following. It is natural to choose the represenation diagonal in $`𝐫`$, $`z_a`$ and $`\overline{z}_a`$, and realize the canonically conjugate momenta in the form of differential operators. In accordance with constraint (7.1), the physical states have to be of the form $`\mathrm{\Psi }_{phys}=\delta ^{(3)}(𝒓z𝝈\overline{z})\psi (z,\overline{z})`$. Then like in the model of massless particle with spin , the quantum analog of the constraint (7.3), $$\left(\frac{1}{2}\left(\overline{z}_a\frac{}{\overline{z}_a}z_a\frac{}{z_a}\right)\nu \right)\mathrm{\Psi }_{phys}=0,$$ (7.9) and the requirement of single-valuedness of the wave functions results in the quantization of the charge-monopole constant, $`2\nu =n`$, $`n\text{ZZ}`$. Finally, the physical states subject to Eq. (7.9) will be described by the wave functions of the form $$\mathrm{\Psi }_{phys}=\delta ^{(3)}(𝒓z𝝈\overline{z})\psi _{phys}(z,\overline{z}),\psi _{phys}(z,\overline{z})=\underset{k=\mathrm{}}{\overset{\mathrm{}}{}}\underset{a,b=1,2}{}C_k^{ab}(z_a)^k(\overline{z}_b)^{n+k},$$ (7.10) where $`C_k^{ab}`$ are constants. The action of quantum analogs of classical observables $`𝐫`$ and $`𝚷`$ on physical wave functions (7.10) is reduced to $`𝐫\mathrm{\Psi }_{phys}`$ $`=`$ $`\delta ^{(3)}(𝒓z𝝈\overline{z})z𝝈\overline{z}\psi _{phys}(z,\overline{z}),`$ $`𝚷\mathrm{\Psi }_{phys}`$ $`=`$ $`i\delta ^{(3)}(𝒓z𝝈\overline{z})(2\overline{z}z)^1(z𝝈{\displaystyle \frac{}{z}}+{\displaystyle \frac{}{\overline{z}}}𝝈\overline{z})\psi _{phys}(z,\overline{z}).`$ (7.11) By inverse Legendre transformation, analogously to the case of the massless particle with spin , one can construct the Lagrangian corresponding to Hamiltonian (7.8). Instead of realizing such constructions, we note here that the list of independent observable quantities and the action of their quantum analogs on the physical states indicate on the possibility to exclude $`𝐫`$ and $`𝐩`$ as independent variables and to construct the theory realized only in terms of spinor variables. To get the corresponding twistor form of the Lagrangian, it is convenient to start from the higher-derivative Lagrangian (6.2) by putting in it $`𝒓=z𝝈\overline{z}.`$ The remarkable feature of such a twistor representation is that its substitution into the CS term of the higher derivative form (6.1) reduces the latter to $`i\nu (\dot{\overline{z}}z\overline{z}\dot{z})/\overline{z}z`$, and we arrive at the Lagrangian $$L=\frac{1}{2}\left(\frac{d}{dt}(z𝝈\overline{z})\right)^2+i\nu \frac{\overline{z}\dot{z}\dot{\overline{z}}z}{\overline{z}z}.$$ (7.12) Let us show that (7.12) describes the charge-monopole system. From the definition of the canonical momenta $`𝒫_a=L/\dot{z}_a`$, $`\overline{𝒫}_a=L/\dot{\overline{z}}_a`$, we find that the constraint (7.3) is the primary constraint for the system (7.12) and the total Hamiltonian is $$H=\frac{1}{2z\overline{z}}\pi \overline{\pi }+\lambda \chi ,$$ (7.13) where we have introduced the notation $`\pi _a=𝒫_ai\nu \overline{z}_a(\overline{z}z)^1,`$ $`\overline{\pi }_a=\overline{𝒫}_a+i\nu z_a(\overline{z}z)^1.`$ The constraint (7.3) is conserved by the Hamiltonian and, as a consequence, there are no secondary constraints. Therefore, the constraint (7.3) is the first class constraint and the number of physical phase space degrees of freedom is equal to $`82=6`$. The corresponding 6 independent observables weakly commuting in the Poisson bracket sense with the constraint (7.3) are $$r_i=z\sigma _i\overline{z},\mathrm{\Pi }_i=\frac{1}{2z\overline{z}}(\overline{𝒫}\sigma _i\overline{z}+z\sigma _i𝒫).$$ (7.14) These quantities satisfy the following Poisson bracket relations: $`\{r_i,r_j\}=0`$, $`\{r_i,\mathrm{\Pi }_j\}=\delta _{ij}`$, $`\{\mathrm{\Pi }_i,\mathrm{\Pi }_j\}\nu r^3ϵ_{ijk}r_k`$, where the weak equality means the equality on the surface of the constraint (7.3). Obviously, they have the sense of the charge-monopole variables $`r_i`$ and $`P_i`$ represented in the composite form (7.14) in terms of twistor variables. The variables $`\mathrm{\Pi }_i`$ are nothing else as the classical analogs of the quantum operators corresponding to observables (7.6), whose action is reduced to the surface of the constraints (7.1), i.e. they are the classical analogs of operators (7.11). The conserving $`su(2)`$ generators are represented here in the form $`𝐉=𝐫\times 𝚷\nu 𝐧`$ with $`𝐧=𝐫r^1`$ and $`r_i`$, $`\mathrm{\Pi }_i`$ given by Eq. (7.14). Direct verification shows that the Hamiltonian (7.13) generates the correct equations of motion, $`\ddot{𝐫}=\nu r^3𝐫\times \dot{𝐫}`$. It is worth noting that the system (7.13) can be treated as a reduction of the Hamiltonian system (7.8) to the surface of the constraints (7.1) supplied with the gauge conditions $`p_i\frac{1}{2}(z\overline{z})^1(z\sigma _i𝒫+\overline{𝒫}\sigma _i\overline{z})0`$. To conclude the discussion of the twistor formulation for the charge-monopole system, let us show how the limit of zero mass described here by the Lagrangian $$L=i\nu \frac{\overline{z}\dot{z}\dot{\overline{z}}z}{\overline{z}z}$$ (7.15) gives rise to the pure spin system. Here, as before, it is assumed that the initial configuration space is $`\text{lC}^2\{0\}`$. This system can be quantized by the method advocated in ref. , but the symplectic two-form corresponding to Lagrangian (7.15) is singular, $`\omega =\omega _{ab}d\overline{z}_adz_b`$, $`\omega _{ab}=2i\nu (\overline{z}z)^1(\delta _{ab}z_a\overline{z}_b(\overline{z}z)^1)`$, $`\omega _{ab}\overline{z}_a=\omega _{ab}z_b=0`$, that complicates the analysis. There is a more short way to reveal a spin nature of the system (7.15) by showing its equivalence to the system (3.10). For the purpose we note that in the regions where either $`z_10`$ (chart $`U_1`$) or $`z_20`$ (chart $`U_2`$), one can define the complex coordinate $`Z`$ as $`Z=z_2/z_1`$ or $`Z=z_1/z_2`$, respectively. Then Lagrangian (7.15) is represented equivalently as $`L=i\nu (\overline{Z}\dot{Z}\dot{\overline{Z}}Z)(1+\overline{Z}Z)^12\nu \dot{\phi },`$ where $`\phi `$ is the phase of the coordinate $`z_a`$ in the chart $`U_a`$, $`a=1,2`$. The first term coincides exactly with the Lagrangian (3.10) corresponding to the pure spin system. The second total derivative term is not important classically as well as quantum mechanically since its contribution to the action, $`\mathrm{\Delta }S=4\nu \pi `$, for periodical trajectories is trivialized, $`\mathrm{\Delta }S=0(mod\mathrm{\hspace{0.17em}2}\pi )`$, if we take into account the quantization condition $`2\nu \text{ZZ}`$. Therefore, the system (7.15) is equivalent to the pure spin system (3.10). The equivalence of the system (7.15) to the system (3.10) is encoded in its gauge symmetries. In correspondence with the above mentioned degeneracy of the symplectic form, the system (7.15) possesses two gauge invariances: its action is invariant under the transformations $`z_a\rho z_a`$, $`\rho =\rho (t)>0`$, generated by the constraint $`\rho =\overline{𝒫}\overline{z}+z𝒫0`$, and $`z_ae^{i\phi }z_a`$, $`\phi =\phi (t)\text{IR }`$, generated by the helicity constraint of the form (7.3). This means that the points $`z_a`$ and $`\zeta z_a`$, $`\zeta \text{lC},`$ $`\zeta 0`$, in configuration space are physically equivalent and should be identified, $`z_a\zeta z_a`$. As a result, the configuration space of the system is the projective complex plane $`\text{lC}P^1=(\text{lC}^2\{0\})/`$, on which the coordinate $`Z`$ introduced above plays a role of the “inhomogeneous” coordinate . ## 8 Charge-monopole in a spherical geometry and spin Let us restrict the motion of the charge in the monopole field to the sphere $`𝐫^2=1`$. This can be done by treating the relation $`𝐫^21=0`$ as a Lagrangian constraint, $`L=\frac{1}{2}\dot{𝐫}{}_{}{}^{2}+e\dot{𝐫}𝐀(𝐫)\frac{\lambda }{2}(𝐫^21),`$ whose condition of conservation generates the Hamiltonian constraint $`p_r=0`$. The reduction to the surface of these second class constraints excludes the radial variables $`r`$ and $`p_r`$, and we arrive at the Hamiltonian describing the charge on the sphere with the monopole in its center<sup>2</sup><sup>2</sup>2See also refs. , where some aspects of the charge-monopole system in a spherical geometry were discussed., $`H=\frac{1}{2}(𝐉\times 𝐧)^2.`$ Here the dynamical variables $`𝐉`$ and $`𝐧`$ satisfy the Poisson bracket relations (2.22) and are subject to the conditions (2.23), i.e. the reduced system is a pure reduced E(3) system. The same procedure of reduction can be realized in a spinning top picture, i.e. by adding the condition $`𝐫^21=0`$ to Lagrangian (5.5) as a Lagrangian constraint, or by putting $`r=1`$ directly in Lagrangian (5.5): $$L=\frac{1}{2}\dot{𝐧}{}_{}{}^{2}+\frac{\nu }{2}(𝐞_1\dot{𝐞}_2𝐞_2\dot{𝐞}_1),$$ (8.1) where we suppose that $`𝐧=𝐞_1\times 𝐞_2`$ and vectors $`𝐞_1`$ and $`𝐞_2`$ are orthonormal. In the case of higher derivative treatment of the charge-monopole system the corresponding reduced Lagrangian is $$L=\frac{1}{2}\dot{𝐧}{}_{}{}^{2}\frac{\nu }{\dot{𝐧}^2}𝐧(\dot{𝐧}\times \ddot{𝐧}),$$ (8.2) whereas in the twistor picture the Lagrangian takes the form $$L=2(\dot{z}\sigma _2z)(\dot{\overline{z}}\sigma _2\overline{z})+i\nu (\overline{z}\dot{z}\dot{\overline{z}}z),\overline{z}z=1.$$ (8.3) The condition $`\overline{z}z=1`$ can be omitted by normalizing appropriately the Lagrangian, i.e. by multiplying the first and second terms by $`(\overline{z}z)^2`$ and $`(\overline{z}z)^1`$, respectively. The CS term is not changed by the reduction procedure due to its scale-invariance. But this term in addition is reparametrization invariant, whereas the total Lagrangian has no reparametrization invariance. We can change the non-invariant second order in velocity term $`\frac{1}{2}\dot{𝐧}^2`$ for the first order term $`\sqrt{\dot{𝐧}^2}`$, that gives rise to the reparametrization invariant action. Let us analyze the physical content of such reparametrization action considering, e.g., the modification of the Lagrangian (8.1), $$L=\gamma \sqrt{\dot{𝐧}^2}+\frac{\nu }{2}(𝐞_1\dot{𝐞}_2𝐞_2\dot{𝐞}_1),$$ (8.4) where $`\gamma >0`$ is a dimensionless parameter. The canonical Hamiltonian of the system (8.4) is equal to zero, and from the Hamiltonian point of view, the difference of the system (8.4) from the system (8.1) consists in the presence of the constraint $`𝐉^2\kappa ^20,`$ $`\kappa ^2\gamma ^2+\nu ^2,`$ generating reparametrizations in addition to the isospin U(1) gauge symmetry generated by the constraint (5.12). The system (8.4) has the same physical content as the spin system of section 3 given by only the CS term. Indeed, applying the quantization scheme of section 5, the physical subspace is separated here by Eq. (5.15) and by $`(𝐉^2\kappa ^2)\mathrm{\Psi }_{phys}(\alpha ,\beta ,\gamma )=0.`$ These two equations have nontrivial solutions only when the parameters $`\kappa `$ and $`\nu `$ are quantized: $`\kappa ^2=j(j+1)`$ and $`\nu =k`$, where $`k`$, $`|k|j`$, is integer (half-integer) for $`j`$ integer (half-integer), and the corresponding physical state is $`\mathrm{\Psi }_{phys}(\alpha ,\beta ,\gamma )=_{s=j}^jC_sD_{s,k}^j(\alpha ,\beta ,\gamma ),`$ where $`C_s`$ are constants. This wave function describes an arbitrary state of fixed spin $`j`$ in the form alternative to the holomorphic functions of section 3. Therefore, the reparametrization-invariant system (8.4) is equivalent to the spin system described by only the CS term. The charge-monopole system in spherical geometry can be treated as a partially gauge fixed version of the reparametrization and scale invariant spin system (8.4). To see this, we rewrite the Lagrangian (8.4) in equivalent form by introducing the einbein $`v`$: $$L=\frac{\dot{𝐧}^2}{2v}+\frac{v}{2}\gamma ^2+\frac{\nu }{2}(𝐞_1\dot{𝐞}_2𝐞_2\dot{𝐞}_1).$$ (8.5) Reparametrization invariance of the system (8.5) can be fixed (locally, see ref. ) via introducing the appropriate gauge-fixing conditions for the constraint $`𝐉^2\kappa ^20,`$ and for the constraint $`p_v0`$, where $`p_v`$ is the momentum canonically conjugate to $`v`$. On the other hand, introducing only the condition $`v=1`$, we obtain the partially gauge fixed version of the system (8.5) . The Lagrangian (8.5) with $`v=1`$ is the Lagrangian (8.1) shifted for the inessential constant. Therefore, the charge-monopole system in a spherical geometry can formally be treated as a partially gauge fixed version of the spin system (8.5). Analogously, we can change the kinetic term $`\frac{1}{2}\dot{𝐫}^2`$ in the initial charge-monopole Lagrangian (2.1) for the reparametrization invariant term $`\sqrt{\dot{𝐫}^2}`$, the latter is a kinetic term of relativistic particle in 3D Euclidean space. As a result, we get the reparametrization invariant action $$S=L_r𝑑t,L_r=\frac{\dot{𝐫}^2}{2v}+\frac{v}{2}+e𝐀\dot{𝐫}.$$ (8.6) Its partial gauge fixed version ($`v=1`$) will give the Lagrangian coinciding up to inessential constant with the initial Lagrangian (2.1). From this point of view, the time translation, the time dilation and special conformal transformation symmetries produced canonically by the $`so(2,1)`$ generators $`H`$, $`D`$ and $``$ , can be treated as a relic of the reparametrization symmetry of the system (8.6) surviving the described formal Lagrangian gauge fixing procedure. ## 9 Charge-monopole system and anyons Let us discuss the relationship between the charge-monopole system and (2+1)-dimensional anyons in the light of the obtained results. Earlier, the analogy with the charge-monopole system played an important role in constructing the theory of anyons as spinning particles -. We have observed that the charge-monopole system in many aspects is similar to the 3D free particle. In (2+1)-dimensions, spin is a (pseudo)scalar and, as a consequence, the anyon of fixed spin has the same number of degrees of freedom as a spinless free massive particle. The relationship between the charge-monopole and anyon can be understood better within the framework of the canonical description of these two systems. As we have seen, the charge-monopole system essentially is a reduced E(3) system. In the case of anyons, the E(3) group is changed for the Poincaré group ISO(2,1). The translation generators of the corresponding groups are $`𝐧`$ and $`p_\mu `$, the latter being the energy-momentum vector of the anyon, and the corresponding Casimir central elements are fixed by the relations $`𝐧^2=1`$ and $$p^2+m^20,$$ (9.1) where $`m`$ is a mass of the anyon. The rotation (Lorentz) generators are given by $`𝐉`$ and by $$𝒥_\mu =ϵ_{\mu \nu \lambda }x^\nu p^\lambda +J_\mu ,$$ (9.2) where $`J_\mu `$ are the translation invariant $`so(2,1)`$ generators satisfying the algebra of the form (2.17), $`\{J_\mu ,J_\nu \}=ϵ_{\mu \nu \lambda }J^\lambda `$, and subject to the relation $`J_\mu J^\mu =\alpha ^2=const`$. The representation (9.2) for the anyon total angular momentum vector is, obviously, the analog of the relation $`𝐉=𝐋+𝐬`$ appearing under interpretation of the charge-monopole system as a particle with spin. In the charge-monopole system, the second Casimir element of E(3) is fixed either strongly, $`\mathrm{𝐉𝐧}=\nu `$, or in the form of the weak relation $`\mathrm{𝐉𝐧}+\nu 0`$ (see Eqs. (4.2), (4.6)). In the anyon model, spin also can be fixed either strongly, $`𝒥p=Jp=\alpha m`$, or in the form of the weak (constraint) relation $$\chi _aJp+\alpha m0.$$ (9.3) When the helicity is fixed strongly, for the charge-monopole system the symplectic form corresponding to the Poisson brackets (2.8), has a nontrivial contribution describing noncommuting quantities $`P_i`$ being the components of the charge’s velocity: $`\omega =dP_idr_i+\frac{\nu }{2r^3}ϵ_{ijk}r_idr_jdr_k`$. In the anyon case, the strong spin fixing gives rise to the nontrivial Poisson structure for the particle’s coordinate’s -, $$\{x_\mu ,x_\nu \}=\alpha (p^2)^{3/2}ϵ_{\mu \nu \lambda }p^\lambda .$$ (9.4) On the other hand, when we treat the charge-monopole system as a particle with spin (helicity is fixed weakly), one can work in terms of the canonical symplectic structure for the charge’s coordinates and momenta (see Eq. (4.1)), but the canonical momenta $`𝐩`$ are not observables due to their non-commutativity with the helicity constraint, whereas the gauge-invariant extension of $`𝐩`$ given by Eq. (4.3) plays the role of the non-commuting quantities $`P_i`$. Exactly the same picture takes place in the case of anyon when its spin is fixed weakly: the coordinates $`x_\mu `$ commute in this case, $`\{x_\mu ,x_\nu \}=0`$, but they have nontrivial Poisson brackets with the spin constraint (9.3), whereas their gauge-invariant extension, $`X_\mu =x_\mu +\frac{1}{p^2}ϵ_{\mu \nu \lambda }p^\nu J^\lambda `$ (cf. with Eq. (4.3)), $`\{X_\mu ,\chi _a\}0`$, are non-commuting and reproduce the Poisson bracket relation (9.4). Like in the anyon case, the advantage of the extended formulation for the charge-monopole system (when we treat it as a particle with spin), is in the existence of canonical charge’s coordinates $`r_i`$ and momenta $`p_i`$. Within the initial minimal formulation (given in terms of $`r_i`$ and gauge-invariant variables $`P_i`$), the canonical momenta $`p_i`$ are reconstructed from $`P_i`$ only locally, $`p_i=P_i+eA_i`$, due to the global Dirac string singularities hidden in the monopole vector potential. Having in mind the gauge invariant nature and non-commutativity of $`P_i`$ or their analogs $`\mathrm{\Pi }_i`$ from the extended formulation, we conclude that they, like anyon coordinates $`X_\mu `$, are the charge-monopole’s analogs of the Foldy-Wouthuysen coordinates of the Dirac particle . ## 10 Concluding remarks The discussed classical $`so(2,1)`$ symmetry can be quantized in an abstract way proceeding from the classical relations (2.17) and (2.19). In such a way the infinite-dimensional unitary half-bounded $`sl(2,R)`$ representations of the discrete series $`D_\alpha ^+`$ will be obtained, which are characterized by the quantum Casimir element $`𝒞=\alpha (\alpha 1)`$, $`0<\alpha \text{IR }`$ and by the eigenvalues $`j_0=\alpha +n`$, $`n=0,1,\mathrm{}`$, of the operator $`𝒥_0`$ . Since $`\alpha >0`$ is arbitrary, such a quantization procedure of the $`so(2,1)`$ symmetry algebra does not introduce any restrictions for the charge-monopole coupling constant, and does not fix correctly the spectrum of the operator $`𝐉^2`$. The latter information, as we saw, is encoded in the corresponding $`so(3)`$ algebra and classical condition $`𝐉^2>\nu ^2`$. The quantization of the parameter $`\nu `$ and the quantum spectrum of $`𝐉^2`$ could be obtained in principle by applying the geometric quantization to the classical E(3) system from section 2.4. On the other hand, the same information could be extracted from the quantization of classical $`so(3,1)`$ symmetry of the charge-monopole system described in section 2.3 with taking into account the classical relation $`𝐉^2>\nu ^2`$. We are going to consider the geometric quantization of $`so(3,1)`$ charge-monopole symmetry elsewhere. The observation of the similarity between the charge-monopole and the 3D free particle systems realized in section 2.3 in the context of the vector integrals of motion has been applied recently in ref. for explaining the nature of the nonstandard fermion-monopole supersymmetry . In sections 4 and 5 we have discussed the two different interpretations of the charge-monopole system as a particle with spin and as a spinning top system. It is interesting to find the corresponding map between these pictures at the Hamiltonian level. In the first picture, the “rotational” part of the phase space is given by the spin vector $`𝐬`$, $`𝐬^2=\nu ^2`$, and by the orbital angular momentum $`𝐋`$ and associated unitary vector $`𝐧`$. In the spinning top picture we have the angular momentum vector $`𝐉`$ and the set of orthonormal vectors $`𝐞_a`$. The vector $`𝐉`$ of the second formulation is identified with the total angular momentum vector $`𝐋+𝐬`$ from the first formulation, and the vector $`𝐞_3`$ giving the symmetry axis of the top is naturally identified with $`𝐧`$. Therefore, to find the mapping between the two formulations, it is sufficient to construct from the variables $`𝐬`$, $`𝐋`$ and $`𝐧`$ the orthonormal vectors $`𝐞_{1,2}`$, $`𝐞_1\times 𝐞_2=𝐧`$, satisfying the necessary Poisson bracket relations. Unfortunately, we did not succeed in realization of such a construction. It seems interesting to investigate analogously other systems of particles (strings) in the background of external gauge or gravitational fields from the point of view of their possible alternative description as the particles (strings) with internal reduced degrees of freedom. Acknowledgements I am grateful to J. Zanelli and D. Sorokin for interest to this work and helpful discussions. I thank J. Alfaro and J. Gamboa for bringing refs. to my attention and M. Bañados for discussion of some related issues. The work has been supported by the grant 1980619 from FONDECYT (Chile) and by DICYT (USACH).
warning/0004/quant-ph0004037.html
ar5iv
text
# “Violating” Clauser-Horne inequalities within classical mechanics ## 1. Introduction Many years have elapsed since Bell , nailed down his prolegomena to any future hidden-variables theory, and Horne and Clauser worked out their inequalities , that shifted the whole issue from the realm of the Gedankenexperimente to the optical laboratories. And from these workshops of empirical evidence the response came, more and more convincing with the lapse of the years and the consequent refinement of the techniques for experimentation, aimed at closing all the possible loopholes , , . Needless to say, the experiments without exceptions have confirmed, although not to everybody’s satisfaction, the predictions stemming from quantum mechanics. Had not this been the case, theoretical physicists would have found something very interesting to do during the last years. But alas, quantum mechanics works finely even with these pairs of photons, whatever a photon may turn out to be <sup>1</sup><sup>1</sup>1 In a letter of December 12, 1951 Albert Einstein wrote to his long-time friend Michele Besso: “The whole fifty years of conscious brooding have not brought me nearer to the answer to the question ‘What are light quanta?’. Nowadays every scalawag believes that he knows what they are, but he deceives himself.” This English translation of the excerpt can be found in an article by J. Stachel .. Under this respect, the above mentioned experiments did not have much to add to what was already common wisdom for the overwhelming majority of physicists since many decades: quantum mechanics is the theory of choice for describing the microscopic reality in its interaction with macroscopic detectors. But of course this is only one aspect of the question. The reason why the outcome of such delicate experiences is scrutinised by the theoreticians with so much attention depends also on their epistemological charme: not only is quantum mechanics once more confirmed by the experiments, but also (apart from residual loopholes) the so called local hidden-variables theories appear deprived of any residual hope for challenging quantum mechanics as the sole ruler of the microscopic realm . At first sight, one does not grasp wherefrom the pressure for settling so abstract a question could come, for no credible pretenders to the role presently kept by quantum theory have emerged up to now. For this reason the attempt at outlining a priori the features of such would be pretenders is fraught with the unavoidable dangers of vagueness. Despite these risks, much effort has been spent for endowing the phantom pretenders with formal attributes that could allow for more and more general arguments against them. Much less work has been done (with some notable exceptions, like the ones accounted for in , ) for understanding what the good old classical physics <sup>2</sup><sup>2</sup>2Die Physik der Modelle”, as Schrödinger nostalgically dubbed it . might have to suggest, through the analysis of particular instances, as hypothetical patterns of behaviour. ## 2. The Clauser-Horne inequalities With the optical experiment in mind, Clauser and Horne have envisaged the behaviour that a wide class of stochastic local theories, designated by them as objective local theories, would display in the considered experimental circumstances. For completeness we summarise here the features of the model situation contemplated in . A common source emits pairs of correlated entities that fly or propagate in opposite directions. When one of them reaches a detector, a characteristic feature of the entity, say the way it lies in a plane normal to the direction of motion, can be measured. Since the putative theory has to be an objective one, the entity is supposed to possess this feature in a way independent of any act of measurement. The overall behaviour of the single pair is characterized by a “hidden” parameter $`\lambda `$ in the following sense: $`\lambda `$ would specify the initial state of the pair and its subsequent development when the mutual interaction of the components has stopped, were it not for a residual element of uncertainty. The authors of do not specify the nature of this uncertainty <sup>3</sup><sup>3</sup>3The relation of equivalence between deterministic and stochastic hidden-variables models was first elucidated by A. Fine.. Anyway, a residual randomness is left, maybe intrinsic, maybe just due to our laziness in investigating the system. Therefore it makes sense to speak of the conditional probability $`p(A|\lambda )`$ that the entity that propagates, say, to the left, when reaching the detector set on the left will display there a characteristic feature, for instance an angle $`A`$, when the hidden parameter has the value $`\lambda `$. Let $`p(B|\lambda )`$ be the conditional probability that the entity going to the right be detected to display as characteristic feature an angle $`B`$ when the hidden parameter has the value $`\lambda `$. Clauser and Horne make an assumption, which they claim to be “a natural expression of a field-theoretical point of view, which in turn is an extrapolation from the common-sense view that there is no action at a distance”. They stipulate that the local objective theories should obey the following factorisability condition: (2.1) $$p(A,B|\lambda )=p(A|\lambda )p(B|\lambda )$$ for the joint probability $`p(A,B|\lambda )`$ of detecting the left entity as displaying the angle $`A`$ and the right entity as displaying the angle $`B`$ when the hidden parameter has the value $`\lambda `$. Let $`A`$, $`A^{}`$, and $`B`$, $`B^{}`$ be the four angles that happen to be displayed by the propagating entities at their respective acts of detection when the hidden parameter has the value $`\lambda `$. It is an easy matter to show that the inequalities $`1p(A|\lambda )p(B|\lambda )p(A|\lambda )p(B^{}|\lambda )+p(A^{}|\lambda )p(B|\lambda )+p(A^{}|\lambda )p(B^{}|\lambda )`$ (2.2) $`p(A^{}|\lambda )p(B|\lambda )0`$ must hold. By integrating $`p(A|\lambda )`$ and $`p(A,B|\lambda )`$ over $`\lambda `$ with a suitably normalised weight function $`\rho (\lambda )`$ one defines the probability (2.3) $$p(A)\rho (\lambda )p(A|\lambda )𝑑\lambda ,$$ the joint probability (2.4) $$p(A,B)\rho (\lambda )p(A|\lambda )p(B|\lambda )𝑑\lambda ,$$ and eventually finds the now famous Clauser-Horne inequalities (2.5) $`1p(A,B)p(A,B^{})+p(A^{},B)+p(A^{},B^{})p(A^{})p(B)0`$ as a necessary condition that all the objective local theories must fulfil. One cannot help agreeing with Clauser and Horne that the factorisability condition (2.1) is a necessary one in the objective local theories; one can add that, apart from the above mentioned exceptions , , die Physik der Modelle generally requires the satisfaction of such a condition, hence the general validity of the Clauser-Horne inequalities. When the quantum mechanical probabilities, let us say $`q(A)`$, $`q(A,B)`$, etc., calculated for particular instances dealing either with correlated spins or with correlated light quanta are substituted for the corresponding probabilities in (2.5), it is found that, if the angles are appropriately chosen, the resulting expression does not fulfil the inequalities. According to the established wisdom, this is the hallmark of quantum mechanics, well confirmed by the experimental facts (apart from residual loopholes), a unique feature that cannot be mimicked by classical physics, unless one contemplates either subtle enhancement processes connected e.g. with the existence of a zero-point electromagnetic field or a nonlocal behaviour, like the one stemming from the Lorentz-Abraham-Dirac equation of motion for the classical electron . In recent years, however, rather lonely but persistent voices have been heard , , , , , claiming inter alia that some not so venial sin is committed when substituting the quantum mechanical probabilities for the probabilities $`p(A)`$, $`p(A,B)`$ in (2.5), and that when the wrongdoing is amended, the Clauser-Horne inequalities happen to be no longer violated. More than seven decades have elapsed since the very notion of quantum probabilities has been hesitantly extracted , by Born from Einstein’s idea of the Gespensterfeld, and yet one feels some uneasiness when forced to confront a seemingly simple question like this. One would venture in this sort of discussion with a lighter heart, had the number of interpretations attached since then to quantum mechanics shrinked instead of increasing, as it has been the case, and provided that the quantum measurement problem had already been solved for good, i.e. more relativistico. While waiting for such occurrences, one still feels to walk on firmer ground if once more this admittedly old fashioned and just preliminary approach is attempted: seeking whether a more familiar simile, entirely grounded on classical physics and on classical probability theory, may provide some enlightenment. ## 3. A classical model that “violates” the inequalities Let us consider an experimental device like the one drawn in Fig. 1, whose working is entirely accounted for by classical mechanics: two equal cylindrical bodies, “1” and “2”, are thrown from the spatial origin of an inertial reference frame with opposite velocities $`v_1=v_2`$ and with opposite angular velocities $`\omega _1=\omega _2`$, all directed along the $`y`$ axis, by the action of, say, a pressed and twisted spring interposed between them and suddenly released. The action of the spring lasts for a very short time, after which the two material bodies are free to run and to turn around their axes in the interior of a hollow cylinder of length $`l`$, kept at rest in the considered inertial reference system. The hollow cylinder is centered at the origin of the spatial coordinates and its axis lies in the $`y`$ direction; its scope is to act as measuring device. To this end on its left half, along the whole span $`l/2<y<0`$, two straight marks are engraved, whose projections on the $`x,z`$-plane happen to lay at the angles $`A`$ and $`A^{}`$ with respect to the $`x`$ axis. Two straight marks are engraved also on the right half of the hollow cylinder, along its whole span $`0<y<l/2`$, at the angles $`B`$ and $`B^{}`$ respectively. On the inner rims of the cylindrical bodies “1” and “2” two tiny marks $`m_1`$ and $`m_2`$ are impressed, and we shall abide to the rule that, after the pressing and twisting of the spring is accomplished, the two marks shall happen to coincide at some angle $`\phi `$ measured as previously in the $`x,z`$-plane. We shall also take care to load the spring always in the same way; the rotation angles of the two bodies, when they run from the center to the ends of the measuring device, shall be invariably $`\gamma _1=const.>0`$ for the body running to the left, and $`\gamma _2=\gamma _1`$ for the other one. Let us indulge, with the given apparatus, in the following exercise of “experimental probability”. We pick at random some angle $`\phi `$ from a uniform distribution that extends between $`0`$ and $`2\pi `$, we load and twist the spring interposed between the bodies as prescribed above, and we eventually set the coincident marks $`m_1`$ and $`m_2`$ at the angle $`\phi `$ before releasing the spring. We then check whether the mark on body “1”, that runs to the left, crosses or not the straight lines drawn at the angles $`A`$ and $`A^{}`$, and whether the mark on body “2” reaches or not the angles $`B`$ and $`B^{}`$. After repeating ab ovo the whole procedure $`N`$ times, the statistics of the experiment can be compiled, and the “experimental probability” can be eventually inferred. Of course nobody will really indulge in so trivial an experiment, since the physical situation is quite clear, and the classical probabilities $`p(A)`$, $`p(A,B)`$, etc. can be directly evaluated through a very simple geometric argument. Needless to say, when these probabilities are inserted in the Clauser-Horne inequalities (2.5), the latter happen to be fulfilled for all the possible choices of the angles $`A`$, $`A^{}`$, and $`B`$, $`B^{}`$. Suppose however that we are caught by a virulent form of wishful thinking: we would like to see these inequalities violated, despite the fact that the instruments at our disposal are constituted by purely classical, objective and local entities. Therefore, in order to mimic the quantum mechanical behaviour, we modify the experimental device described above, and we conjure up an illusive “Verschränkung” by: * connecting the bodies “1” and “2” through a mechanical constraint, that hinders their rotations when the mutual rotation angle reaches the value $`\gamma `$, in order to provide a not quite mysterious surrogate to the “spooky action at a distance”, * allowing for an active role of the measuring device, through mechanical stops placed in the interior of the hollow cylinder, respectively at an angle, say $`A`$, on the left side, acting in the span $`l/2<y<0`$, and at an angle $`B`$ on the right side, acting in the span $`0<y<l/2`$. These stops are so contrived as to block the motion of the bodies “1” and “2” whenever the tiny marks $`m_1`$ and $`m_2`$ impressed on their inner rims respectively reach the above mentioned angles. It is intended that we can change the position of the mechanical stops from the pair of angles $`A,B`$ to the pairs $`A,B^{}`$, $`A^{},B`$ and $`A^{},B^{}`$, or even suppress one of the stops, while leaving the other one active and set just at one of the four positions mentioned above. We test the modified device by choosing the angles as shown in Figure 2, i.e. we set $`B^{}=0`$, $`B=\vartheta `$, $`A^{}=\gamma `$, $`A=\gamma +\vartheta `$, with $`0<\vartheta <\gamma `$; the spring is loaded just in the way kept earlier, that entails a clockwise rotation of the body “2”. It is our intention to perform again a sequel of trials with the start angle $`\phi `$ picked at random in a uniform distribution between $`0`$ and $`2\pi `$, as it was appropriate with the earlier version of the apparatus, but our plan meets with a certain difficulty: while previously just one series of trials, corresponding to just one experimental setup, was sufficient for inferring all the “experimental probabilities”, the situation now is completely different. In order to gather the statistics required for inferring the joint probabilities four distinct sequences of trials are needed, one for each of the physically different setups that can occur when both mechanical stops are active. Furthermore, in order to infer the probabilities of the single occurrences, four additional sequences of trials seem needed, one for each of the physically different situations that can occur when one of the mechanical stops is removed. What we infer from the statistics of these eight independent sequences of trials, performed each one on a different physical system, are of course conditional probabilities. They are the probabilities for the bodies to reach certain angles when the stops are placed in a certain way. Also in this case there is no need to perform really the sequences of trials. Let $`p(A,B|a,b)`$ mean the conditional probability that the marks on the two bodies reach the angles $`A`$ and $`B`$ when the mechanical stops are both active and set just at the angles $`a=A`$, $`b=B`$, while $`p(A|a)`$ means the conditional probability that the mark on body “1” reach the angle $`A`$ when only the mechanical stop on the left is active, and set at the angle $`a=A`$. Simple geometric arguments give (3.1) $$p(A,B|a,b)=p(A^{},B^{}|a^{},b^{})=\frac{\gamma }{2\pi },p(A,B^{}|a,b^{})=0,p(A^{},B|a^{},b)=\frac{\gamma \vartheta }{2\pi },$$ while (3.2) $$p(A|a)=p(A^{}|a^{})=p(B|b)=p(B^{}|b^{})=\frac{\gamma }{4\pi }.$$ It is quite clear that we should abstain from the inconsiderate act of plugging these quantities in the Clauser-Horne inequalities (2.5), since each one of them is a conditional probability referring to a different physical system, while the inequalities deal with the probabilites for just one physical system. If we insist in doing so, a nonsensical outcome must be expected, and is in fact obtained. The inequality (2.5) is “violated” on the right, despite the fact that classical mechanics fully accounts in an objective and local way for all the physical happenigs considered here. Moreover, through the same hocus-pocus also the inequality (3.3) $`1p(A^{},B^{})p(A^{},B)+p(A,B^{})+p(A,B)p(A)p(B^{})0,`$ that can be obtained from (2.5) by exchanging the angles with and without a prime, is “violated” on the right. By summing the inequalities stemming from the two “violations” one would get (3.4) $$2p(A,B)+2p(A^{},B^{})p(A)p(A^{})p(B)p(B^{})>0.$$ Of course one expects that in an entirely classical situation at least Bayes’ axiom (3.5) $$p(A,B)=p(A)p(B|A)=p(B)p(A|B)$$ should be respected. But if we insist on this requirement and insert (3.5) in (3.4), we eventually reach an unquestionably absurd result: at least one of the four conditional probabilities $`p(A|B)`$, $`p(B|A)`$, $`p(A^{}|B^{})`$, $`p(B^{}|A^{})`$ should be $`>1`$! By inserting the conditional probabilities (3.1) and (3.2) in the Clauser-Horne inequalities we have just committed a blunder. But the way out is not so arduous, and we could have spared the useless trials with one stop removed. The four trials with the two stops active are in fact sufficient for mastering the problem, if we confront it in the correct way. The measuring device, after the addition of the mechanical stops, is no longer a passive entity. Due to their presence the physical system under investigation is not constituted only by the bodies “1” and “2” and by the mechanical constraint that hinders mutual rotations of the latter larger than the angle $`\gamma `$. The physical system now includes the measuring device with its mechanical stops, and the probabilities that enter the Clauser-Horne inequalities shall deal with this larger system. Let us discard the trials that were performed with only one stop active, and consider the ratios between the number of trials performed with a certain setting of the two stops and the overall number of trials performed with the two stops both active. From these ratios we infer, in retrospect, the probabilities for the various settings of the stops. We call them $`p(a,b)`$, $`p(a,b^{})`$, $`p(a^{},b)`$, $`p(a^{},b^{})`$. Then the probabilities concerning the whole experiment, that we are entitled to insert in the Clauser-Horne inequalities, shall read: $`p(A,B)=p(A,B|a,b)p(a,b),p(A,B^{})=p(A,B^{}|a,b^{})p(a,b^{}),`$ $`p(A^{},B)=p(A^{},B|a^{},b)p(a^{},b),p(A^{},B^{})=p(A^{},B^{}|a^{},b^{})p(a^{},b^{}),`$ $`p(A)=p(A,B)+p(A,B^{}),p(A^{})=p(A^{},B)+p(A^{},B^{}),`$ (3.6) $`p(B)=p(A,B)+p(A^{},B),p(B^{})=p(A,B^{})+p(A^{},B^{}).`$ By inserting (3) in (2.5) the latter reduces to (3.7) $$1p(A,B^{}|a,b^{})p(a,b^{})p(A^{},B|a^{},b)p(a^{},b)0$$ which is of course not violated in general both on its right and on its left side. ## 4. Conclusion We ask for the reader’s forgiveness, because the scrutiny of the detailed working of our artful “Verschränkung” may well have looked a pedantic recitation of things so well known that their further recollection is just a waste of time. But what is given for granted in classical physics does not seem to be so well settled in quantum physics if, after so many years since the appearance of the Clauser-Horne inequalities and after so many papers written on the subject, some authors , , still felt urged to stand up and to remark that a major conceptual error, akin to the one expounded in the previous Section, has been committed in quantum physics. They have shown also that if the quantum probabilities are dealt with as conditional probabilities and the statistics of the use of the detectors is taken into account, quantum mechanics no longer happens to violate the inequalities in the situations that have been so carefully scrutinised, both by theoreticians and by experimentalists. The readings of the above mentioned papers and of the seminal work by C. Brans have convinced us that even a recollection of the obvious, as it has been done with the classical simile of the present paper, could be helpful in intimating what might be the proper use of the Clauser-Horne inequalities in quantum mechanics.
warning/0004/math0004051.html
ar5iv
text
# Spectra and symmetric spectra in general model categories ## Introduction The object of this paper is to give two very general constructions of the passage from unstable homotopy theory to stable homotopy theory. Since homotopy theory in some form appears in many different areas of mathematics, this construction is useful beyond algebraic topology, where these methods originated. In particular, the two constructions we give apply not only to the usual passage from unstable homotopy theory of pointed topological spaces (or simplicial sets) to the stable homotopy theory of spectra, but also to the passage from the unstable $`𝔸^1`$-homotopy theory of Morel-Voevodsky to the stable $`𝔸^1`$-homotopy theory. This example is obviously important, and the fact that it is an example of a widely applicable theory of stabilization may come as a surprise to readers of , where specific properties of sheaves are used. Suppose, then, that we are given a (Quillen) model category $`𝒞`$ and a functor $`G:𝒞\stackrel{}{}𝒞`$ that we would like to invert, analogous to the suspension. We will clearly need to require that $`𝒞`$ be compatible with the model structure; specifically, we require $`G`$ to be a left Quillen functor. We will also need some technical hypotheses on the model category $`𝒞`$, which are complicated to state and to check, but which are satisfied in almost all interesting examples, including $`𝔸^1`$-homotopy theory. It is well-known what one should do to form the category $`Sp^{}(𝒞,G)`$ of spectra, as first written down for topological spaces in . An object of $`Sp^{}(𝒞,G)`$ is a sequence $`X_n`$ of objects of $`𝒞`$ together with maps $`GX_n\stackrel{}{}X_{n+1}`$, and a map $`f:X\stackrel{}{}Y`$ is a sequence of maps $`f_n:X_n\stackrel{}{}Y_n`$ compatible with the structure maps. There is an obvious model structure, called the *projective model structure*, where the weak equivalences are the maps $`f:X\stackrel{}{}Y`$ such that $`f_n`$ is a weak equivalence for all $`n`$. It is not difficult to show that this is a model structure and that there is a left Quillen functor $`G:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$ extending $`G`$ on $`𝒞`$. But, just as in the topological case, $`G`$ will not be a Quillen equivalence. So we must localize the projective model structure on $`Sp^{}(𝒞,G)`$ to produce the *stable* model structure, with respect to which $`G`$ will be a Quillen equivalence. A new feature of this paper is that we are able to construct the stable model structure with minimal hypotheses on $`𝒞`$, using the localization results of Hirschhorn (based on work of Dror-Farjoun ). We must pay a price for this generality, of course. That price is that stable equivalences are not stable homotopy isomorphisms, but instead are cohomology isomorphisms on all cohomology theories, just as for symmetric spectra . If enough hypotheses are put on $`𝒞`$ and $`G`$, then we show that stable equivalences are stable homotopy isomorphisms. Jardine proves that stable equivalences are stable homotopy isomorphisms in the stable $`𝔸^1`$-homotopy theory, using the Nisnevitch descent theorem. His result does not follow from our general theorem for the Morel-Voevodsky motivic model category, because the hypotheses we need do not hold there, but Voevodsky (personal communication) has constructed a simpler model category equivalent to the Morel-Voevodsky one that does satify our hypotheses. As is well-known in algebraic topology, the category $`Sp^{}(𝒞,G)`$ is not sufficient to understand the smash product. That is, if $`𝒞`$ is a symmetric monoidal model category, and $`G`$ is the functor $`XXK`$ for some cofibrant object $`K`$ of $`𝒞`$, it almost never happens that $`Sp^{}(𝒞,G)`$ is symmetric monoidal. We therefore need a different construction in this case. We define a category $`Sp^\mathrm{\Sigma }(𝒞,K)`$ just as in symmetric spectra . That is, an object of $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is a sequence $`X_n`$ of objects of $`𝒞`$ with an action of the symmetric group $`\mathrm{\Sigma }_n`$ on $`X_n`$. In addition, we have $`\mathrm{\Sigma }_n`$-equivariant structure maps $`X_nK\stackrel{}{}X_{n+1}`$, but we must further require that the iterated structure maps $`X_nK^p\stackrel{}{}X_{n+p}`$ are $`\mathrm{\Sigma }_n\times \mathrm{\Sigma }_p`$-equivariant, where $`\mathrm{\Sigma }_p`$ acts on $`K^p`$ by permuting the tensor factors. It is once again straightforward to construct the projective model structure on $`Sp^\mathrm{\Sigma }(𝒞,K)`$. The same localization methods developed for $`Sp^{}(𝒞,G)`$ apply again here to give a stable model structure on which tensoring with $`K`$ is a Quillen equivalence. Once again, stable equivalences are cohomology isomorphisms on all possible cohomology theories, but this time it is very difficult to give a better description of stable equivalences even in the case of simplicial symmetric spectra (but see for the best such result I know). We point out that our construction gives a different construction of the stable model category of simplicial symmetric spectra than the one appearing in . We now have competing stabilizations of $`𝒞`$ under the tensoring with $`K`$ functor when $`𝒞`$ is symmetric monoidal. Naturally, we need to prove they are the same in an appropriate sense. This was done in the topological (actually, simplicial) case in by constructing a functor $`Sp^{}(𝒞,G)\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K)`$, where $`K=S^1`$ and $`G`$ is the tensor with $`S^1`$ functor, and proving it is a Quillen equivalence. We are unable to generalize this argument. Instead, following an idea of Hopkins, we construct a zigzag of Quillen equivalences $`Sp^{}(𝒞,G)\stackrel{}{}𝒞\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K)`$. However, we need to require that the cyclic permutation map on $`KKK`$ be homotopic to the identity by an explicit homotopy for our construction to work. This hypothesis holds in the topological case with $`K=S^1`$ and in the $`𝔸^1`$-local case with $`K`$ equal to either the simplicial circle or the algebraic circle $`𝔸^1\{0\}`$. This section of the paper is by far the most delicate, and it is likely that we do not have the best possible result. We also investigate the properties of these two stabilization constructions. There are some obvious properties one would hope for of a stabilization construction such as $`Sp^{}(𝒞,G)`$ or $`Sp^\mathrm{\Sigma }(𝒞,K)`$. First of all, it should be functorial in the pair $`(𝒞,G)`$. We prove this for both stabilization constructions; the most difficult point is defining what one should mean by a map from $`(𝒞,G)`$ to $`(𝒟,H)`$. Furthermore, it should be homotopy invariant. That is, if the map $`(𝒞,G)\stackrel{}{}(𝒟,H)`$ is a Quillen equivalence, the induced map of stabilizations should also be a Quillen equivalence. We also prove this; one corollary is that the Quillen equivalence class of $`Sp^\mathrm{\Sigma }(𝒞,K)`$ depends only on the homotopy type of $`K`$. Finally, the stabilization map $`𝒞\stackrel{}{}Sp^{}(𝒞,G)`$ should be the initial map to a model category $`𝒟`$ with an extension of $`G`$ to a Quillen equivalence. However, this last statement seems to be asking for too much, because the category of model categories is itself something like a model category. This statement is analogous to asking for an initial map in a model category from $`X`$ to a fibrant object, and such things do not usually exist. The best we can do is to say that if $`G`$ is already a Quillen equivalence, then the map from $`𝒞\stackrel{}{}Sp^{}(𝒞,G)`$ is a Quillen equivalence. This gives a weak form of uniqueness, and is the basis for the comparison between $`Sp^{}(𝒞,G)`$ and $`Sp^\mathrm{\Sigma }(𝒞,K)`$. See also see and for uniqueness results for the usual stable homotopy category. We point out that this paper leaves some obvious questions open. We do not have a good characterization of stable equivalences or stable fibrations in either spectra or symmetric spectra, in general, and we are unable to prove that spectra or symmetric spectra are right proper. We do have such characterizations for spectra when the original model category $`𝒞`$ is sufficiently well-behaved, and the adjoint $`U`$ of $`G`$ preserves sequential colimits. These hypotheses include the cases of ordinary simplicial spectra and spectra in a new motivic model category of Voevodsky (but not the original Morel-Voevodsky motivic model category). We also prove that spectra are right proper in this situation. But we do not have a characterization of stable equivalences of symmetric spectra even with these strong assumptions. Also, we have been unable to prove that symmetric spectra satisfy the monoid axiom. Without the monoid axiom, we do not get model categories of monoids or of modules over an arbitrary monoid, though we do get a model category of modules over a cofibrant monoid. The question of whether commutative monoids form a model category is even more subtle and is not addressed in this paper. See for commutative monoids in symmetric spectra of topological spaces. There is a long history of work on stabilization, much of it not using model categories. As far as this author knows, Boardman was the first to attempt to construct a good point-set version of spectra; his work was never published (but see ), but it was the standard for many years. Generalizations of Boardman’s construction were given by Heller in several papers, including and . Heller has continued work on these lines, most recently in . The review of this paper in Mathematical Reviews by Tony Elmendorf (MR98g:55021) captures the response of many algebraic topologists to Heller’s approach. I believe the central idea of Heller’s approach is that the homotopy theory associated to a model category $`𝒞`$ is the collection of all possible homotopy categories of diagram categories $`\text{ho }𝒞^I`$ and all functors between them. With this definition, one can then forget one had the model category in the first place, as Heller does. Unfortunately, the resulting complexity of definition is overwhelming at present. Of course, there has also been very successful work on stabilization by May and coauthors, the two major milestones being and . At first glance, May’s approach seems wedded to the topological situation, relying as it does on homeomorphisms $`X_n\stackrel{}{}\mathrm{\Omega }X_{n+1}`$. This is the reason we have not tried to use it in this paper. However, there has been considerable recent work showing that this approach may be more flexible than one might have expected. I have mentioned above, but perhaps the most ambitious attempt to generalize $`S`$-modules has been initiated by Mark Johnson . Finally, we point out that Schwede has shown that the methods of Bousfield and Friedlander apply to certain more general model categories. His model categories are always simplicial and proper, and he is always inverting the ordinary suspension functor. Nevertheless, the paper is the first serious attempt to define a general stabilization functor of which the author is aware. This paper is organized as follows. We begin by defining the category $`Sp^{}(𝒞,G)`$ and the associated strict model structure in Section 1. Then there is the brief Section 2 recalling Hirschhorn’s approach to localization of model categories. We then construct the stable model structure modulo certain technical lemmas in Section 3. The technical lemmas we need assert that if a model category $`𝒞`$ is left proper cellular, then so is the strict model structure on $`Sp^{}(𝒞,G)`$, and therefore we can apply the localization technology of Hirschhorn. We prove these technical lemmas, and the analogous lemmas for the strict model structure on $`Sp^\mathrm{\Sigma }(𝒟,K)`$, in an Appendix. In Section 4, we study the simplifications that arise when the adjoint $`U`$ of $`G`$ preserves sequential colimits and $`𝒞`$ is sufficiently well-behaved. We characterize stable equivalences as the appropriate generalization of stable homotopy isomorphisms in this case, and we show the stable model structure is right proper, giving a description of the stable fibrations as well. In Section 5, we prove the functoriality, homotopy invariance, and homotopy idempotence of the construction $`(𝒞,G)Sp^{}(𝒞,G)`$. We also investigate monoidal structure. Section 6 begins the second part of the paper, about symmetric spectra. Since we have developed all the necessary techniques in the first part, the proofs in this part are more concise. In Section 6 we discuss the category of symmetric spectra. In Section 7 we construct the projective and stable model structures on symmetric spectra, and in Section 8, we discuss some properties of symmetric spectra. This includes functoriality, homotopy invariance, and homotopy idempotence of the stable model structure. We conclude the paper in Section 9 by constructing the chain of Quillen equivalences betwee $`Sp^{}(𝒞,G)`$ and $`Sp^\mathrm{\Sigma }(𝒞,K)`$, under the cyclic permutation hypothesis mentioned above. Finally, as stated previously, there is an Appendix verifying that the techniques of Hirschhorn can be applied to the projective model structures on $`Sp^{}(𝒞,G)`$ and $`Sp^\mathrm{\Sigma }(𝒞,K)`$. Obviously, considerable familiarity with model categories will be necessary to understand this paper. The original reference is , but a better introductory reference is . More in depth references include , , and . In particular, we rely heavily on the localization technology in . The author would like to thank Dan Dugger, Phil Hirschhorn, Mike Hopkins, Dan Kan, Stefan Schwede, Brooke Shipley, Jeff Smith, Markus Spitzweck, and Vladimir Voevodsky for helpful conversations about this paper. In particular, to the author’s knowledge, it is Jeff Smith’s vision that one should be able to stabilize an arbitrary model category, a vision that could not be carried out without Phil Hirschhorn’s devotion to getting the localization theory of model categories right. The idea of using almost finitely generated model categories in Section 4 is due to Voevodsky, and the idea of using bispectra to compare symmetric spectra with ordinary spectra (see Section 9) is due to Hopkins. ## 1. Spectra In this section and throughout the paper, $`𝒞`$ will be a model category and $`G:𝒞\stackrel{}{}𝒞`$ will be a left Quillen endofunctor of $`𝒞`$ with right adjoint $`U`$. In this section, we define the category $`Sp^{}(𝒞,G)`$ of spectra and construct its strict model structure. The following definition is a straightforward generalization of the usual notion of spectra . ###### Definition 1.1. Suppose $`G`$ is a left Quillen endofunctor of a model category $`𝒞`$. Define $`Sp^{}(𝒞,G)`$, the category of *spectra*, as follows. A *spectrum* $`X`$ is a sequence $`X_0,X_1,\mathrm{},X_n,\mathrm{}`$ of objects of $`𝒞`$ together with structure maps $`\sigma :GX_n\stackrel{}{}X_{n+1}`$ for all $`n`$. A *map of spectra* from $`X`$ to $`Y`$ is a collection of maps $`f_n:X_n\stackrel{}{}Y_n`$ commuting with the structure maps, so that the diagram $$\begin{array}{ccc}GX_n& \stackrel{\sigma _X}{}& X_{n+1}\\ Gf_n& & f_{n+1}& & \\ GY_n& \underset{\sigma _Y}{}& Y_{n+1}\end{array}$$ is commutative for all $`n`$. Note that if $`𝒞`$ is either the model category of pointed simplicial sets or the model category of pointed topological spaces, and $`G`$ is the suspension functor given by smashing with the circle $`S^1`$, then $`Sp^{}(𝒞,G)`$ is the Bousfield-Friedlander category of spectra . ###### Definition 1.2. Given $`n0`$, the *evaluation functor* $`\mathrm{Ev}_n:Sp^{}(𝒞,G)\stackrel{}{}𝒞`$ takes $`X`$ to $`X_n`$. The evaluation functor has a left adjoint $`F_n:𝒞\stackrel{}{}Sp^{}(𝒞,G)`$ defined by $`(F_nX)_m=G^{mn}X`$ if $`mn`$ and $`(F_nX)_m=0`$ otherwise, where $`0`$ is the initial object of $`𝒞`$. The structure maps are the obvious ones. Note that $`F_0`$ is an full and faithful embedding of the category $`𝒞`$ into $`Sp^{}(𝒞,G)`$. Limits and colimits in the category of spectra are taken objectwise. ###### Lemma 1.3. The category of spectra is bicomplete. ###### Proof. Given a functor $`X:\stackrel{}{}Sp^{}(𝒞,G)`$, we define $`(\mathrm{colim}X)_n=\mathrm{colim}\mathrm{Ev}_nX`$ and $`(\mathrm{lim}X)_n=\mathrm{lim}\mathrm{Ev}_nX`$. Since $`G`$ is a left adjoint, it preserves colimits. The structure maps of the colimit are then the composites $$G(\mathrm{colim}\mathrm{Ev}_nX)\mathrm{colim}(G\mathrm{Ev}_nX)\stackrel{\mathrm{colim}\sigma X}{}\mathrm{colim}\mathrm{Ev}_{n+1}X.$$ Although $`G`$ does not preserve limits, there is still a natural map $`G(\mathrm{lim}Y)\stackrel{}{}\mathrm{lim}GY`$ for any functor $`Y:\stackrel{}{}𝒞`$. Then the structure maps of the limit are the composites $$G(\mathrm{lim}\mathrm{Ev}_nX)\stackrel{}{}\mathrm{lim}(G\mathrm{Ev}_nX)\stackrel{\mathrm{lim}\sigma X}{}\mathrm{lim}\mathrm{Ev}_{n+1}X.\mathit{}$$ ###### Remark 1.4. Note that the evaluation functor $`\mathrm{Ev}_n:Sp^{}(𝒞,G)\stackrel{}{}𝒞`$ preserves colimits, so should have a right adjoint $`R_n:𝒞\stackrel{}{}Sp^{}(𝒞,G)`$. We define $`(R_nX)_m=U^{nm}X`$ if $`mn`$, and $`(R_nX)_m=1`$ if $`m>n`$. The structure map $`GU^{nm}X\stackrel{}{}U^{nm1}X`$ is adjoint to the identity map of $`U^{nm}X`$ when $`m<n`$. We leave it to the reader to verify that $`R_n`$ is the right adjoint of $`\mathrm{Ev}_n`$. We now show that the functors $`G`$ and $`U`$ extend to functors on $`Sp^{}(𝒞,G)`$. ###### Lemma 1.5. Suppose $`F:𝒞\stackrel{}{}𝒞`$ is a functor and $`\tau :GF\stackrel{}{}FG`$ is a natural transformation. Then there is an induced functor $`F:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$, called the *prolongation* of $`F`$. ###### Proof. Given $`XSp^{}(𝒞,G)`$, we define $`(FX)_n=FX_n`$, with structure maps $`GFX_n\stackrel{𝜏}{}FGX_n\stackrel{F\sigma }{}FX_{n+1}`$. Given a map $`f:X\stackrel{}{}Y`$, we define $`Ff`$ by $`(Ff)_n=Ff_n`$. ∎ Note that the prolongation of $`F`$ depends on the choice of natural transformation $`\tau `$. Usually there is an obvious choice of $`\tau `$. ###### Corollary 1.6. The functors $`G`$ and $`U`$ prolong to adjoint functors $`G`$ and $`U`$ on $`Sp^{}(𝒞,G)`$. ###### Proof. To prolong $`G`$, take $`\tau `$ to be the identity in Lemma 1.5. To prolong its right adjoint $`U`$, take $`\tau `$ to be the composite $`GUX\stackrel{}{}X\stackrel{}{}UGX`$ of the counit and unit of the adjunction. It is a somewhat involved exercise in adjoint functors to verify that the resulting prolongations $`G^{}`$ and $`U^{}`$ are still adjoint to each other. Like all the exercises in adjoint functors in this paper, the simplest way to proceed is to write the adjoint $`Y\stackrel{}{}UZ`$ of a map $`f:GY\stackrel{}{}Z`$ as the composition $$Y\stackrel{𝜂}{}UGY\stackrel{Uf}{}UZ$$ and to use the standard properties of the unit and counit of an adjunction. ∎ The following remark is critically important to the understanding of our approach to spectra. ###### Remark 1.7. The definition we have just given of the prolongation of $`G`$ to an endofunctor of $`Sp^{}(𝒞,G)`$ is the only possible definition under our very general hypotheses. However, this definition does not generalize the definition of the suspension when $`𝒞`$ is the category of pointed topological spaces and $`GX=XS^1`$. Indeed, recall from that the suspension of a spectrum $`X`$ in this case is defined by $`(XS^1)_n=X_nS^1`$, with structure map given by $$X_nS^1S^1\stackrel{1T}{}X_nS^1S^1\stackrel{\sigma 1}{}X_{n+1}S^1,$$ where $`T`$ is the twist isomorphism. On the other hand, if we apply our definition of the prolongation of $`G`$ above, we get a functor $`XXS^1`$ defined by $`(XS^1)_n=X_nS^1`$ with structure map $$X_nS^1S^1\stackrel{\sigma 1}{}X_{n+1}S^1.$$ Said another way, Bousfield and Friedlander choose the natural transformation $`\tau :XS^1S^1\stackrel{}{}XS^1S^1`$ to be $`1T`$, while we are taking it to be the identity. This is a crucial and subtle difference whose ramifications we will study in Section 9. We now show that $`Sp^{}(𝒞,G)`$ inherits a model structure from $`𝒞`$, called the *projective model structure*. The functor $`G:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$ will be a left Quillen functor with respect to the projective model structure, but it will not be a Quillen equivalence. Our approach to the projective model structure owes much to and \[11, Section 5.1\]. At this point, we will slip into the standard model category terminology and notation, all of which can be found in , mostly in Section 2.1. ###### Definition 1.8. A map $`fSp^{}(𝒞,G)`$ is a *level equivalence* if each map $`f_n`$ is a weak equivalence in $`𝒞`$. Similarly, $`f`$ is a *level fibration* (resp.*level cofibration*, *level trivial fibration*, *level trivial cofibration*) if each map $`f_n`$ is a fibration (resp. cofibration, trivial fibration, trivial cofibration) in $`𝒞`$. The map $`f`$ is a *projective cofibration* if $`f`$ has the left lifting property with respect to every level trivial fibration. Note that level equivalences satisfy the two out of three property, and each of the classes defined above is closed under retracts. Thus we should be able to construct a model structure using these classes. To do so, we need the small object argument, and hence we assume that $`𝒞`$ is cofibrantly generated (see \[10, Section 2.1\] for a discussion of cofibrantly generated model categories). ###### Definition 1.9. Suppose $`𝒞`$ is a cofibrantly generated model category with generating cofibrations $`I`$ and generating trivial cofibrations $`J`$. Suppose $`G`$ is a left Quillen endofunctor of $`𝒞`$, and form the category of spectra $`Sp^{}(𝒞,G)`$. Define sets of maps in $`Sp^{}(𝒞,G)`$ by $`I_G=_nF_nI`$ and $`J_G=_nF_nJ`$. The sets $`I_G`$ and $`J_G`$ will be the generating cofibrations and trivial cofibrations for a model structure on $`Sp^{}(𝒞,G)`$. There is a standard method for proving this, based on the small object argument \[10, Theorem 2.1.14\]. The first step is to show that the domains of $`I_G`$ and $`J_G`$ are small, in the sense of \[10, Definition 2.1.3\]. ###### Proposition 1.10. Suppose $`A`$ is small relative to the cofibrations (resp. trivial cofibrations) in $`𝒞`$, and $`n0`$. Then $`F_nA`$ is small relative to the level cofibrations (resp. level trivial cofibrations) in $`Sp^{}(𝒞,G)`$. ###### Proof. The main point is that $`\mathrm{Ev}_n`$ commutes with colimits. We leave the remainder of the proof to the reader. ∎ To apply this to the domains of $`I_G`$, we need to know that the maps of $`I_G\text{-cof}`$ are level cofibrations. Recall the right adjoint $`R_n`$ of $`\mathrm{Ev}_n`$ constructed in Remark 1.4. ###### Lemma 1.11. A map in $`Sp^{}(𝒞,G)`$ is a level cofibration (resp. level trivial cofibration) if and only if it has the left lifting property with respect to $`R_ng`$ for all $`n0`$ and all trivial fibrations (resp. fibrations) $`g`$ in $`𝒞`$. ###### Proof. By adjunction, a map $`f`$ has the left lifting property with respect to $`R_ng`$ if and only if $`\mathrm{Ev}_nf`$ has the left lifting property with respect to $`g`$. Since a map is a cofibration (resp. trivial cofibration) in $`𝒞`$ if and only if it has the left lifting property with respect to all trivial fibrations (resp. fibrations), the lemma follows. ∎ ###### Proposition 1.12. Every map in $`I_G\text{-cof}`$ is a level cofibration. Every map in $`J_G\text{-cof}`$ is a level trivial cofibration. ###### Proof. Since $`G`$ is a left Quillen functor, every map in $`I_G`$ is a level cofibration. By Lemma 1.11, this means that $`R_ngI_G\text{-inj}`$ for all $`n0`$ and all trivial fibrations $`g`$. Since a map in $`I_G\text{-cof}`$ has the left lifting property with respect to every map in $`I_G\text{-inj}`$, in particular it has the left lifting property with respect to $`R_ng`$. Another application of Lemma 1.11 completes the proof for $`I_G\text{-cof}`$. The proof for $`J_G\text{-cof}`$ is similar. ∎ Proposition 1.10 and Proposition 1.12 immediately imply the following corollary. ###### Corollary 1.13. The domains of $`I_G`$ are small relative to $`I_G\text{-cof}`$. The domains of $`J_G`$ are small relative to $`J_G\text{-cof}`$. ###### Theorem 1.14. Suppose $`𝒞`$ is cofibrantly generated. Then the projective cofibrations, the level fibrations, and the level equivalences define a cofibrantly generated model structure on $`Sp^{}(𝒞,G)`$, with generating cofibrations $`I_G`$ and generating trivial cofibrations $`J_G`$. We call this the *projective model structure*. The projective model structure is left proper (resp. right proper, proper) if $`𝒞`$ is left proper (resp. right proper, proper). Note that if $`𝒞`$ is either the model category of pointed simplicial sets or pointed topological spaces, and $`G`$ is the suspension functor, the projective model structure on $`𝒞_G`$ is the strict model structure on the Bousfield-Friedlander category of spectra . ###### Proof. The retract and two out of three axioms are immediate, as is the lifting axiom for a projective cofibration and a level trivial fibration. By adjointness, a map is a level trivial fibration if and only if it is in $`I_G\text{-inj}`$. Hence a map is a projective cofibration if and only if it is in $`I_G\text{-cof}`$. The small object argument \[10, Theorem 2.1.14\] applied to $`I_G`$ then produces a functorial factorization into a projective cofibration followed by a level trivial fibration. Adjointness implies that a map is a level fibration if and only if it is in $`J_G\text{-inj}`$. We have already seen in Proposition 1.12 that the maps in $`J_G\text{-cof}`$ are level equivalences, and they are projective cofibrations since they have the left lifting property with respect to all level fibrations, and in particular level trivial fibrations. Hence the small object argument applied to $`J_G`$ produces a functorial factorization into a projective cofibration and level equivalence followed by a level fibration. Conversely, we claim that any projective cofibration and level equivalence $`f`$ is in $`J_G\text{-cof}`$, and hence has the left lifting property with respect to level fibrations. To see this, write $`f=pi`$ where $`i`$ is in $`J_G\text{-cof}`$ and $`p`$ is in $`J_G\text{-inj}`$. Then $`p`$ is a level fibration. Since $`f`$ and $`i`$ are both level equivalences, so is $`p`$. Thus $`f`$ has the left lifting property with respect to $`p`$, and so $`f`$ is a retract of $`i`$ by the retract argument \[10, Lemma 1.1.9\]. In particular $`fJ_G\text{-cof}`$. Since colimits and limits in $`Sp^{}(𝒞,G)`$ are taken levelwise, and since every projective cofibration is in particular a level cofibration, the statements about properness are immediate. ∎ We also characterize the projective cofibrations. We denote the pushout of two maps $`A\stackrel{}{}B`$ and $`A\stackrel{}{}C`$ by $`B_AC`$. ###### Proposition 1.15. A map $`i:A\stackrel{}{}B`$ is a projective (trivial) cofibration if and only if the induced maps $`A_0\stackrel{}{}B_0`$ and $`A_n_{GA_{n1}}GB_{n1}\stackrel{}{}B_n`$ for $`n1`$ are (trivial) cofibrations. ###### Proof. We only prove the cofibration case, leaving the similar trivial cofibration case to the reader. First suppose $`i:A\stackrel{}{}B`$ is a projective cofibration. We have already seen in Proposition 1.12 that $`A_0\stackrel{}{}B_0`$ is a cofibration. Suppose $`p:X\stackrel{}{}Y`$ is a trivial fibration in $`𝒞`$, and suppose we have a commutative diagram $$\begin{array}{ccc}A_n_{GA_{n1}}GB_{n1}& & X\\ & & p& & \\ B_n& & Y\end{array}$$ We must construct a lift in this diagram. By adjointness, it suffices to construct a lift in the induced diagram $$\begin{array}{ccc}A& & R_nX\\ i& & & & \\ B& & R_nY\times _{R_{n1}UY}R_{n1}UX\end{array}$$ where $`R_n`$ is the right adjoint of $`\mathrm{Ev}_n`$. Using the description of $`R_n`$ given in Remark 1.4, one can easily check that the map $`R_nX\stackrel{}{}R_nY\times _{R_{n1}UY}R_{n1}UX`$ is a level trivial fibration, so a lift exists. Conversely, suppose that the map $`i`$ satisfies the conditions in the statement of the proposition. Suppose $`p:X\stackrel{}{}Y`$ is a level trivial fibration in $`Sp^{}(𝒞,G)`$, and suppose the diagram $$\begin{array}{ccc}A& \stackrel{f}{}& X\\ i& & p& & \\ B& \underset{g}{}& Y\end{array}$$ commutes. We construct a lift $`h_n:B_n\stackrel{}{}X_n`$, compatible with the structure maps, by induction on $`n`$. There is no difficulty defining $`h_0`$, since $`i_0`$ has the left lifting property with respect to the trivial fibration $`p_0`$. Suppose we have defined $`h_j`$ for $`j<n`$. Then by lifting in the induced diagram $$\begin{array}{ccc}A_n_{GA_{n1}}GB_{n1}& \stackrel{(f_n,\sigma Gh_{n1})}{}& X_n\\ & & p_n& & \\ B_n& \underset{g_n}{}& Y_n\end{array}$$ we find the required map $`h_n:B_n\stackrel{}{}X_n`$. ∎ Finally, we point out that the prolongation of $`G`$ is still a Quillen functor. ###### Proposition 1.16. Give $`Sp^{}(𝒞,G)`$ the projective model structure. Then the prolongation $`G:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$ of $`G`$ is a Quillen functor. Furthermore, the functor $`F_n:𝒞\stackrel{}{}Sp^{}(𝒞,G)`$ is a Quillen functor. ###### Proof. The functor $`\mathrm{Ev}_n`$ obviously takes level fibrations to fibrations and level trivial fibrations to trivial fibrations. Hence $`\mathrm{Ev}_n`$ is a right Quillen functor, and so its left adjoint $`F_n`$ is a left Quillen functor. Similarly, the prolongation of $`U`$ to a functor $`U:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$ preserves level fibrations and level trivial fibrations, so its left adjoint $`G`$ is a Quillen functor. ∎ ## 2. Bousfield localization We will define the stable model structure on $`Sp^{}(𝒞,G)`$ in Section 3 as a Bousfield localization of the projective model structure on $`Sp^{}(𝒞,G)`$. In this section we recall the theory of Bousfield localization of model categories from . To do so, we need some preliminary remarks related to function complexes. Details can be found in \[10, Chapter 5\], , and \[9, Chapter 19\]. Given an object $`A`$ in a model category $`𝒞`$, there is a functorial cosimplicial resolution of $`A`$ induced by the functorial factorizations of $`𝒞`$. By mapping out of this cosimplicial resolution we get a simplicial set $`\mathrm{Map}_{\mathrm{}}(A,X)`$. Similarly, there is a functorial simplicial resolution of $`X`$, and by mapping into it we get a simplicial set $`\mathrm{Map}_r(A,X)`$. One should think of these as replacements for the simplicial structure present in a simplicial model category. Let us define the homotopy function complex $`\mathrm{map}(A,X)=\mathrm{Map}_r(QA,RX)`$. Then $`\mathrm{map}(A,X)`$ is canonically isomorphic in the homotopy category of simplicial sets to $`\mathrm{Map}_{\mathrm{}}(QA,RX)`$, and makes $`\mathrm{Ho}𝒞`$ an enriched category over the homotopy category $`\mathrm{Ho}\mathrm{𝐒𝐒𝐞𝐭}`$ of simplicial sets. In fact, $`\mathrm{Ho}𝒞`$ is naturally tensored and cotensored over $`\mathrm{Ho}\mathrm{𝐒𝐒𝐞𝐭}`$, as well as enriched over it. In particular, if $`G`$ is an arbitrary left Quillen functor between model categories with right adjoint $`U`$, we have $`\mathrm{map}((LG)X,Y)\mathrm{map}(X,(RU)Y)`$ in $`\mathrm{Ho}\mathrm{𝐒𝐒𝐞𝐭}`$, where $`(LG)X=GQX`$ is the total left derived functor of $`G`$ and $`(RU)Y=URY`$ is the total right derived functor of $`U`$. ###### Definition 2.1. Suppose we have a set $`𝒮`$ of maps in a model category $`𝒞`$. 1. A *$`𝒮`$-local object* of $`𝒞`$ is a fibrant object $`W`$ such that, for every $`f:A\stackrel{}{}B`$ in $`𝒮`$, the induced map $`\mathrm{map}(B,W)\stackrel{}{}\mathrm{map}(A,W)`$ is a weak equivalence of simplicial sets. 2. A *$`𝒮`$-local equivalence* is a map $`g:A\stackrel{}{}B`$ in $`𝒞`$ such that the induced map $`\mathrm{map}(B,W)\stackrel{}{}\mathrm{map}(A,W)`$ is a weak equivalence of simplicial sets for all $`𝒮`$-local objects $`W`$. Note that $`𝒮`$-local equivalences between $`𝒮`$-local objects are in fact weak equivalences. Then the main theorem of is the following. We will define cellular model categories, a special class of cofibrantly generated model categories, in the Appendix. ###### Theorem 2.2. Suppose $`𝒮`$ is a set of maps in a left proper cellular model category $`𝒞`$. Then there is a left proper cellular model structure on $`𝒞`$ where the weak equivalences are the $`𝒮`$-local equivalences and the cofibrations remain unchanged. The $`𝒮`$-local objects are the fibrant objects in this model structure. We denote this new model category by $`L_𝒮𝒞`$ and refer to it as the *Bousfield localization* of $`𝒞`$ with respect to $`𝒮`$. Left Quillen functors from $`L_𝒮𝒞`$ to $`𝒟`$ are in one to one correspondence with left Quillen functors $`F:𝒞\stackrel{}{}𝒟`$ such that $`F(Qf)`$ is a weak equivalence for all $`f𝒮`$. We will also need the following fact about localizations, which is implicit in \[9, Chapter 4\]. ###### Proposition 2.3. Suppose $`𝒞`$ and $`𝒟`$ are left proper cellular model categories, $`𝒮`$ is a set of maps in $`𝒞`$, and $`𝒯`$ is a set of maps in $`𝒟`$. Suppose $`F:𝒞\stackrel{}{}𝒟`$ is a Quillen equivalence with right adjoint $`U`$, and suppose $`F(Qf)`$ is a $`𝒯`$-local equivalence for all $`f𝒮`$. Then $`F`$ induces a Quillen *equivalence* $`F:L_𝒮𝒞\stackrel{}{}L_𝒯𝒟`$ if and only if, for every $`𝒮`$-local $`X𝒞`$, there is a $`𝒯`$-local $`Y`$ in $`𝒟`$ such that $`X`$ is weakly equivalent in $`𝒞`$ to $`UY`$. This condition will hold if, for all fibrant $`Y`$ in $`𝒟`$ such that $`UY`$ is $`𝒮`$-local, $`Y`$ is $`𝒯`$-local. ###### Proof. Suppose first that $`F`$ does induce a Quillen equivalence on the localizations, and suppose that $`X`$ is $`𝒮`$-local. Then $`QX`$ is also $`𝒮`$-local. Let $`L_𝒯`$ denote a fibrant replacement functor in $`L_𝒯𝒟`$. Then, because $`F`$ is a Quillen equivalence on the localizations, the map $`QX\stackrel{}{}UL_𝒯FQX`$ is a weak equivalence in $`L_𝒮𝒞`$ (see \[10, Section 1.3.3\]). But both $`QX`$ and $`UL_𝒯FQX`$ are $`𝒮`$-local, so $`QX\stackrel{}{}UL_𝒯FQX`$ is a weak equivalence in $`𝒞`$. Hence $`X`$ is weakly equivalent in $`𝒞`$ to $`UY`$, where $`Y`$ is the $`𝒯`$-local object $`L_𝒯FQX`$. Before proving the converse, note that, since $`F`$ is a Quillen equivalence before localizing, the map $`FQUX\stackrel{}{}X`$ is a weak equivalence for all fibrant $`X`$. Since the functor $`Q`$ does not change upon localization, and every $`𝒯`$-local object of $`𝒟`$ is in particular fibrant, this condition still holds after localization. Thus $`F`$ is a Quillen equivalence after localization if and only if $`F`$ reflects local equivalences between cofibrant objects, by \[10, Corollary 1.3.16\]. Suppose $`f:A\stackrel{}{}B`$ is a map between cofibrant objects such that $`Ff`$ is a $`𝒯`$-local equivalence. We must show that $`\mathrm{map}(f,X)`$ is a weak equivalence for all $`𝒮`$-local $`X`$. Adjointness implies that $`\mathrm{map}(f,UY)`$ is a weak equivalence for all $`𝒯`$-local $`Y`$, and our condition then guarantees that this is enough to conclude that $`\mathrm{map}(f,X)`$ is a weak equivalence for all $`𝒮`$-local $`X`$. We still need to prove the last statement of the proposition. So suppose $`X`$ is $`𝒮`$-local. Then $`QX`$ is also $`𝒮`$-local, and, in $`𝒞`$, we have a weak equivalence $`QX\stackrel{}{}URFQX`$. Our assumption then guarantees that $`Y=RFQX`$ is $`𝒯`$-local, and $`X`$ is in indeed weakly equivalent to $`UY`$. ∎ The fibrations in $`L_𝒮𝒞`$ are not completely understood \[9, Section 3.6\]. The $`𝒮`$-local fibrations between $`𝒮`$-local fibrant objects are just the usual fibrations. In case both $`𝒞`$ and $`L_𝒮𝒞`$ are right proper, there is a characterization of the $`𝒮`$-local fibrations in terms of homotopy pullbacks analogous to the characterization of stable fibrations of spectra in . However, $`L_𝒮𝒞`$ need not be right proper even if $`𝒞`$ is, as is shown by the example of $`\mathrm{\Gamma }`$-spaces in , where it is also shown that the expected characterization of $`𝒮`$-local fibrations does not hold. ## 3. The stable model structure Our plan now is to apply Bousfield localization to the projective model structure on $`Sp^{}(𝒞,G)`$ to obtain a model structure with respect to which $`G`$ is a Quillen equivalence. In order to do this, we will have to prove that the projective model structure makes $`Sp^{}(𝒞,G)`$ into a cellular model category when $`𝒞`$ is left proper cellular. We will prove this technical result in the appendix. In this section, we will assume that $`Sp^{}(𝒞,G)`$ is cellular, find a good set $`𝒮`$ of maps to form the stable model structure as the $`𝒮`$-localization of the projective model structure, and prove that $`G`$ is a Quillen equivalence with respect to the stable model structure. Just as in symmetric spectra , we want the stable equivalences to be maps which induce isomorphisms on all cohomology theories. Cohomology theories will be represented by the appropriate analogue of $`\mathrm{\Omega }`$-spectra. ###### Definition 3.1. A spectrum $`X`$ is a *$`U`$-spectrum* if $`X`$ is level fibrant and the adjoint $`X_n\stackrel{\stackrel{~}{\sigma }}{}UX_{n+1}`$ of the structure map is a weak equivalence for all $`n`$. Of course, if $`𝒞`$ is the category of pointed simplicial sets or pointed topological spaces, and $`G`$ is the suspension functor, $`U`$-spectra are just $`\mathrm{\Omega }`$-spectra. We will find a set $`𝒮`$ of maps of $`Sp^{}(𝒞,G)`$ such that the $`𝒮`$-local objects are the $`U`$-spectra. To do so, note that if $`\mathrm{map}(A,X_n)\stackrel{}{}\mathrm{map}(A,UX_{n+1})`$ is a weak equivalence of simplicial sets for all cofibrant $`A`$ in $`𝒞`$, then $`X_n\stackrel{}{}UX_{n+1}`$ will be a weak equivalence as required. Since $`𝒞`$ is cofibrantly generated, we should not need all cofibrant $`A`$, but only those $`A`$ related to the generating cofibrations. This is true, but the proof is somewhat technical; the reader might be well-advised to skip the following proof. ###### Proposition 3.2. Suppose $`𝒞`$ is a left proper cofibrantly generated model category with generating cofibrations $`I`$, and $`f:X\stackrel{}{}Y`$ is a map. Then $`f`$ is a weak equivalence if and only if $`\mathrm{map}(C,X)\stackrel{}{}\mathrm{map}(C,Y)`$ is a weak equivalence for all domains and codomains $`C`$ of maps of $`I`$. ###### Proof. The only if half is clear. Since every cofibrant object is a retract of a cell complex (i.e. an object $`A`$ such that the map $`0\stackrel{}{}A`$ is a transfinite composition of pushouts of maps of $`I`$), it suffices to show that $`\mathrm{map}(A,f)`$, or, equivalently, $`\mathrm{Map}_r(A,Rf)`$, is a weak equivalence for all cell complexes $`A`$. Given a cell complex $`A`$, there is an ordinal $`\lambda `$ and a $`\lambda `$-sequence $$0=A_0\stackrel{}{}A_1\stackrel{}{}\mathrm{}\stackrel{}{}A_\beta \stackrel{}{}\mathrm{}$$ with colimit $`A_\lambda =A`$. We will show by transfinite induction on $`\beta `$ that $`\mathrm{Map}_r(A_\beta ,Rf)`$ is a weak equivalence for all $`\beta \lambda `$. Since $`A_0=0`$, getting started is easy. For the successor ordinal case of the induction, suppose $`\mathrm{map}(A_\beta ,f)`$ is a weak equivalence. We have a pushout square $$\begin{array}{ccc}C& & A_\beta \\ g& & i_\beta & & \\ D& & A_{\beta +1}\end{array}$$ where $`g`$ is a map of $`I`$. Factor the composite $`QC\stackrel{}{}C\stackrel{}{}D`$ into a cofibration $`\stackrel{~}{g}:QC\stackrel{}{}\stackrel{~}{D}`$ followed by a trivial fibration $`\stackrel{~}{D}\stackrel{}{}D`$. In the terminology of , $`\stackrel{~}{g}`$ is a cofibrant approximation to $`g`$. By \[9, Proposition 12.3.2\], there is a cofibrant approximation $`\stackrel{~}{i_\beta }:\stackrel{~}{A_\beta }\stackrel{}{}\stackrel{~}{A_{\beta +1}}`$ to $`i_\beta `$ which is a pushout of $`\stackrel{~}{g}`$. This is where we need our model category to be left proper. For any fibrant object $`Z`$, the functor $`\mathrm{Map}_r(,Z)`$ converts colimits to limits and cofibrations to fibrations \[10, Corollary 5.4.4\]. Hence we have two pullback squares of fibrant simplicial sets $$\begin{array}{ccc}\mathrm{Map}_r(\stackrel{~}{A_{\beta +1}},RZ)& & \mathrm{Map}_r(\stackrel{~}{D},RZ)\\ & & & & \\ \mathrm{Map}_r(\stackrel{~}{A_\beta },RZ)& & \mathrm{Map}_r(QC,RZ)\end{array}$$ where $`Z=X`$ and $`Z=Y`$, respectively. Here the vertical maps are fibrations. There is a map from the square with $`Z=X`$ to the square with $`Z=Y`$ induced by $`f`$. By hypothesis, this map is a weak equivalence on every corner except possibly the upper left. But then Dan Kan’s cube lemma (see \[10, Lemma 5.2.6\], where the dual of the version we need is proved, or ) implies that the map on the upper left corner is also a weak equivalence, and hence that $`\mathrm{Map}_r(A_{\beta +1},Rf)`$ is a weak equivalence. Now suppose $`\beta `$ is a limit ordinal and $`\mathrm{Map}_r(A_\gamma ,Rf)`$ is a weak equivalence for all $`\gamma <\beta `$. Then, for $`Z=X`$ or $`Z=Y`$, the simplicial sets $`\mathrm{Map}_r(A_\gamma ,RZ)`$ define a limit-preserving functor $`\beta ^{\text{op}}\stackrel{}{}\mathrm{𝐒𝐒𝐞𝐭}`$ such that each map $`\mathrm{Map}_r(A_{\gamma +1},RZ)\stackrel{}{}\mathrm{Map}_r(A_\gamma ,RZ)`$ is a fibration of fibrant simplicial sets. There is a natural transformation from the functor with $`Z=X`$ to the functor with $`Z=Y`$, and by hypothesis this map is a weak equivalence at every stage. It follows that it is a weak equivalence on the inverse limits, as one can see in different ways. The simplest is probably to note that the inverse limit is a right Quillen functor \[10, Corollary 5.1.6\]. Thus $`\mathrm{Map}_r(A_\beta ,Rf)`$ is a weak equivalence, as required. This completes the transfinite induction and the proof. ∎ Note that the left properness assumption in Proposition 3.2 is unnecessary when the domains of the generating cofibrations are themselves cofibrant, since there is then no need to apply cofibrant approximation. In view of Proposition 3.2, we need to choose our set $`𝒮`$ so as to make $$\mathrm{map}(C,X_n)\stackrel{}{}\mathrm{map}(C,UX_{n+1})$$ a weak equivalence for all $`𝒮`$-local objects $`X`$ and all domains and codomains $`C`$ of the generating cofibrations $`I`$. Adjointness implies that, if $`X`$ is level fibrant, $`\mathrm{map}(C,X_n)\mathrm{map}(F_nQC,X)`$ in $`\mathrm{Ho}\mathrm{𝐒𝐒𝐞𝐭}`$, since $`F_nQC=(LF_n)C`$, where $`LF_n`$ is the total left derived functor of $`F_n`$. Also, $`\mathrm{map}(C,UX_{n+1})\mathrm{map}(F_{n+1}GQC,X)`$. In view of this, we make the following definition. ###### Definition 3.3. Suppose $`𝒞`$ is a left proper cellular model category with generating cofibrations $`I`$, and $`G`$ is a left Quillen endofunctor of $`𝒞`$. Define the set $`𝒮`$ of maps in $`Sp^{}(𝒞,G)`$ as $`\{F_{n+1}GQC\stackrel{s_n^{QC}}{}F_nQC\}`$, as $`C`$ runs through the set of domains and codomains of the maps of $`I`$ and $`n`$ runs through the non-negative integers. Here the map $`s_n^{QC}`$ is adjoint to the identity map of $`GQC`$. Define the *stable model structure* on $`Sp^{}(𝒞,G)`$ to be the localization of the projective model structure on $`Sp^{}(𝒞,G)`$ with respect to this set $`𝒮`$. We refer to the $`𝒮`$-local weak equivalences as *stable equivalences*, and to the $`𝒮`$-local fibrations as *stable fibrations*. ###### Theorem 3.4. Suppose $`𝒞`$ is a left proper cellular model category and $`G`$ is a left Quillen endofunctor of $`𝒞`$. Then the stably fibrant objects in $`Sp^{}(𝒞,G)`$ are the $`U`$-spectra. Furthermore, for all cofibrant $`A𝒞`$ and for all $`n0`$, the map $`F_{n+1}GA\stackrel{s_n^A}{}F_nA`$ is a stable equivalence. ###### Proof. By definition, $`X`$ is $`𝒮`$-local if and only if $`X`$ is level fibrant and $$\mathrm{map}(F_nQC,X)\stackrel{}{}\mathrm{map}(F_{n+1}GQC,X)$$ is a weak equivalence for all $`n0`$ and all domains and codomains $`C`$ of maps of $`I`$. By the comments preceding Definition 3.3, this is equivalent to requiring that $`X`$ be level fibrant and that the map $`\mathrm{map}(C,X_n)\stackrel{}{}\mathrm{map}(C,UX_{n+1})`$ be a weak equivalence for all $`n0`$ and all domains and codomains $`C`$ of maps of $`I`$. By Proposition 3.2, this is equivalent to requiring that $`X`$ be a $`U`$-spectrum. Now, by definition, $`s_n^A`$ is a stable equivalence if and only if $`\mathrm{map}(s_n^A,X)`$ is a weak equivalence for all $`U`$-spectra $`X`$. But by adjointness, $`\mathrm{map}(s_n^A,X)`$ can be identified with $`\mathrm{map}(A,X_n)\stackrel{}{}\mathrm{map}(A,UX_{n+1})`$. Since $`X_n\stackrel{}{}UX_{n+1}`$ is a weak equivalence between fibrant objects, so is $`\mathrm{map}(s_n^A,X)`$. ∎ We would now like to claim that the stable model structure on $`Sp^{}(𝒞,G)`$ that we have just defined is a generalization of the stable model structure on spectra of topological spaces or simplicial sets defined in . This cannot be a trivial observation, however, both because our approach is totally different and because of Remark 1.7. ###### Corollary 3.5. If $`𝒞`$ is either the category of pointed simplicial sets or pointed topological spaces, and $`G`$ is the suspension functor given by smashing with $`S^1`$, then the stable model structure on $`Sp^{}(𝒞,G)`$ coincides with the stable model structure on the category of Bousfield-Friedlander spectra . ###### Proof. We know already that the cofibrations are the same in the stable model structure on $`Sp^{}(𝒞,G)`$ and the stable model structure of . We will show that the weak equivalences are the same. In any model category at all, a map $`f`$ is a weak equivalence if and only if $`\mathrm{map}(f,X)`$ is a weak equivalence of simplicial sets for all fibrant $`X`$. Construction of $`\mathrm{map}(f,X)`$ requires replacing $`f`$ by a cofibrant approximation $`f^{}`$ and building cosimplicial resolutions of the domain and codomain of $`f^{}`$. In the case at hand, we can do the cofibrant replacement and build the cosimplicial resolutions in the strict model category of spectra, since the cofibrations do not change under localization. Thus $`\mathrm{map}(f,X)`$ is the same in both the stable model structure on $`Sp^{}(𝒞,G)`$ and in the stable model category of Bousfield and Friedlander. Since the stably fibrant objects are also the same, the corollary holds. ∎ The purpose of the stable model structure is to make the prolongation of $`G`$ into a Quillen equivalence. We begin the process of proving this with the following corollary. ###### Corollary 3.6. Suppose $`𝒞`$ is a left proper cellular model category and $`G`$ is a left Quillen endofunctor of $`𝒞`$. Then the prolongation of $`G`$ to a functor $`G:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$ is a Quillen functor with respect to the stable model structure. ###### Proof. In view of Hirschhorn’s localization theorem 2.2, we must show that $`G(Qf)`$ is a stable equivalence for all $`f𝒮`$. Since the domains and codomains of the maps of $`𝒮`$ are already cofibrant, it is equivalent to show that $`Gf`$ is a stable equivalence for all $`fS`$. Since $`GF_n=F_nG`$, we have $`Gs_n^A=s_n^{GA}`$. In view of Theorem 3.4, this map is a weak equivalence whenever $`A`$, and hence $`GA`$, is cofibrant. Taking $`A=QC`$, where $`C`$ is a domain or codomain of a map of $`I`$, completes the proof. ∎ We will now show that $`G`$ is in fact a Quillen *equivalence* with respect to the stable model structure. To do so, we introduce the shift functors. ###### Definition 3.7. Suppose $`𝒞`$ is a model category and $`G`$ is a left Quillen endofunctor of $`𝒞`$. Define the *shift functors* $`t:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$ and $`s:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$ by $`(sX)_n=X_{n+1}`$ and $`(tX)_n=X_{n1}`$, $`(tX)_0=0`$, with the evident structure maps. Note that $`t`$ is left adjoint to $`s`$. It is clear that $`s`$ preserves level equivalences and level fibrations, so $`t`$ is a left Quillen functor with respect to the strict model structure on $`Sp^{}(𝒞,G)`$, and $`s`$ is a right Quillen functor. Also, $`sU=Us`$, so $`tG=Gt`$. Similarly, $`\mathrm{Ev}_ns=\mathrm{Ev}_{n+1}`$, so $`tF_n=F_{n+1}`$. It follows that $`ts_n^A=s_{n+1}^A`$, so that $`t`$ is a Quillen functor with respect to the stable model structure as well. We have now come to the main advantage of our approach to spectra. We can test whether a spectrum $`X`$ is a $`U`$-spectrum by checking that $`X`$ is level fibrant and by checking that the map $`X\stackrel{}{}sUX`$, adjoint to the structure map of $`X`$, is a level equivalence. The analogous statement is false in the Bousfield-Friedlander category , because the extra twist map they use (see Remark 1.7) means that there is no map of spectra $`X\stackrel{}{}sUX`$ adjoint to the structure map of $`X`$! Our interpretation of this is that the Bousfield-Friedlander approach, while excellent at what it does, is probably not the right general construction. Further evidence for this is provided by the extreme simplicity of the following proof. ###### Theorem 3.8. Suppose $`𝒞`$ is a left proper cellular model category and $`G`$ is a left Quillen endofunctor of $`𝒞`$. Then the functors $`G:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$ and $`t:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$ are Quillen equivalences with respect to the stable model structures. Furthermore, $`Rs`$ is naturally isomorphic to $`LG`$, and $`RU`$ is naturally isomorphic to $`Lt`$. ###### Proof. There is a a natural map $`X\stackrel{}{}sUX`$ which is a weak equivalence when $`X`$ is a stably fibrant object of $`Sp^{}(𝒞,G)`$. This means that the total right derived functor $`R(sU)`$ is naturally isomorphic to the identity functor on $`\mathrm{Ho}Sp^{}(𝒞,G)`$ (where we use the stable model structure). On the other hand, $`R(sU)`$ is naturally isomorphic to $`RsRU`$ and also to $`RURs`$, since $`s`$ and $`U`$ commute with each other. Thus the natural isomorphism from the identity to $`R(sU)`$ gives rise to an natural isomorphism $`1\stackrel{}{}RsRU`$ and a natural isomorphism $`RURs\stackrel{}{}1`$, displaying $`RU`$ and $`Rs`$ as adjoint equivalences of categories. It follows that $`U`$ and $`s`$ are both Quillen equivalences, as required. Since $`RU`$ is adjoint to $`LG`$ and $`Rs`$ is adjoint to $`Lt`$, we must also have $`RU`$ naturally isomorphic to $`Lt`$ and $`Rs`$ naturally isomorphic to $`LG`$. ∎ ## 4. The almost finitely generated case The reader may well object at this point that we have defined the stable model structure on $`Sp^{}(𝒞,G)`$ without ever defining stable homotopy groups. This is because stable homotopy groups do not detect stable equivalences in general. The usual simplicial and topological situation is very special. The goal of this section is to put some hypotheses on $`𝒞`$ and $`G`$ so that the stable model structure on $`Sp^{}(𝒞,G)`$ behaves similarly to the stable model structure on ordinary simplicial spectra. In particular, we show that, if $`𝒞`$ is almost finitely generated (defined below), the usual $`Q`$ construction gives a stable fibrant replacement functor; thus, a map $`f`$ is a stable equivalence if and only if $`Qf`$ is a level equivalence. This allows us to characterize $`\mathrm{Ho}Sp^{}(𝒞,G)(F_0A,X)`$ for well-behaved $`A`$ as the usual sort of colimit $`\mathrm{colim}\mathrm{Ho}𝒞(G^nA,X_n)`$. It also allows us to prove that the stable model structure is right proper, so we get the expected characterization of stable fibrations. Most of the results in this section do not depend on the existence of the stable model structure on $`Sp^{}(𝒞,G)`$, so we do not usually need to assume $`𝒞`$ is left proper cellular. We now define almost finitely generated model categories, as suggested to the author by Voevodsky. ###### Definition 4.1. An object $`A`$ of a category $`𝒞`$ is called *finitely presented* if the functor $`𝒞(A,)`$ preserves direct limits of sequences $`X_0\stackrel{}{}X_1\stackrel{}{}\mathrm{}\stackrel{}{}X_n\stackrel{}{}\mathrm{}`$. A cofibrantly generated model category $`𝒞`$ is said to be *finitely generated* if the domains and codomains of the generating cofibrations and the generating trivial cofibrations are finitely presented. A cofibrantly generated model category is said to be *almost finitely generated* if the domains and codomains of the generating cofibrations are finitely presented, and if there is a set of trivial cofibrations $`J^{}`$ with finitely presented domains and codomains such that a map $`f`$ *whose codomain is fibrant* is a fibration if and only if $`f`$ has the right lifting property with respect to $`J^{}`$. This definition differs slightly from other definitions. In particular, an object $`A`$ is usually said to be finitely presented if $`𝒞(A,)`$ preserves all directed (or, equivalently, filtered) colimits. We are trying to assume the minimum necessary. Finitely generated model categories were introduced in \[10, Section 7.4\], but in that definition we assumed only that $`𝒞(A,)`$ preserves (transfinitely long) direct limits of sequences of *cofibrations*. The author would now prefer to call such model categories *compactly generated*. Thus, the model category of simplicial sets is finitely generated, but the model category on topological spaces is only compactly generated. Since we will only be working with (almost) finitely generated model categories in this section, our results will not apply to topological spaces. We will indicate where our results fail for compactly generated model categories, and a possible way to amend them in the compactly generated case. The definition of an almost finitely generated model category was suggested by Voevodsky. The problem with finitely generated, or, indeed, compactly generated, model categories is that they are not preserved by localization. That is, if $`𝒞`$ is a finitely generated left proper cellular model category, and $`S`$ is a set of maps, then the Bousfield localization $`L_S𝒞`$ will not be finitely generated, because we lose all control over the generating trivial cofibrations in $`L_S𝒞`$. However, if $`S`$ is a set of cofibrations such that $`XK`$ is finitely presented for every domain or codomain $`X`$ of a map of $`S`$ and every finite simplicial set $`K`$, then $`L_S𝒞`$ will still be almost finitely generated. (We use a framing on $`𝒞`$ to construct $`XK`$). Indeed, the horns $$(A\mathrm{\Delta }[n])_{A\mathrm{\Lambda }^k[n]}(B\mathrm{\Lambda }^k[n])\stackrel{}{}B\mathrm{\Delta }[n]$$ on the maps $`A\stackrel{}{}B`$ of $`S`$ are used to detect $`S`$-local fibrant objects, and an $`S`$-local fibration between $`S`$-local fibrant objects is just an ordinary fibration. We can therefore take the set $`J^{}`$ to consist of the horns on the maps of $`S`$ together with the old set of generating trivial cofibrations. In particular, Voevodsky has informed the author that he can make an unstable motivic model category that is almost finitely generated, using this approach. For the reader’s benefit, we summarize his construction. The category $`𝒞`$ is the category of simplicial presheaves (of sets) on the category of smooth schemes over some base scheme $`k`$. There is a projective model structure on this category, where weak equivalences and fibrations are defined objectwise from weak equivalences and fibrations of simplicial sets. The projective model structure is finitely generated (using the fact that smooth schemes over $`k`$ is an essentially small category). There is an embedding of smooth schemes into $`𝒞`$ as representable functors. We need to localize this model structure to take into account both the Nisnevich topology and the fact that the functor $`XX\times 𝔸^1`$ should be a Quillen equivalence. To do so, we define a set $`S^{}`$ to consist of the maps $`X\times 𝔸^1\stackrel{}{}X`$ for every smooth scheme $`X`$ and maps $`P\stackrel{}{}X`$ for every pullback square of smooth schemes $$\begin{array}{ccc}B& & Y\\ & & p& & \\ A& \underset{j}{}& X\end{array}$$ where $`p`$ is etale, $`j`$ is an open embedding, and $`p^1(XA)\stackrel{}{}XA`$ is an isomorphism. Here $`P`$ is the mapping cylinder $`(BY)_{BB}(A\times \mathrm{\Delta }[1])`$. We then define $`S`$ to consist of mapping cylinders on the maps of $`S^{}`$. The maps of $`S`$ are then cofibrations whose domains and codomains are finitely presented (and remain so after tensoring with any finite simplicial set), so the Bousfield localization will be almost finitely generated. There is then some work involving properties of the Nisnevich topology to show that this model category is equivalent to the Morel-Voevodsky motivic model category of , and to the model category used by Jardine . The reason we need almost finitely generated model categories is because, in an almost finitely generated model category $`𝒞`$, sequential colimits preserve trivial fibrations, fibrant objects, and fibrations between fibrant objects. Indeed, suppose we have a map of sequences $`p_n:X_n\stackrel{}{}Y_n`$ that is a fibration between fibrant objects for all $`n`$. We show $`\mathrm{colim}X_n`$ is fibrant by testing that $`X\stackrel{}{}0`$ has the right lifting property with respect to $`J^{}`$. We then test that $`\mathrm{colim}p_n`$ is a fibration by testing that it has the right lifting property with respect to $`J^{}`$. The proof that sequential colimits preserve trivial fibrations is similar. Now, given a spectrum $`X`$, there is an obvious candidate for a stable fibrant replacement of $`X`$. ###### Definition 4.2. Suppose $`G`$ is a left Quillen endofunctor of a model category $`𝒞`$ with right adjoint $`U`$. Define $`R:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒞,G)`$ to be the functor $`sU`$, where $`s`$ is the shift functor. Then we have a natural map $`\iota _X:X\stackrel{}{}RX`$, and we define $$R^{\mathrm{}}X=\mathrm{colim}(X\stackrel{\iota _X}{}RX\stackrel{R\iota _X}{}R^2X\stackrel{R^2\iota _X}{}\mathrm{}\stackrel{R^{n1}\iota _X}{}R^nX\stackrel{R^n\iota _X}{}\mathrm{}).$$ Let $`j_X:X\stackrel{}{}R^{\mathrm{}}X`$ denote the obvious natural transformation. The following lemma, though elementary, is crucial. ###### Lemma 4.3. The maps $`\iota _{RX},R\iota _X:RX\stackrel{}{}R^2X`$ coincide. ###### Proof. The map $`\iota _{RX}`$ is the adjoint of the structure map $$GUX_{n+1}\stackrel{𝜀}{}X_{n+1}\stackrel{𝜂}{}UGX_{n+1}\stackrel{U\sigma }{}UX_{n+2}$$ of $`RX`$, where $`\epsilon `$ denotes the counit of the adjunction, $`\eta `$ denotes the unit, and $`\sigma `$ denotes the structure map of $`X`$. Thus $`\iota _{RX}`$ is the composite $$UX_{n+1}\stackrel{𝜂}{}UGUX_{n+1}\stackrel{U\epsilon }{}UX_{n+1}\stackrel{U\eta }{}U^2GX_{n+1}\stackrel{U^2\sigma }{}U^2X_{n+2}.$$ Since $`U\epsilon \eta `$ is the identity, it follows that $`\iota _{RX}=R\iota _X`$. ∎ We stress that Lemma 4.3 fails for symmetric spectra, and it is the major reason we must work with finitely generated model categories rather than compactly generated model categories. Indeed, in the compactly generated case, $`R^{\mathrm{}}`$ is not a good functor, since maps out of one of the domains of the generating cofibrations will not preserve the colimit that defines $`R^{\mathrm{}}`$. The obvious thing to try is to replace the functor $`R`$ by a functor $`W`$, obtained by factoring $`X\stackrel{}{}RX`$ into a projective cofibration $`X\stackrel{}{}WX`$ followed by a level trivial fibration $`WX\stackrel{}{}RX`$. The difficulty with this plan is that we do not see how to prove Lemma 4.3 for $`W`$. An alternative plan would be to use the mapping cylinder $`X\stackrel{}{}W^{}X`$ on $`X\stackrel{}{}RX`$; this might make Lemma 4.3 easier to prove, but the map $`X\stackrel{}{}W^{}X`$ will not be a cofibration. The map $`X\stackrel{}{}W^{}X`$ may, however, be good enough for the required smallness properties to hold. It is a closed inclusion if $`𝒞`$ is topological spaces, for example. The author knows of no good general theorem in the compactly generated case. This lemma leads immediately to the following proposition. ###### Proposition 4.4. Suppose $`G`$ is a left Quillen endofunctor of a model category $`𝒞`$, and suppose that its right adjoint $`U`$ preserves sequential colimits. Then the map $`\iota _{R^{\mathrm{}}X}:R^{\mathrm{}}X\stackrel{}{}R(R^{\mathrm{}}X)`$ is an isomorphism. In particular, if $`X`$ is level fibrant, $`𝒞`$ is almost finitely generated, and $`U`$ preserves sequential colimits, then $`R^{\mathrm{}}X`$ is a $`U`$-spectrum. ###### Proof. The map $`\iota _{R^{\mathrm{}}X}`$ is the colimit of the vertical maps in the diagram below. $$\begin{array}{ccccccccccc}X& \stackrel{\iota _X}{}& RX& \stackrel{R\iota _X}{}& R^2X& \stackrel{R^2\iota _X}{}& \mathrm{}& \stackrel{R^{n1}\iota _X}{}& R^nX& \stackrel{R^n\iota _X}{}& \mathrm{}\\ \iota _X& & \iota _{RX}& & \iota _{R^2X}& & & & \iota _{R^nX}& & \\ RX& \underset{R\iota _X}{}& R^2X& \underset{R^2\iota _X}{}& R^3X& \underset{R^3\iota _X}{}& \mathrm{}& \underset{R^n\iota _X}{}& R^{n+1}X& \underset{R^{n+1}\iota _X}{}& \mathrm{}\end{array}$$ Since the vertical and horizontal maps coincide, the result follows. For the second statement, we note that if $`X`$ is level fibrant, each $`R^nX`$ is level fibrant since $`R`$ is a right Quillen functor (with respect to the projective model structure). Since sequential colimits in $`𝒞`$ preserve fibrant objects, $`R^{\mathrm{}}X`$ is level fibrant, and hence a $`U`$-spectrum. ∎ ###### Proposition 4.5. Suppose $`G`$ is a left Quillen endofunctor of a model category $`𝒞`$ with right adjoint $`U`$. If $`𝒞`$ is almost finitely generated, and $`X`$ is a $`U`$-spectrum, then the map $`j_X:X\stackrel{}{}R^{\mathrm{}}X`$ is a level equivalence. ###### Proof. By assumption, the map $`\iota _X:X\stackrel{}{}RX`$ is a level equivalence between level fibrant objects. Since $`R`$ is a right Quillen functor, $`R^n\iota _X`$ is a level equivalence as well. Then the method of \[10, Corollary 7.4.2\] completes the proof. Recall that this method is to use factorization to construct a sequence of projective trivial cofibrations $`Y_n\stackrel{}{}Y_{n+1}`$ with $`Y_0=X`$ and a level trivial fibration of sequences $`Y_n\stackrel{}{}R^nX`$. Then the map $`X\stackrel{}{}\mathrm{colim}Y_n`$ will be a projective trivial cofibration. Since sequential colimits in $`𝒞`$ preserve trivial fibrations, the map $`\mathrm{colim}Y_n\stackrel{}{}R^{\mathrm{}}X`$ will still be a level trivial fibration. ∎ Proposition 4.5 gives us a slightly better method of detecting stable equivalences. ###### Corollary 4.6. Suppose $`G`$ is a left Quillen endofunctor of a model category $`𝒞`$ with right adjoint $`U`$. Suppose $`𝒞`$ is almost finitely generated and $`U`$ preserves sequential colimits. Then a map $`f:A\stackrel{}{}B`$ is a stable equivalence in $`Sp^{}(𝒞,G)`$ if and only if $`\mathrm{map}(f,X)`$ is a weak equivalence for all level fibrant spectra $`X`$ such that $`\iota _X:X\stackrel{}{}RX`$ is an isomorphism. ###### Proof. By definition, $`f`$ is a stable equivalence if and only if $`\mathrm{map}(f,Y)`$ is a weak equivalence for all $`U`$-spectra $`Y`$. But we have a level equivalence $`Y\stackrel{}{}R^{\mathrm{}}Y`$ by Proposition 4.5, and so it suffices to know that $`\mathrm{map}(f,R^{\mathrm{}}Y)`$ is a weak equivalence for all $`U`$-spectra $`Y`$. But, by Proposition 4.4, $`\iota _{R^{\mathrm{}}Y}`$ is an isomorphism. ∎ This corollary, in turn, allows us to prove that $`R^{\mathrm{}}`$ detects stable equivalences. The following theorem is similar to \[11, Theorem 3.1.11\]. ###### Theorem 4.7. Suppose $`G`$ is a left Quillen endofunctor of a model category $`𝒞`$ with right adjoint $`U`$. Suppose that $`𝒞`$ is almost finitely generated and sequential colimits in $`𝒞`$ preserve finite products. Suppose also that $`U`$ preserves sequential colimits. If $`f:A\stackrel{}{}B`$ is a map in $`Sp^{}(𝒞,G)`$ such that $`R^{\mathrm{}}f`$ is a level equivalence, then $`f`$ is a stable equivalence. ###### Proof. Suppose $`X`$ is a $`U`$-spectrum such that the map $`\iota _X:X\stackrel{}{}RX`$ is an isomorphism. We will show that $`\mathrm{map}(f,X)`$ as a retract of $`\mathrm{map}(R^{\mathrm{}}f,R^{\mathrm{}}X)`$; this will obviously complete the proof. We first note that there is a natural map $`\mathrm{map}(R^{\mathrm{}}C,R^{\mathrm{}}X)\stackrel{}{}\mathrm{map}(C,X)`$ obtained by precomposition with $`C\stackrel{}{}R^{\mathrm{}}C`$ and postcomposition with $`k:R^{\mathrm{}}X\stackrel{}{}X`$. Here $`k`$ is the inverse of the map $`j_X:X\stackrel{}{}R^{\mathrm{}}X`$, which is an isomorphism since $`\iota _X`$ is so. On the other hand, we claim that there is also a natural map $`\mathrm{map}(C,X)\stackrel{}{}\mathrm{map}(R^{\mathrm{}}C,R^{\mathrm{}}X)`$ obtained by applying the total right derived functor of $`R^{\mathrm{}}`$. This is not obvious, since we are asserting that the total right derived functor of $`R^{\mathrm{}}`$ preserves the enrichment of the projective homotopy category of $`Sp^{}(𝒞,G)`$ over the homotopy category of simplicial sets, even though $`R^{\mathrm{}}`$ is not a right Quillen functor. Nevertheless, if we assume that this natural map exists, it follows easily that the composite $`\mathrm{map}(C,X)\stackrel{}{}\mathrm{map}(R^{\mathrm{}}C,R^{\mathrm{}}X)\stackrel{}{}\mathrm{map}(C,X)`$ is the identity (in the homotopy category of simplicial sets), and therefore that $`\mathrm{map}(f,X)`$ is a retract of $`\mathrm{map}(R^{\mathrm{}}f,R^{\mathrm{}}X)`$. It remains to show that the total right derived functor of $`R^{\mathrm{}}`$ preserves the enriched structure. We first point out that $`R^{\mathrm{}}`$ preserves level fibrations between level fibrant objects and all level trivial fibrations, because $`𝒞`$ is almost finitely generated. It follows from Ken Brown’s lemma \[10, Lemma 1.1.12\] that $`R^{\mathrm{}}`$ preserves level equivalences between level fibrant objects. Because sequential colimits in $`𝒞`$ preserve finite products, $`R^{\mathrm{}}`$ also preserves finite products. We claim that, for any functor $`H`$ on a model category $`𝒟`$ that preserves fibrations between fibrant objects, weak equivalences between fibrant objects, and finite products, the total right derived functor of $`H`$ preserves the enriched structure over $`\mathrm{Ho}\mathrm{𝐒𝐒𝐞𝐭}`$. Indeed, analysis of the definition of this enrichment \[10, Chapter 5\] shows that it suffices to check that such a functor $`H`$ preserves simplicial frames on a fibrant object $`Y`$. A simplicial frame on $`Y`$ is a factorization $`\mathrm{}_{}Y\stackrel{𝛼}{}Y_{}\stackrel{𝛽}{}r_{}Y`$ in the diagram category of simplicial objects in $`𝒟`$, where $`\mathrm{}_{}Y`$ is the constant simplicial diagram on $`Y`$, the $`n`$th space of $`r_{}Y`$ is the product $`Y^{n+1}`$, $`\alpha `$ is a level equivalence, and $`\beta `$ is a level fibration. We further require that $`\beta `$ is an isomorphism in degree $`0`$. Since $`H`$ preserves products, weak equivalences between fibrant objects, and fibrations between fibrant objects, it follows that $`HY_{}`$ is a simplicial frame on $`HY`$. ∎ ###### Corollary 4.8. Suppose $`G`$ is a left Quillen endofunctor of a model category $`𝒞`$ with right adjoint $`U`$. Suppose that $`𝒞`$ is almost finitely generated and sequential colimits in $`𝒞`$ preserve finite products. Suppose also that $`U`$ preserves sequential colimits. Then $`j_A:A\stackrel{}{}R^{\mathrm{}}A`$ is a stable equivalence for all $`ASp^{}(𝒞,G)`$. ###### Proof. One can easily check that $`R^{\mathrm{}}j_A`$ is an isomorphism, using Proposition 4.4. ∎ Finally, we get the desired characterization of stable equivalences. ###### Theorem 4.9. Suppose $`G`$ is a left Quillen endofunctor of a model category $`𝒞`$ with right adjoint $`U`$. Suppose that $`𝒞`$ is almost finitely generated and sequential colimits in $`𝒞`$ preserve finite products. Suppose as well that $`U`$ preserves sequential colimits. Let $`L^{}`$ denote a fibrant replacement functor in the projective model structure on $`Sp^{}(𝒞,G)`$. Then, for all $`ASp^{}(𝒞,G)`$, the map $`A\stackrel{}{}R^{\mathrm{}}L^{}A`$ is a stable equivalence into a $`U`$-spectrum. Also, a map $`f:A\stackrel{}{}B`$ is a stable equivalence if and only if $`R^{\mathrm{}}L^{}f`$ is a level equivalence. ###### Proof. The first statement follows immediately from Proposition 4.4 and Corollary 4.8. By the first statement, if $`f`$ is a stable equivalence, so is $`R^{\mathrm{}}L^{}f`$. Since $`R^{\mathrm{}}L^{}f`$ is a map between $`U`$-spectra, it is a stable equivalence if and only if it is a level equivalence. The converse follows from Theorem 4.7. ∎ Since we did not need the existence of the stable model structure to prove Theorem 4.9, one can imagine attempting to construct it from the functor $`R^{\mathrm{}}L^{}`$. This is, of course, the original approach of Bousfield-Friedlander , and this approach has been generalized by Schwede . Also, if one has some way to detect level equivalences in $`𝒞`$, say using appropriate generalizations of homotopy groups, Theorem 4.9 implies that stable equivalences in $`Sp^{}(𝒞,G)`$ are detected by the appropriate generalizations of stable homotopy groups. One can see these generalizations in the following corollary as well. ###### Corollary 4.10. Suppose $`𝒞`$ is a pointed, left proper, cellular, almost finitely generated model category where sequential colimits preserve finite products. Suppose $`G:𝒞\stackrel{}{}𝒞`$ is a left Quillen functor whose right adjoint $`U`$ commutes with sequential colimits. Finally, suppose $`A`$ is a finitely presented cofibrant object of $`𝒞`$ that has a finitely presented cylinder object $`A\times I`$. Then $$\mathrm{Ho}Sp^{}(𝒞,G)(F_kA,Y)=\mathrm{colim}_m\mathrm{Ho}𝒞(A,U^mY_{k+m}).$$ for all level fibrant $`YSp^{}(𝒞,G)`$. Here we are using the stable model structure to form $`\mathrm{Ho}Sp^{}(𝒞,G)`$, of course. ###### Proof. We have $`\mathrm{Ho}Sp^{}(𝒞,G)(F_kA,Y)=Sp^{}(𝒞,G)(F_kA,R^{\mathrm{}}Y)/`$, by Theorem 4.9, where $``$ denotes the left homotopy relation. We can use the cylinder object $`F_k(A\times I)`$ as the source for our left homotopies. Then adjointness implies that $`Sp^{}(𝒞,G)(F_kA,R^{\mathrm{}}Y)/=𝒞(A,\mathrm{Ev}_kR^{\mathrm{}}Y)/`$. Since $`A`$ and $`A\times I`$ are finitely presented, we get the required result. ∎ By assuming slightly more about $`𝒞`$, we can also characterize the stable fibrations. ###### Corollary 4.11. Suppose $`𝒞`$ is a pointed, proper, cellular, almost finitely generated model category such that sequential colimits preserve pullbacks. Suppose $`G:𝒞\stackrel{}{}𝒞`$ is a left Quillen functor whose right adjoint $`U`$ commutes with sequential colimits. Then the stable model structure on $`Sp^{}(𝒞,G)`$ is proper. In particular, a map $`f:X\stackrel{}{}Y`$ is a stable fibration if and only if $`f`$ is a level fibration and the diagram $$\begin{array}{ccc}X& & R^{\mathrm{}}L^{}X\\ f& & Lf& & \\ Y& & R^{\mathrm{}}L^{}Y\end{array}$$ is a homotopy pullback square in the projective model structure, where $`L^{}`$ is a fibrant replacement functor in the projective model structure. ###### Proof. We wil actually show that, if $`p:X\stackrel{}{}Y`$ is a level fibration and $`f:B\stackrel{}{}Y`$ is a stable equivalence, the pullback $`B\times _AY\stackrel{}{}X`$ is a stable equivalence. The first step is to use the right properness of the projective model structure on $`Sp^{}(𝒞,G)`$ to reduce to the case where $`B`$ and $`Y`$ are level fibrant. Indeed, let $`Y^{}=L^{}Y`$, $`B^{}=L^{}B`$, and $`f^{}=L^{}f`$. Then factor the composite $`X\stackrel{}{}Y\stackrel{}{}Y^{}`$ into a projective trivial cofibration $`X^{}\stackrel{}{}Y^{}`$ followed by a level fibration $`p^{}:X^{}\stackrel{}{}Y^{}`$. Then we have the commutative diagram below, $$\begin{array}{ccccc}B& \stackrel{f}{}& Y& \stackrel{p}{}& X\\ & & & & & & \\ B^{}& \underset{f^{}}{}& Y^{}& \underset{p^{\prime \prime }}{}& X^{}\end{array}$$ where the vertical maps are level equivalences. Then Proposition 12.2.4 and Corollary 12.2.8 of , which depend on the projective model structure being right proper, imply that the induced map $`B\times _YX\stackrel{}{}B^{}\times _Y^{}X^{}`$ is a level equivalence. Hence $`B\times _YX\stackrel{}{}X`$ is a stable equivalence if and only if $`B^{}\times _Y^{}X^{}\stackrel{}{}X^{}`$ is a stable equivalence, and so we can assume $`B`$ and $`X`$ are level fibrant. Now let $`S`$ denote the pullback square below. $$\begin{array}{ccc}B\times _YX& & X\\ & & p& & \\ B& \underset{f}{}& Y\end{array}$$ Then $`R^nS`$ is a pullback square for all $`n`$, and there are maps $`R^nS\stackrel{R^n\iota _S}{}R^{n+1}S`$. Since pullbacks commute with sequential colimits, $`R^{\mathrm{}}S`$ is a pullback square. Furthermore, $`R^{\mathrm{}}p`$ is a level fibration, since sequential colimits in $`𝒞`$ preserve fibrations between level fibrant objects. Since $`f`$ is a stable equivalence between level fibrant spectra, $`R^{\mathrm{}}f`$ is a level equivalence by Theorem 4.9. So, since the projective model structure is right proper, the map $`R^{\mathrm{}}(B\times _YX\stackrel{}{}X)`$ is a level equivalence, and thus $`B\times _YX\stackrel{}{}X`$ is a stable equivalence. The characterization of stable fibrations then follows from \[9, Proposition 3.6.8\]. ∎ ## 5. Properties of the stabilization functor In this section we explore some of the properties of the correspondence $`(𝒞,G)Sp^{}(𝒞,G)`$, where, throughout this section, we mean the stable model structure on $`Sp^{}(𝒞,G)`$. We begin by showing that, if $`G`$ is already a Quillen equivalence, then the embedding $`𝒞\stackrel{}{}Sp^{}(𝒞,G)`$ is a Quillen equivalence. This important fact is as close as we can get to proving that $`Sp^{}(𝒞,G)`$ is the initial, up to homotopy, stabilization of $`𝒞`$ with respect to $`G`$. We also show that $`Sp^{}(𝒞,G)`$ is functorial in the pair $`(𝒞,G)`$, with a suitable definition of maps of pairs. Under mild hypotheses, we show that $`Sp^{}(𝒞,G)`$ preserves Quillen equivalences in the pair $`(𝒞,G)`$. In particular, this means, for example, that the Bousfield-Friedlander category of spectra of simplicial sets does not depend, up to Quillen equivalence, on which model of the circle $`S^1`$ one chooses. We conclude the section by pointing out that our stabilization condition preserves some monoidal structure. For example, if $`𝒞`$ is a simplicial model category, and $`G`$ is a simplicial functor, then $`Sp^{}(𝒞,G)`$ is a gain a simplicial model category, and the extension of $`G`$ is again a simplicial functor. However, if $`𝒞`$ is monoidal, and $`G`$ is a monoidal functor, $`Sp^{}(𝒞,G)`$ will almost never be a monoidal category; this is the reason we need the symmetric spectra introduced in the next section. ###### Theorem 5.1. Suppose $`𝒞`$ is a left proper cellular model category, and suppose $`G`$ is a left Quillen endofunctor of $`𝒞`$ that is a Quillen equivalence. Then $`F_0:𝒞\stackrel{}{}Sp^{}(𝒞,G)`$ is a Quillen equivalence, where $`Sp^{}(𝒞,G)`$ has the stable model structure. ###### Proof. We first point out that the right adjoint $`\mathrm{Ev}_0:Sp^{}(𝒞,G)\stackrel{}{}𝒞`$ reflects weak equivalences between fibrant objects. Indeed, suppose $`X`$ and $`Y`$ are $`U`$-spectra, and $`f:X\stackrel{}{}Y`$ is a map such that $`\mathrm{Ev}_0f=f_0`$ is a weak equivalence. Then, because $`X`$ and $`Y`$ are $`U`$-spectra, $`U^nf_n`$ is a weak equivalence for all $`n`$. Since $`G`$ is a Quillen equivalence, $`U`$ reflects weak equivalences between fibrant objects by \[10, Corollary 1.3.16\]. Thus $`f_n`$ is a weak equivalence for all $`n`$, and so $`f`$ is a level equivalence and hence a stable equivalence, as required. In view of \[10, Corollary 1.3.16\], to complete the proof it suffices to show that $`X\stackrel{}{}(L_𝒮F_0X)_0`$ is a weak equivalence for all cofibrant $`X𝒞`$, where $`L_𝒮`$ denotes a stably fibrant replacement functor in $`Sp^{}(𝒞,G)`$. Let $`R^{}`$ denote a fibrant replacement functor in the strict model structure on $`Sp^{}(𝒞,G)`$. Then $`X\stackrel{}{}(R^{}F_0X)_0`$ is certainly a weak equivalence. We claim that $`R^{}F_0X`$ is already a $`U`$-spectrum. Suppose for the moment that this is true; then by lifting we can construct a stable equivalence $`R^{}F_0X\stackrel{}{}L_𝒮F_0X`$, and since $`R^{}F_0X`$ is a $`U`$-spectrum, this map is in fact a level equivalence. Hence the map $`X\stackrel{}{}(L_𝒮F_0X)_0`$ is a weak equivalence, as required. It remains to prove that $`R^{}F_0X`$ is a $`U`$-spectrum. Since $`G`$ is a Quillen equivalence, the map $`(F_0X)_n=G^nX\stackrel{}{}URG^{n+1}X=UR(F_0X)_{n+1}`$ is a weak equivalence. By lifting, we can factor the weak equivalence $`(F_0X)_{n+1}\stackrel{}{}(R^{}F_0X)_{n+1}`$ through the trivial cofibration $`(F_0X)_{n+1}\stackrel{}{}R(F_0X)_{n+1}`$. This implies that the map $`(F_0X)_n\stackrel{}{}U(R^{}F_0X)_{n+1}`$ is a weak equivalence, and hence that $`R^{}F_0X`$ is a $`U`$-spectrum, as required. ∎ In particular, this theorem means that the passage $`(𝒞,G)(Sp^{}(𝒞,G),G)`$ is idempotent, up to Quillen equivalence. This suggests that we are doing some kind of fibrant replacement of $`(𝒞,G)`$, but the author knows no way of making this precise. We now examine the functoriality of the stable model structure on $`Sp^{}(𝒞,G)`$. ###### Definition 5.2. Suppose $`𝒞`$ and $`𝒟`$ are left proper cellular model categories, $`G`$ is a left Quillen endofunctor of $`𝒞`$, and $`H`$ is a left Quillen endofunctor of $`𝒟`$. A *map of pairs* $`(\mathrm{\Phi },\tau ):(𝒞,G)\stackrel{}{}(𝒟,H)`$ is a left Quillen functor $`\mathrm{\Phi }:𝒞\stackrel{}{}𝒟`$ and a natural transformation $`\tau :\mathrm{\Phi }G\stackrel{}{}H\mathrm{\Phi }`$ such that $`\tau _A`$ is a weak equivalence for all cofibrant $`A𝒞`$. Note that there is an obvious associative and unital composition of maps of pairs. ###### Proposition 5.3. Suppose $`(\mathrm{\Phi },\tau ):(𝒞,G)\stackrel{}{}(𝒟,H)`$ is a map of pairs. Then there is an induced map of pairs $`(Sp^{(\mathrm{\Phi })},Sp^{(\tau )}):(Sp^{}(𝒞,G),G)\stackrel{}{}(Sp^{}(𝒟,H),H)`$ such that $`Sp^{(\mathrm{\Phi })}F_n=F_n\mathrm{\Phi }`$. This induced map of pairs is compatible with composition and identities. ###### Proof. Suppose $`G`$ has right adjoint $`U`$, $`H`$ has right adjoint $`V`$, and $`\mathrm{\Phi }`$ has right adjoint $`\mathrm{\Gamma }`$. The natural transformation $`\tau `$ induces a dual natural transformation $`D\tau :\mathrm{\Gamma }V\stackrel{}{}U\mathrm{\Gamma }`$. Define $`Sp^{(\mathrm{\Gamma })}:Sp^{}(𝒟,H)\stackrel{}{}Sp^{}(𝒞,G)`$ by $`(Sp^{(\mathrm{\Gamma })}Y)_n=\mathrm{\Gamma }Y_n`$, with structure maps adjoint to the composite $$\mathrm{\Gamma }Y_n\stackrel{\mathrm{\Gamma }\stackrel{~}{\sigma }}{}\mathrm{\Gamma }VY_{n+1}\stackrel{D\tau }{}U\mathrm{\Gamma }Y_{n+1}$$ where $`\stackrel{~}{\sigma }`$ is adjoint to the structure map of $`Y`$. The functor $`Sp^{(\mathrm{\Gamma })}`$ is analogous to restriction in the theory of group representations, and we must now define the analog to induction $`Sp^{(\mathrm{\Phi })}`$. To do so, first note that $`\tau `$ defines natural transformations $`\tau ^q:\mathrm{\Phi }G^q\stackrel{}{}H^q\mathrm{\Phi }`$ for all $`q`$, by iteration. Define $`(Sp^{(\mathrm{\Phi })}X)_n`$ to be the coequalizer of the two maps $$\underset{p+q+r=n}{}H^p\mathrm{\Phi }G^qX_r\underset{p+q=n}{}H^p\mathrm{\Phi }X_q$$ where the top map is induced by $`H^p\mathrm{\Phi }G^qX_r\stackrel{}{}H^p\mathrm{\Phi }X_{q+r}`$ and the bottom map is induced by $`H^p\mathrm{\Phi }G^qX_r\stackrel{H^p\tau ^q}{}H^{p+q}\mathrm{\Phi }X_r`$. To define the structure map of $`Sp^{(\mathrm{\Phi })}X`$, note that the coequalizer diagram for $`H(Sp^{(\mathrm{\Phi })}X)_n`$ is just the subdiagram of the coequalizer diagram for $`(Sp^{(\mathrm{\Phi })}X)_{n+1}`$ consisting of all terms with a positive power of $`H`$. The inclusion of diagrams induces the desired structure map $`H(Sp^{(\mathrm{\Phi })}X)_n\stackrel{}{}(Sp^{(\mathrm{\Phi })}X)_{n+1}`$. We leave to the reader the exercise in adjointness required to prove that $`Sp^{(\mathrm{\Phi })}`$ is left adjoint to $`Sp^{(\mathrm{\Gamma })}`$. The functor $`Sp^{(\mathrm{\Gamma })}`$ clearly preserves level fibrations and level trivial fibrations, so $`Sp^{(\mathrm{\Phi })}`$ is a left Quillen functor with respect to the projective model structures. Also, since $`\mathrm{Ev}_nSp^{(\mathrm{\Gamma })}=\mathrm{\Gamma }\mathrm{Ev}_n`$, we have $`F_n\mathrm{\Phi }=Sp^{(\mathrm{\Phi })}F_n`$. To show that $`Sp^{(\mathrm{\Phi })}`$ is a left Quillen functor with respect to the stable model structures, we must show that $`Sp^{(\mathrm{\Phi })}s_n^A`$ is a stable equivalence for all cofibrant $`A`$, by Theorem 2.2. Using the fact that $`Sp^{(\mathrm{\Phi })}F_n=F_n\mathrm{\Phi }`$ and the fact that $`\tau _A`$ is a weak equivalence, we reduce to showing that $`s_n\mathrm{\Phi }A`$ is a stable equivalence in $`Sp^{}(𝒟,H)`$. This follows from Theorem 3.4, so $`Sp^{(\mathrm{\Phi })}`$ is a left Quillen functor with respect to the stable model structures. the map $`\mathrm{\Phi }s_n^A`$ is a stable equivalence. Thus $`\mathrm{\Phi }:Sp^{}(𝒞,G)\stackrel{}{}Sp^{}(𝒟,H)`$ is a left Quillen functor with respect to the stable model structures. We define $`Sp^{(\tau )}`$ by defining its adjoint $`DSp^{(\tau )}:Sp^{(\mathrm{\Gamma })}V\stackrel{}{}USp^{(\mathrm{\Gamma })}`$. Indeed, $`D\tau `$ is just the prolongation of the adjoint $`D\tau `$ of $`\tau `$. Since $`\tau `$ is a weak equivalence on all cofibrant objects of $`𝒞`$, $`D\tau `$ is a weak equivalence on all fibrant objects of $`𝒟`$. To see this, note that $`\tau `$ induces a natural isomorphism in the homotopy category. Adjointness implies that $`D\tau `$ also induces a natural isomorphism in the homotopy category, and it follows that $`D\tau `$ is a weak equivalence on all fibrant objects of $`𝒟`$. Thus $`DSp^{(\tau )}`$ will be a level equivalence on all level fibrant objects of $`Sp^{}(𝒟,H)`$, so $`Sp^{(\tau )}`$ is a level equivalence on all cofibrant objects of $`Sp^{}(𝒞,G)`$. We leave it to the reader to check compatibility of $`(Sp^{(\mathrm{\Phi })},Sp^{(\tau )})`$ with compositions and identities. ∎ Proposition 5.3 and Theorem 5.1 give us a weak universal property of $`Sp^{}(𝒞,G)`$. ###### Corollary 5.4. Suppose $`(\mathrm{\Phi },\tau ):(𝒞,G)\stackrel{}{}(𝒟,H)`$ is a map of pairs such that $`H`$ is a Quillen equivalence. Then there is a functor $`\stackrel{~}{\mathrm{\Phi }}:\mathrm{Ho}Sp^{}(𝒞,G)\stackrel{}{}𝒟`$ such that $`\stackrel{~}{\mathrm{\Phi }}LF_0=L\mathrm{\Phi }:\mathrm{Ho}𝒞\stackrel{}{}\mathrm{Ho}𝒟`$. This corollary is trying to say that $`(Sp^{}(𝒞,G),G)`$ is homotopy initial among maps of pairs $`(𝒞,G)\stackrel{}{}(𝒟,H)`$ where $`H`$ is a Quillen equivalence. Though the statement of the corollary is the best statement of this concept we have been able to find, we suspect there is a better one. ###### Proof. By Proposition 5.3 there is a map of pairs $$(Sp^{(\mathrm{\Phi })},Sp^{(\tau )}):(Sp^{}(𝒞,G),G)\stackrel{}{}(Sp^{}(𝒟,H),H)$$ induced by $`(\mathrm{\Phi },\tau )`$. By Theorem 5.1, $`F_0:𝒟\stackrel{}{}Sp^{}(𝒟,H)`$ is a Quillen equivalence. Define $`\stackrel{~}{\mathrm{\Phi }}`$ to be the composite $`R\mathrm{Ev}_0LSp^{(\mathrm{\Phi })}`$. ∎ We have now shown that the correspondence $`(𝒞,G)(Sp^{}(𝒞,G),G)`$ is functorial. We would like to know that it is homotopy invariant. In particular, we would like to know that $`(Sp^{(\mathrm{\Phi })},Sp^{(\tau )})`$ is a Quillen equivalence of pairs when $`\mathrm{\Phi }`$ is a Quillen equivalence. Our proof of this seems to require some hypotheses. ###### Theorem 5.5. Suppose $`(\mathrm{\Phi },\tau ):(𝒞,G)\stackrel{}{}(𝒟,H)`$ is a map of pairs such that $`\mathrm{\Phi }`$ is a Quillen equivalence. Suppose as well that either the domains of the generating cofibrations for $`𝒞`$ can be taken to be cofibrant, or that $`\tau _X`$ is a weak equivalence for all $`X`$. Then, in the induced map of pairs $`(Sp^{(\mathrm{\Phi })},Sp^{(\tau )}):(Sp^{}(𝒞,G),G)\stackrel{}{}(Sp^{}(𝒟,H),H)`$, the Quillen functor $`Sp^{(\mathrm{\Phi })}`$ is a Quillen equivalence. ###### Proof. We will first show that $`Sp^{(\mathrm{\Phi })}`$ is a Quillen equivalence on the projective model structures. Use the same notation as in the proof of Proposition 5.3, so that $`\mathrm{\Gamma }`$ denotes the right adjoint of $`\mathrm{\Phi }`$. Then, since $`\mathrm{\Phi }`$ is a Quillen equivalence, $`\mathrm{\Gamma }`$ reflects weak equivalences between fibrant objects, by \[10, Corollary 1.3.16\]. It follows that $`Sp^{(\mathrm{\Gamma })}`$ reflects level equivalences between level fibrant objects. Hence to show that $`Sp^{(\mathrm{\Phi })}`$ is a Quillen equivalence of the projective model structures, it suffices to show that $`X\stackrel{}{}Sp^{(\mathrm{\Gamma })}RSp^{(\mathrm{\Phi })}X`$ is a level equivalence for all cofibrant $`X`$, where $`R`$ is a fibrant replacement functor in the projective model structure on $`Sp^{}(𝒟,H)`$. Thus, we need only show that $`X_n\stackrel{}{}\mathrm{\Gamma }R(Sp^{(\mathrm{\Phi })}X)_n`$ is a weak equivalence for all $`n`$ and all cofibrant $`X`$, where now $`R`$ is a fibrant replacement functor in $`𝒟`$. Since $`X_n`$ is cofibrant and $`\mathrm{\Phi }`$ is a Quillen equivalence, it suffices to show that $`\mathrm{\Phi }X_n\stackrel{}{}(Sp^{(\mathrm{\Phi })}X)_n`$ is a weak equivalence for all $`n`$ and all cofibrant $`X`$. In fact, we can assume that $`X`$ is an $`I_G`$-cell complex. Write $`X`$ as the colimit of a $`\lambda `$-sequence $$0=X^0\stackrel{}{}X^1\stackrel{}{}X^2\stackrel{}{}\mathrm{}\stackrel{}{}X^\beta \stackrel{}{}\mathrm{}\stackrel{}{}X^\lambda =X$$ where each map $`X^\beta \stackrel{}{}X^{\beta +1}`$ is a pushout of a map of $`I_G`$. We will prove that, for all $`\beta \lambda `$, $`\mathrm{\Phi }X_n^\beta \stackrel{}{}(Sp^{(\mathrm{\Phi })}X^\beta )_n`$ is a weak equivalence for all $`n`$, by transfinite induction on $`\beta `$. Getting started is easy. The limit ordinal part of the induction follows from \[9, Proposition 17.9.12\], since each of the maps $`\mathrm{\Phi }X_n^\beta \stackrel{}{}\mathrm{\Phi }X_n^{\beta +1}`$ and each of the maps $`(Sp^{(\mathrm{\Phi })}X^\beta )_n\stackrel{}{}(Sp^{(\mathrm{\Phi })}X^{\beta +1})_n`$ is a cofibration of cofibrant objects. For the succesor ordinal part of the induction, suppose $`X^\beta \stackrel{}{}X^{\beta +1}`$ is a pushout of the map $`F_mC\stackrel{F_mf}{}F_mD`$ of $`I_G`$. Then we have a pushout diagram $$\begin{array}{ccc}\mathrm{\Phi }(F_mC)_n& & \mathrm{\Phi }(F_mD)_n\\ & & & & \\ \mathrm{\Phi }X_n^\beta & & \mathrm{\Phi }X_n^{\beta +1}\end{array}$$ and another pushout diagram $$\begin{array}{ccc}(Sp^{(\mathrm{\Phi })}F_mC)_n& & (Sp^{(\mathrm{\Phi })}F_mD)_n\\ & & & & \\ (Sp^{(\mathrm{\Phi })}X^\beta )_n& & (Sp^{(\mathrm{\Phi })}X^{\beta +1})_n\end{array}$$ in $`𝒟`$. Note that $`\mathrm{\Phi }(F_mC)_n=\mathrm{\Phi }G^{nm}C`$, where we interpret $`G^{nm}C`$ to be the initial object if $`n<m`$. Similarly, $`(Sp^{(\mathrm{\Phi })}F_mC)_n=(F_m\mathrm{\Phi }C)_n=H^{nm}\mathrm{\Phi }C`$. Thus the natural transformation $`\tau `$ induces a map from the first of these pushout squares to the second. If $`C`$ (and hence also $`D`$) is cofibrant, then this map of pushout squares is a weak equivalence at both the upper left and upper right corners. Or, if $`\tau `$ is a weak equivalence for all $`X`$, then again this map is a weak equivalence at both the upper left and upper right corners. It is also a weak equivalence at the lower left corner, by the induction hypothesis. Since the top horizontal map is a cofibration in $`𝒟`$, Dan Kan’s cube lemma \[10, Lemma 5.2.6\] implies that the map is a weak equivalence on the lower right corner. This completes the induction. We have now proved that $`Sp^{(\mathrm{\Phi })}`$ is a Quillen equivalence with respect to the projective model structures. In view of Proposition 2.3, to show that $`Sp^{(\mathrm{\Phi })}`$ is a Quillen equivalence with respect to the stable model structures, we need to show that if $`Y`$ is level fibrant in $`Sp^{}(𝒟,H)`$ and $`Sp^{(\mathrm{\Gamma })}Y`$ is a $`U`$-spectrum, then $`Y`$ is a $`V`$-spectrum. To see this, note that, since $`Sp^{(\mathrm{\Gamma })}Y`$ is a $`U`$-spectrum, the natural map $`\mathrm{\Gamma }Y_n\stackrel{}{}U\mathrm{\Gamma }Y_{n+1}`$ is a weak equivalence for all $`n`$. There is a natural transformation $`D\tau :\mathrm{\Gamma }V\stackrel{}{}U\mathrm{\Gamma }`$ dual to $`\tau `$. Furthermore, $`(D\tau )_X`$ is a weak equivalence for all fibrant $`X`$, as we have seen in the proof of Proposition 5.3. Thus, the natural map $`\mathrm{\Gamma }Y_n\stackrel{}{}\mathrm{\Gamma }VY_{n+1}`$ is a weak equivalence for all $`n`$. Since $`\mathrm{\Gamma }`$ reflects weak equivalences between fibrant objects, it follows that $`Y`$ is a $`V`$-spectrum, as required. ∎ As an example of Theorem 5.5, suppose we take a pointed simplicial set $`K`$ weakly equivalent to $`S^1`$. Then there is a weak equivalence $`K\stackrel{}{}RS^1`$, where $`R`$ is the fibrant replacement functor. This induces a natural transformation of left Quillen functors $`\tau :K\stackrel{}{}RS^1`$. In Theorem 5.5, take $`𝒟=𝒞`$ equal to the model category of pointed simplicial sets, take $`\mathrm{\Phi }`$ to be the identity, and take $`\tau `$ to be this natural transformation. Then we get a Quillen equivalence between the stable model categories of spectra obtained by inverting $`K`$ and inverting $`RS^1`$. Therefore, the choice of simplicial circle does not matter, up to Quillen equivalence, for Bousfield-Friedlander spectra. We now investigate to what extent the correspondence that takes $`(𝒞,G)`$ to the stable model structure on $`Sp^{}(𝒞,G)`$ preserves monoidal structure. We begin by assuming that $`𝒞`$ is a $`𝒟`$-model category, for some symmetric monoidal model category $`𝒟`$. This means that $`𝒟`$ is a symmetric monoidal category with a compatible model structure. Since we will need to work with this compatibility, we remind the reader of the precise definition (see also \[10, Chapter 4\]). ###### Definition 5.6. Suppose $`𝒟`$ is a monoidal category. Given maps $`f:A\stackrel{}{}B`$ and $`g:C\stackrel{}{}D`$, we define the *pushout product* $`fg`$ of $`f`$ and $`g`$ to be the map $`fg:(AD)_{AC}(BC)\stackrel{}{}BD`$ induced by the commutative square $$\begin{array}{ccc}AC& \stackrel{f1}{}& BC\\ 1g& & 1g& & \\ AD& \underset{f1}{}& BD\end{array}$$ The compatibility condition we require is then that, if $`f`$ and $`g`$ are cofibrations, then so is $`fg`$, and furthermore, if one of $`f`$ and $`g`$ is a trivial cofibration, so is $`fg`$. We must also require that, if $`S`$ is the unit of $``$ and $`QS\stackrel{}{}S`$ is a cofibrant approximation, then $`QSX\stackrel{}{}X`$ is still a weak equivalence. Then, by saying that $`𝒞`$ is a $`𝒟`$-model category, we mean that $`𝒞`$ is tensored, cotensored, and enriched over $`𝒟`$, compatibly with the model structure. This compatibility is precisely analogous to the compatibility above. We then have the following theorem. ###### Theorem 5.7. Let $`𝒟`$ be a cofibrantly generated monoidal model category, and suppose the domains of the generating cofibrations are cofibrant. Suppose $`𝒞`$ is a left proper cellular $`𝒟`$-model category, and that $`G`$ is a left $`𝒟`$-Quillen endofunctor of $`𝒟`$. This means that $`G(XK)GXK`$, coherently, for $`X𝒞`$ and $`K𝒟`$. Then $`Sp^{}(𝒞,G)`$, with the stable model structure, is again a $`𝒟`$-model category, and the extension of $`G`$ is a $`𝒞`$-Quillen self-equivalence of $`Sp^{}(𝒞,G)`$. Of course, the Quillen functors $`F_n:𝒞\stackrel{}{}Sp^{}(𝒞,G)`$ will be $`𝒟`$-Quillen functors as well. ###### Proof. We define the action of $`𝒟`$ on $`Sp^{}(𝒞,G)`$ levelwise. That is, given $`XSp^{}(𝒞,G)`$ and $`K𝒟`$, we define $`(XK)_n=X_nK`$. The structure map is given by $$G(X_nK)GX_nK\stackrel{}{}X_{n+1}K.$$ One can easily verify that this makes $`Sp^{}(𝒞,G)`$ tensored over $`𝒟`$. Similarly, define $`(X^K)_n=X_n^K`$, with structure maps $`G(X_n^K)\stackrel{}{}X_{n+1}^K`$ adjoint to the composite $$G(X_n^K)KG(X_n^KK)\stackrel{G(\text{ev})}{}GX_n\stackrel{}{}X_{n+1}$$ where $`\text{ev}:X_n^KK`$ is the evaluation map, adjoint to the identity of $`X_n^K`$. This makes $`Sp^{}(𝒞,G)`$ cotensored over $`𝒟`$. Finally, given $`X`$ and $`Y`$ in $`Sp^{}(𝒞,G)`$, define $`\mathrm{Map}(X,Y)𝒟`$ to be the equalizer of the two maps $$\alpha ,\beta :\underset{n}{}\mathrm{Map}(X_n,Y_n)\stackrel{}{}\underset{n}{}\mathrm{Map}(X_n,UY_{n+1})$$ where $`\alpha `$ is the product of the maps $`\mathrm{Map}(X_n,Y_n)\stackrel{}{}\mathrm{Map}(X_n,UY_{n+1})`$ induced by the adjoint of the structure map of $`Y`$, and $`\beta `$ is the product of the maps $$\mathrm{Map}(X_{n+1},Y_{n+1})\stackrel{}{}\mathrm{Map}(UX_{n+1},UY_{n+1})\stackrel{}{}\mathrm{Map}(X_n,UY_{n+1}).$$ Here the first map exists since $`G`$ preserves the $`𝒟`$ action, and the second map is induced by the structure map of $`X`$. This functor makes $`Sp^{}(𝒞,G)`$ enriched over $`𝒟`$. We must now check that these structures are compatible with the model structure. We begin with the projective model structure on $`Sp^{}(𝒞,G)`$. One can easily check that $`F_nfg=F_n(fg)`$. Thus, if $`f`$ is one of the generating cofibrations of the projective model structure on $`Sp^{}(𝒞,G)`$, and $`g`$ is a cofibration in $`𝒟`$, then $`fg`$ is a cofibration in $`Sp^{}(𝒞,G)`$. It follows that $`fg`$ is a cofibration for $`f`$ an arbitrary cofibration of $`Sp^{}(𝒞,G)`$ (see \[21, Lemma 2.3\] and \[11, Corollary 5.3.5\]). A similar argument shows that $`fg`$ is a projective trivial cofibration in $`Sp^{}(𝒞,G)`$ if either $`f`$ is a projective cofibration in $`Sp^{}(𝒞,G)`$ or $`g`$ is a trivial cofibration in $`𝒟`$. Finally, if $`QS\stackrel{}{}S`$ is a cofibrant approximation to the unit $`S`$ in $`𝒟`$, and $`X`$ is cofibrant in $`Sp^{}(𝒞,G)`$, then each $`X_n`$ is cofibrant in $`𝒞`$, so the map $`XQS\stackrel{}{}X`$ is a level equivalence as required. Thus $`Sp^{}(𝒞,G)`$ with its projective model structure is a $`𝒟`$-model category. To show that $`Sp^{}(𝒞,G)`$ with its stable model structure is also a $`𝒟`$-model category, we need to show that, if $`f`$ is a stable trivial cofibration and $`g`$ is a cofibration in $`𝒟`$, then $`fg`$ is a stable equivalence. It suffices to check this for $`g:K\stackrel{}{}L`$ one of the generating trivial cofibrations of $`𝒟`$. In this case, by hypothesis, $`K`$ and $`L`$ are cofibrant in $`𝒟`$. Thus the functor $`K`$ is a Quillen functor with respect to the projective model structure on $`Sp^{}(𝒞,G)`$, and similarly for $`L`$. Furthermore, if $`s_n^{QC}:F_{n+1}GQC\stackrel{}{}F_nQC`$ is an element of the set $`𝒮`$, then $`s_n^{QC}Ks_n^{QCK}`$, since $`G`$ preserves the $`𝒟`$-action. In view of Theorem 3.4, the map $`s_n^{QCK}`$ is a stable equivalence. Theorem 2.2 then implies that $`K`$ is a Quillen functor with respect to the stable model structure on $`Sp^{}(𝒞,G)`$, and similarly for $`L`$. Thus, if $`f`$ is a stable trivial cofibration, so are $`fK`$ and $`fL`$. It follows from the two out of three property that $`fg`$ is a stable equivalence, as required. ∎ ###### Remark 5.8. Suppose that the functor $`G`$ is actually given by $`GX=XK`$ for some cofibrant object $`K`$ of $`𝒟`$. We then have two different ways of tensoring with $`K`$ on $`Sp^{}(𝒞,G)`$. The first way is the extension of $`G`$ to a Quillen equivalence of $`Sp^{}(𝒞,G)`$. Recall from Remark 1.7 that this functor, which we denote by $`XX\overline{}K`$, does not use the twist map. On the other hand, we also have the functor $`XXK`$ that is part of the $`𝒟`$-action on $`Sp^{}(𝒞,G)`$ constructed in Theorem 5.7. This functor *does* use the twist map as part of its structure map; indeed, in order to construct the isomorphism $`G(XK)GXK`$ we need to permute the two different copies of $`K`$. Therefore, we do not know that $`XXK`$ is a Quillen equivalence, even though $`XX\overline{}K`$ is. We will have to deal with this point more thoroughly in Section 9, when we compare $`Sp^{}(𝒞,G)`$ with symmetric spectra. Theorem 5.7 gives us a functorial stabilization. We first simplify the notation. Suppose $`K`$ is a cofibrant object of a symmetric monoidal model category $`𝒟`$. Then $`G=K`$ is a left Quillen functor on any $`𝒟`$-model category $`𝒞`$. In this case, we denote $`Sp^{}(𝒞,G)`$ by $`Sp^{}(𝒞,K)`$. ###### Corollary 5.9. Suppose $`K`$ is a cofibrant object of a cofibrantly generated symmetric monoidal model category $`𝒟`$ where the domains of the generating cofibrations can be taken to be cofibrant. Then the correspondence $`𝒞Sp^{}(𝒞,K)`$ defines an endofunctor of the category of left proper cellular $`𝒟`$-model categories. Note that the “category” of left proper cellular $`𝒟`$-model categories is not really a category, because the $`\mathrm{Hom}`$-sets need not be sets. It is really a $`2`$-category, and the correspondence $`𝒞Sp^{}(𝒞,K)`$ is actually a $`2`$-functor. See for a description of this point of view on model categories. ###### Proof. Given a left proper cellular $`𝒟`$-model category $`𝒞`$, we have seen in Theorem 5.7 that $`Sp^{}(𝒞,K)`$ is a $`𝒟`$-model category. Just as in Lemma 1.5, a $`𝒟`$-Quillen functor $`H:𝒞\stackrel{}{}𝒞^{}`$ induces a functor $`H:Sp^{}(𝒞,K)\stackrel{}{}Sp^{}(𝒞^{},K)`$, as does its right adjoint $`V`$. Since $`H`$ is defined levelwise, it preserves the action of $`𝒟`$. It is easy to check that $`V`$ preserves level fibrations and level trivial fibrations, so that $`H`$ is a $`𝒞`$-Quillen functor with respect to the projective model structures. Furthermore, we have $`Hs_n^{QC}=s_n^{HQC}`$, so Theorem 3.4 and Theorem 2.2 imply that $`H`$ is a $`𝒞`$-Quillen functor with respect to the stable model structures as well. ∎ We now point out that, if $`𝒟`$ is a symmetric monoidal model category, and $`G=K`$ for some cofibrant object $`K`$, the category $`Sp^{}(𝒟,G)`$ is almost never itself monoidal, though, as we have seen, it has an action of $`𝒟`$. To see this, consider the category $`𝒟^{}`$ of sequences from $`𝒟`$. An object of $`𝒟^{}`$ is a sequence $`X_n`$ of objects of $`𝒟`$, and a map $`f:X\stackrel{}{}Y`$ is a sequence of maps $`f_n:X_n\stackrel{}{}Y_n`$. Then $`𝒟^{}`$ is a symmetric monoidal category, where we define $`(XY)_n=_{p+q=n}X_pY_q`$. The functor $`G`$ defines a monoid $`T=(S^0,GS^0,G^2S^0,\mathrm{},G^nS^0,\mathrm{})`$ in this category, using the fact that $`G`$ preserves the $`𝒟`$-action. ###### Lemma 5.10. Suppose $`𝒟`$ is a symmetric monoidal model category and $`G`$ is a left $`𝒟`$-Quillen functor. Then $`Sp^{}(𝒟,G)`$ is the category of left modules over the monoid $`T=(S^0,GS^0,\mathrm{},G^nS^0,\mathrm{})`$. We leave the proof of this lemma to the reader. The important point is that the monoid $`T`$ is almost never commutative, and therefore $`Sp^{}(𝒟,G)`$ can not be a symmetric monoidal category with unit $`T`$. Indeed, let $`K=GS𝒟`$, so that $`G=K`$. Then $`T`$ is commutative if and only if the commutativity isomorphism on $`KK`$ is the identity. This happens only very rarely. ## 6. Symmetric spectra We have just seen that the stabilization functor $`Sp^{}(𝒞,G)`$ is not good enough in case $`𝒟`$ is a symmetric monoidal model category and $`G`$ is a $`𝒟`$-Quillen functor, because $`Sp^{}(𝒞,G)`$ is not usually itself a symmetric monoidal model category. In this section, we begin the construction of a better stabilization functor $`Sp^\mathrm{\Sigma }(𝒟,K)`$ for this case. We will concentrate on the category theory in this section, leaving the model structures for the next section. The terms used for the algebra of symmetric monoidal categories and modules over them are all defined in \[10, Section 4.1\]. Through most of this section, then, $`𝒟`$ will be a bicomplete closed symmetric monoidal category with unit $`S`$, and $`K`$ will be an object of $`𝒟`$, The category $`𝒞`$ will be a bicomplete category enriched, tensored, and cotensored over $`𝒟`$. Note that any $`𝒟`$-functor $`G`$ on $`𝒟`$ is of the form $`G(X)=XK`$ for $`K=GS`$, so we will only consider such functors. Because of this, we will drop the letter $`G`$ from our notations and replace it with $`K`$. This section is based on the symmetric spectra and sequences of . The main idea of symmetric spectra is that the commutativity isomorphism of $`𝒟`$ makes $`K^n`$ into a $`\mathrm{\Sigma }_n`$-object of $`𝒟`$, where $`\mathrm{\Sigma }_n`$ is the symmetric group on $`n`$ letters. We must keep track of this action if we expect to get a symmetric monoidal category of $`K`$-spectra. The following definition is \[11, Definition 2.1.1\]. ###### Definition 6.1. Let $`\mathrm{\Sigma }=_{n0}\mathrm{\Sigma }_n`$ be the category whose objects are the sets $`\overline{n}=\{1,2,\mathrm{},n\}`$ for $`n0`$, where $`\overline{0}=\mathrm{}`$. The morphisms of $`\mathrm{\Sigma }`$ are the isomorphisms of $`\overline{n}`$. Given a category $`𝒞`$, a *symmetric sequence* in $`𝒞`$ is a functor $`\mathrm{\Sigma }\stackrel{}{}𝒞`$. The category of symmetric sequences is the functor category $`𝒞^\mathrm{\Sigma }`$. A symmetric sequence in $`𝒞`$ is a sequence $`X_0,X_1,\mathrm{},X_n,\mathrm{}`$ of objects of $`𝒞`$ with an action of $`\mathrm{\Sigma }_n`$ on $`X_n`$. It is sometimes more useful to consider a symmetric sequence as a functor from the category of finite sets and isomorphisms to $`𝒞`$; since the category $`\mathrm{\Sigma }`$ is a skeleton of the category of finite sets and isomorphisms, there is no difficulty in doing so. As a functor category, the category of symmetric sequences in $`𝒞`$ is bicomplete if $`𝒞`$ is so; limits and colimits are taken objectwise. Furthermore, if $`𝒟`$ is a closed symmetric monoidal category, so is $`𝒟^\mathrm{\Sigma }`$, as explained in \[11, Section 2.1\]. Recall that the monoidal structure is given by $$(XY)(C)=\underset{AB=C,AB=\mathrm{}}{}X(A)Y(B)$$ where we think of $`X`$, $`Y`$, and $`XY`$ as functors from finite sets to $`𝒟`$. Equivalently, though less canonically, we have $$(XY)_n=\underset{p+q=n}{}\mathrm{\Sigma }_n\times _{\mathrm{\Sigma }_p\times \mathrm{\Sigma }_q}(X_pY_q).$$ The unit of the monoidal structure is the symmetric sequence $`(S,0,\mathrm{},0,\mathrm{})`$, where $`0`$ is the initial object of $`𝒞`$. To define the closed structure, we first define $`\mathrm{Hom}_{\mathrm{\Sigma }_n}(X,Y)`$ for $`X,Y𝒟^{\mathrm{\Sigma }_n}`$ in the usual way, as an equalizer of the two obvious maps $`\mathrm{Hom}(X,Y)\stackrel{}{}\mathrm{Hom}(X\times \mathrm{\Sigma }_n,Y)`$. The closed structure is then given by $$\mathrm{Hom}(X,Y)_k=\underset{n}{}\mathrm{Hom}_{\mathrm{\Sigma }_n}(X_n,Y_{n+k}).$$ If $`𝒞`$ is enriched, tensored, and cotensored over $`𝒟`$, then $`𝒞^\mathrm{\Sigma }`$ is enriched, tensored, and cotensored over $`𝒟^\mathrm{\Sigma }`$. Indeed, the same definition as above works to define the tensor structure. The cotensor structure is defined as follows. First we define $`\mathrm{Hom}_{\mathrm{\Sigma }_n}(K,X)`$ for $`X𝒞^{\mathrm{\Sigma }_n}`$ and $`K𝒟^{\mathrm{\Sigma }_n}`$ as an appropriate equalizer. Then, for $`X𝒞^\mathrm{\Sigma }`$ and $`K𝒟^\mathrm{\Sigma }`$, we define $`X_k^K=_n\mathrm{Hom}_{\mathrm{\Sigma }_n}(K_n,X_{n+k})`$. The enrichment $`\mathrm{Map}(X,Y)`$ is defined similarly. In the same way, if $`𝒞`$ is an enriched monoidal category over $`𝒟`$, then $`𝒞^\mathrm{\Sigma }`$ is an enriched monoidal category over $`𝒟^\mathrm{\Sigma }`$. Consider the free commutative monoid $`\mathrm{Sym}(K)`$ on the object $`(0,K,0,\mathrm{},0,\mathrm{})`$ of $`𝒟^\mathrm{\Sigma }`$. One can easily check that $`\mathrm{Sym}(K)`$ is the symmetric sequence $`(S^0,K,KK,\mathrm{},K^n,\mathrm{})`$ where $`\mathrm{\Sigma }_n`$ acts on $`K^n`$ by the commutativity isomorphism, as in \[11, Section 4.4\]. ###### Definition 6.2. Suppose $`𝒟`$ is a symmetric monoidal model category, $`𝒞`$ is a $`𝒟`$-model category, and $`K`$ is an object of $`𝒟`$. The category of *symmetric spectra* $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is the category of modules in $`𝒞^\mathrm{\Sigma }`$ over the commutative monoid $`\mathrm{Sym}(K)`$ in $`𝒟^\mathrm{\Sigma }`$. That is, a symmetric spectrum is a sequence of objects $`X_n𝒞^{\mathrm{\Sigma }_n}`$ and $`\mathrm{\Sigma }_n`$-equivariant maps $`X_nK\stackrel{}{}X_{n+1}`$, such that the composite $$X_nK^p\stackrel{}{}X_{n+1}K^{p1}\stackrel{}{}\mathrm{}\stackrel{}{}X_{n+p}$$ is $`\mathrm{\Sigma }_n\times \mathrm{\Sigma }_p`$-equivariant for all $`n,p0`$. A map of symmetric spectra is a collection of $`\mathrm{\Sigma }_n`$-equivariant maps $`X_n\stackrel{}{}Y_n`$ compatible with the structure maps of $`X`$ and $`Y`$. Because $`\mathrm{Sym}(K)`$ is a commutative monoid, the category $`Sp^\mathrm{\Sigma }(𝒟,K)`$ is a bicomplete closed symmetric monoidal category, with $`\mathrm{Sym}(K)`$ itself as the unit (see Lemma 2.2.2 and Lemma 2.2.8 of ). We denote the monoidal structure by $`XY=X_{\mathrm{Sym}(K)}Y`$, and the closed structure by $`\mathrm{Hom}_{\mathrm{Sym}(K)}(X,Y)`$. Similarly, $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is bicomplete, enriched, tensored, and cotensored over $`Sp^\mathrm{\Sigma }(𝒟,K)`$ with the tensor structure denoted $`XY`$ again, and, if $`𝒞`$ is a $`𝒟`$-monoidal model category, then $`Sp^\mathrm{\Sigma }(𝒞,K)`$ will be a monoidal category enriched over $`Sp^\mathrm{\Sigma }(𝒟,K)`$. Of course, if we take $`𝒞=\mathrm{𝐒𝐒𝐞𝐭}_{}`$ and $`K=S^1`$, we recover the definition of symmetric spectra given in , except that we are using right modules instead of left modules. ###### Definition 6.3. Given $`n0`$, the *evaluation functor* $`\mathrm{Ev}_n:Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}𝒞`$ takes $`X`$ to $`X_n`$. the evaluation functor has a left adjoint $`F_n:𝒞\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K)`$, defined by $`F_nX=\stackrel{~}{F}_nX\mathrm{Sym}(K)`$, where $`\stackrel{~}{F}_nX`$ is the symmetric sequence $`(0,\mathrm{},0,\mathrm{\Sigma }_n\times X,0,\mathrm{})`$. Note that $`F_0X=(X,XK,\mathrm{},XK^n,\mathrm{})`$, and in particular $`F_0S=\mathrm{Sym}(K)`$. Also, if $`X𝒞`$ and $`Y𝒟`$, there is a natural isomorphism $`F_nXF_mYF_{n+m}(XY)`$, just as in \[11, Proposition 2.2.6\]. In particular, $`F_0:𝒟\stackrel{}{}Sp^\mathrm{\Sigma }(𝒟,K)`$ is a (symmetric) monoidal functor, and so $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is naturally enriched, tensored, and cotensored over $`𝒟`$. In fact, this structure is very simple. Indeed, if $`XSp^\mathrm{\Sigma }(𝒞,K)`$ and $`L𝒟`$, $`XL=X_{\mathrm{Sym}(K)}F_0L`$ is just the symmetric sequence whose $`n`$th term is $`X_nL`$. The structure map is the composite $$X_nLK\stackrel{1\tau }{}X_nKL\stackrel{}{}X_{n+1}L.$$ Note the presence of the twist map; this is required even when $`L=K`$ to get a symmetric spectrum, unlike the case of ordinary spectra. Similarly, $`X^L=\mathrm{Hom}_{\mathrm{Sym}(K)}(F_0L,X)`$ is the symmetric sequence whose $`n`$th term is $`X_n^L`$, with the twist map again appearing as part of the structure map. ###### Remark 6.4. Just as in the spectrum case, the functor $`\mathrm{Ev}_n`$ has a right adjoint $`R_n:𝒞\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K)`$. Indeed, $`R_nX=\mathrm{Hom}(\mathrm{Sym}(K),\stackrel{~}{R}_nL)`$, where $`\stackrel{~}{R}_nL`$ is the symmetric sequence concentrated in degree $`n`$ whose $`n`$th term is $`X^{\mathrm{\Sigma }_n}`$. ## 7. Model structures on symmetric spectra Throughout this section, $`𝒟`$ will denote a left proper cellular symmetric monoidal model category, $`𝒞`$ will denote a left proper cellular $`𝒟`$-model category, and $`K`$ will denote a cofibrant object of $`𝒟`$. In this section, we discuss the projective and stable model structures on the category $`Sp^\mathrm{\Sigma }(𝒞,K)`$ of symmetric spectra. The results in this section are very similar to the corresponding results for spectra, so we will leave most of the proofs to the reader. ###### Definition 7.1. A map $`fSp^\mathrm{\Sigma }(𝒞,K)`$ is a *level equivalence* if each map $`f_n`$ is a weak equivalence in $`𝒞`$. Similarly, $`f`$ is a *level fibration* (resp. *level cofibration*, *level trivial fibration*, *level trivial cofibration*) if each map $`f_n`$ is a fibration (resp. cofibration, trivial fibration, trivial cofibration) in $`𝒞`$. The map $`f`$ is a *projective cofibration* if $`f`$ has the left lifting property with respect to every level trivial fibration. Then, just as in Definition 1.9, if we denote the generating cofibrations of $`𝒞`$ by $`I`$ and the generating trivial cofibrations by $`J`$, we define $`I_K=_nF_nI`$ and $`J_K=_nF_nJ`$. We have analogues of 1.101.13 with the same proofs. This gives us the projective model structure. ###### Theorem 7.2. The projective cofibrations, the level fibrations, and the level equivalences define a left proper cellular model structure on $`Sp^\mathrm{\Sigma }(𝒞,K)`$. The set $`I_K`$ is the set of generating cofibrations of the projective model structure, and $`J_K`$ is the set of generating trivial cofibrations. The cellularity of the projective model structure is proved in the Appendix. Note that $`\mathrm{Ev}_n`$ takes level (trivial) fibrations to (trivial) fibrations, so $`\mathrm{Ev}_n`$ is a right Quillen functor and $`F_n`$ is a left Quillen functor. ###### Theorem 7.3. The category $`Sp^\mathrm{\Sigma }(𝒟,K)`$, with the projective model structure, is a symmetric monoidal model category. The category $`Sp^\mathrm{\Sigma }(𝒞,K)`$, with its projective model structure, is a $`Sp^\mathrm{\Sigma }(𝒟,K)`$-model category. ###### Proof. We first show that the pushout product $`fg`$ is a (trivial) cofibration when $`f`$ is a cofibration in $`Sp^\mathrm{\Sigma }(𝒞,K)`$, and $`g`$ is a cofibration in $`Sp^\mathrm{\Sigma }(𝒟,K)`$ (and one of them is a level equivalence). As explained in \[10, Corollary 2.5\]. we may as well assume that $`f`$ and $`g`$ belong to the sets of generating cofibrations or generating trivial cofibrations. In either case, we have $`f=F_mf^{}`$ and $`g=F_ng^{}`$. But then $`fg=F_{m+n}(f^{}g^{})`$. Since $`F_{m+n}`$ is a Quillen functor, the result follows. Now let $`QS`$ denote a cofibrant replacement for the unit $`S`$ in $`𝒟`$. Then $`F_0QS`$ is a cofibrant replacement for $`F_0S=\mathrm{Sym}(K)`$ in $`Sp^\mathrm{\Sigma }(𝒟,K)`$. Indeed, $`F_0QS`$ is cofibrant, and $`\mathrm{Ev}_nF_0QS`$ is just $`QSK^n`$. Since $`K`$ is cofibrant and $`𝒟`$ is a monoidal model category, the map $`F_0QS\stackrel{}{}F_0S`$ is a level equivalence. Now, if $`X`$ is cofibrant in $`Sp^\mathrm{\Sigma }(𝒞,K)`$, then each $`X_n`$ is cofibrant. Hence the map $`X_nQS\stackrel{}{}X_n`$ is a weak equivalence for all $`n`$, and so the map $`XF_0QS\stackrel{}{}X`$ is a level equivalence, as required. ∎ We point out here that one can show that the projective model structure on $`Sp^\mathrm{\Sigma }(𝒟,K)`$ satisfies the monoid axiom of , assuming that $`𝒟`$ itself does so. This means there is a projective model structure on the category of monoids in $`Sp^\mathrm{\Sigma }(𝒟,K)`$ and on the category of modules over any monoid. We do not include the proofs of these statement since we have not been able to prove the analogous statements for the stable model structure. The projective cofibrations of symmetric spectra are more complicated than they are in the case of ordinary spectra. ###### Definition 7.4. Define the symmetric spectrum $`\overline{\mathrm{Sym}(K)}`$ in $`Sp^\mathrm{\Sigma }(𝒟,K)`$ to be $`0`$ in degree $`0`$ and $`K^n`$ in degree $`n`$, for $`n>0`$, with the obvious structure maps. Define the $`n`$th *latching space* $`L_nX`$ of $`XSp^\mathrm{\Sigma }(𝒞,K)`$ by $`L_nX=\mathrm{Ev}_n(X\overline{\mathrm{Sym}(K)})`$. The obvious map $`i:\overline{\mathrm{Sym}(K)}\stackrel{}{}\mathrm{Sym}(K)`$ induces a $`\mathrm{\Sigma }_n`$-equivariant natural transformation $`L_nX\stackrel{}{}X`$. Note that the latching space is a $`\mathrm{\Sigma }_n`$-object of $`𝒞`$. There is a model structure on $`\mathrm{\Sigma }_n`$-objects of $`𝒞`$ where the fibrations and weak equivalences are the underlying ones. This model structure is cofibrantly generated: if $`I`$ is the set of generating cofibrations of $`𝒞`$, then the set of generating cofibrations of $`𝒞^{\mathrm{\Sigma }_n}`$ is the set $`\mathrm{\Sigma }_n\times I`$. Here, for an object $`A`$, $`\mathrm{\Sigma }_n\times A`$ is the coproduct of $`n!`$ copies of $`A`$, given the obvious $`\mathrm{\Sigma }_n`$-structure. ###### Proposition 7.5. A map $`f:X\stackrel{}{}Y`$ in $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is a projective (trivial) cofibration if and only if the induced map $`\mathrm{Ev}_n(fi):X_n_{L_nX}L_nY\stackrel{}{}Y_n`$ is a (trivial) cofibration in $`𝒞^{\mathrm{\Sigma }_n}`$ for all $`n`$. ###### Proof. We only prove the cofibration case, as the trivial cofibration case is analogous. If each map $`X_n_{L_nX}L_nY\stackrel{}{}Y_n`$ is a cofibration, then we can show that $`f`$ is a projective cofibration by showing $`f`$ has the left lifting property with respect to level trivial fibrations. Indeed, we construct a lift by induction, just as in the proof of Proposition 1.15. To prove the converse, it suffices to show that $`\mathrm{Ev}_n(fi)`$ is a $`\mathrm{\Sigma }_n`$-cofibration for $`fI_K`$, since $`\mathrm{Ev}_n`$ is itself a left Quillen functor. Then we can write $`f=F_mg`$, and we find that $`\mathrm{Ev}_n(fi)`$ is an isomorphism when $`nm`$, and is the map $`\mathrm{\Sigma }_m\times g`$ when $`n=m`$. This is a $`\mathrm{\Sigma }_m`$-cofibration, as required. ∎ We must now localize the projective model structure to obtain the stable model structure. ###### Definition 7.6. A symmetric spectrum $`XSp^\mathrm{\Sigma }(𝒞,K)`$ is an *$`\mathrm{\Omega }`$-spectrum* if $`X`$ is level fibrant and the adjoint $`X_n\stackrel{}{}X_{n+1}^K`$ of the structure map is a weak equivalence for all $`n`$. Just as with Bousfield-Friedlander spectra, we would like the $`\mathrm{\Omega }`$-spectra to be the fibrant objects in the stable model structure. We invert the same maps we did in that case. ###### Definition 7.7. Define the set of maps $`𝒮`$ in $`Sp^\mathrm{\Sigma }(𝒞,K)`$ to be $`\{F_{n+1}(QCK)\stackrel{s_n^{QC}}{}F_nQC\}`$ as $`C`$ runs through the domains and codomains of the generating cofibrations of $`𝒞`$, and $`n0`$. The map $`s_n^{QC}`$ is adjoint to the map $`QCK\stackrel{}{}\mathrm{Ev}_{n+1}F_nQC=\mathrm{\Sigma }_{n+1}\times (QCK)`$ corresponding to the identity of $`\mathrm{\Sigma }_{n+1}`$. Define the *stable model structure* on $`Sp^\mathrm{\Sigma }(𝒞,K)`$ to be the Bousfield localization with respect to $`𝒮`$ of the projective model structure on $`Sp^\mathrm{\Sigma }(𝒞,K)`$. The $`𝒮`$-local weak equivalences are called the *stable equivalences*, and the $`𝒮`$-local fibrations are called the *stable fibrations*. The following theorem is then analogous to Theorem 3.4, and has the same proof. ###### Theorem 7.8. The stably fibrant symmetric spectra are the $`\mathrm{\Omega }`$-spectra. Furthermore, for all cofibrant $`A𝒞`$ and for all $`n0`$, the map $`F_{n+1}(AK)\stackrel{s_n^A}{}F_nA`$ is a stable equivalence. Just as in Corollary 3.5, this theorem implies that, when $`𝒞=\mathrm{𝐒𝐒𝐞𝐭}_{}`$ or $`\mathrm{𝐓𝐨𝐩}_{}`$, $`Sp^\mathrm{\Sigma }(𝒟,K)`$ is the same as the stable model category on (simplicial or topological) symmetric spectra discussed in . The analog of Corollary 3.6 also holds, with the same proof, so that tensoring with $`K`$ is a Quillen endofunctor of $`Sp^\mathrm{\Sigma }(𝒞,K)`$. Of course, we want this functor to be a Quillen equivalence. As in Definition 3.7, we prove this by introducing the shift functors. ###### Definition 7.9. Define the *right shift functor* $`s:Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K)`$ by $`sX=\mathrm{Hom}_{\mathrm{Sym}(K)}(F_1S,X)`$. Thus $`(sX)_n=X_{n+1}`$, where the $`\mathrm{\Sigma }_n`$-action on $`X_{n+1}`$ is induced by the usual inclusion $`\mathrm{\Sigma }_n\stackrel{}{}\mathrm{\Sigma }_{n+1}`$. The structure maps of $`sX`$ are the same as the structure maps of $`X`$. Define the *left shift functor* $`t:Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K)`$ by $`tX=X_{\mathrm{Sym}(K)}F_1S`$, so that $`t`$ is left adjoint of $`s`$. We have $`(tX)_n=\mathrm{\Sigma }_n\times _{\mathrm{\Sigma }_{n1}}X_{n1}`$, with the induced structure maps. Note that adjointness gives natural isomorphisms $$(sX)^K\mathrm{Hom}_{\mathrm{Sym}(K)}(F_1K,X)s(X^K).$$ There is a map $`F_1K\stackrel{}{}F_0S`$ which is the identity in degree $`1`$. By adjointness, this map induces a map $$X=\mathrm{Hom}_{\mathrm{Sym}(K)}(F_0S,X)\stackrel{}{}\mathrm{Hom}_{\mathrm{Sym}(K)}(F_1K,X)=(sX)^K.$$ $`X`$ is an $`\mathrm{\Omega }`$-spectrum if and only if this map is a level equivalence and $`X`$ is level fibrant. Therefore, the same method used to prove Theorem 3.8 also proves the following theorem. ###### Theorem 7.10. The functors $`XXK`$ and $`t`$ are Quillen equivalences on $`Sp^\mathrm{\Sigma }(𝒞,K)`$. Furthermore, $`Rs`$ is naturally isomorphic to $`L(K)`$ and $`R(()^K)`$ is naturally isomorphic to $`Lt`$. We have now shown that $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is a $`K`$-stabilization of $`𝒞`$. However, for this construction to be better than ordinary spectra, we must show that $`Sp^\mathrm{\Sigma }(𝒟,K)`$ is a symmetric monoidal model category. ###### Theorem 7.11. Suppose that the domains of the generating cofibrations of both $`𝒞`$ and $`𝒟`$ are cofibrant. Then the category $`Sp^\mathrm{\Sigma }(𝒟,K)`$ is a symmetric monoidal model category, and the category $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is a $`Sp^\mathrm{\Sigma }(𝒟,K)`$-model category. ###### Proof. We prove this theorem in the same way as Theorem 5.7. Since the cofibrations in the stable model structure are the same as the cofibrations in the projective model structure, the only thing to check is that $`fg`$ is a stable equivalence when $`f`$ and $`g`$ are cofibrations and one of them is a stable equivalence. We may as well assume that $`f:F_nA\stackrel{}{}F_nB`$ is a generating cofibration in $`Sp^\mathrm{\Sigma }(𝒞,K)`$ and $`g`$ is a stable trivial cofibration in $`Sp^\mathrm{\Sigma }(𝒟,K)`$; the argument for $`f`$ a stable trivial cofibration and $`g`$ a generating cofibration in $`Sp^\mathrm{\Sigma }(𝒟,K)`$ is the same. Then, by hypothesis, $`A`$ and $`B`$ are cofibrant in $`𝒞`$. We claim that $`F_nA`$ is a Quillen functor $`Sp^\mathrm{\Sigma }(𝒟,K)`$ to $`Sp^\mathrm{\Sigma }(𝒞,K)`$ with their stable model structures, and similarly for $`F_nB`$. Indeed, in view of Theorem 2.2, it suffices to show that $`F_nAs_m^{QC}`$ is a stable equivalence for all $`m0`$ and all domains or codomains $`C`$ of the generating cofibrations of $`𝒞`$. But one can easily check that $`F_nAs_m^{QC}=s_{n+m}^{AQC}`$. Then Theorem 7.8 implies that this map is a stable equivalence, as required. Thus, both functors $`F_nA`$ and $`F_nB`$ are Quillen functors in the stable model structures. A two out of three argument, as in Theorem 5.7, then shows that $`fg`$ is a stable equivalence, as required. ∎ Note that the functor $`F_0:𝒟\stackrel{}{}Sp^\mathrm{\Sigma }(𝒟,K)`$ is a symmetric monoidal Quillen functor, so of course $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is a $`𝒟`$-model category as well, under the hypotheses of Theorem 7.11. In fact, we only need the domains of the generating cofibrations of $`𝒟`$ to be cofibrant to conclude that $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is a $`𝒟`$-model category, using the argument of Theorem 7.11. As we mentioned above, we do not know if the stable model structure satisfies the monoid axiom. Given a particular monoid $`R`$, one could attempt to localize the projective model structure on $`R`$-modules to obtain a stable model structure. However, for this to work one would need to know that the projective model structure is cellular, and the author does not see how to prove this. This plan will certainly fail for the category of monoids, since the projective model structure on monoids will not be left proper in general. We also point out that it may be possible to prove some of the results of Section 4 for symmetric spectra. All of those results cannot hold, since stable homotopy isomorphisms do not coincide with stable equivalences even for symmetric spectra of simplicial sets. Nevertheless, in that case, every stable homotopy isomorphism is a stable equivalence \[11, Theorem 3.1.11\], and there is a replacement for the functor $`R^{\mathrm{}}`$ constructed in . We do not know if these results hold for symmetric spectra over a general well-behaved finitely generated model category. ## 8. Properties of symmetric spectra In this section, we point out that the arguments of Section 5 also apply to symmetric spectra. In particular, if smashing with $`K`$ is already a Quillen equivalence on $`𝒞`$, then $`F_0:𝒞\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K)`$ is a Quillen equivalence. This means that, under mild hypotheses, the homotopy category of $`𝒞`$ is enriched, tensored, and cotensored over $`\mathrm{Ho}Sp^\mathrm{\Sigma }(𝒟,K)`$. We also show symmetric spectra are functorial in an appropriate sense. In particular, we show that the Quillen equivalence class of $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is an invariant of the homotopy type of $`K`$. Throughout this section, $`𝒟`$ will denote a left proper cellular symmetric monoidal model category, $`𝒞`$ will denote a left proper cellular $`𝒞`$-model category, and $`K`$ will denote a cofibrant object of $`𝒟`$. The proof of the following theorem is the same as the proof of Theorem 5.1. ###### Theorem 8.1. Suppose smashing with $`K`$ is a Quillen equivalence on $`𝒞`$. Then $`F_0:𝒞\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K)`$ is a Quillen equivalence. ###### Corollary 8.2. Suppose that the domains of the generating cofibrations of both $`𝒞`$ and $`𝒟`$ are cofibrant, and suppose that smashing with $`K`$ is a Quillen equivalence on $`𝒞`$. Then $`\mathrm{Ho}𝒟`$ is enriched, tensored, and cotensored over $`\mathrm{Ho}Sp^\mathrm{\Sigma }(𝒟,K)`$. ###### Proof. Note that $`\mathrm{Ho}Sp^\mathrm{\Sigma }(𝒞,K)`$ is certainly enriched, tensored, and cotensored over $`\mathrm{Ho}Sp^\mathrm{\Sigma }(𝒟,K)`$. Now use the equivalence of categories $`LF_0:\mathrm{Ho}𝒞\stackrel{}{}\mathrm{Ho}Sp^\mathrm{\Sigma }(𝒞,K)`$ to transport this structure back to $`𝒞`$. ∎ Recall that the homotopy category of any model category is naturally enriched, tensored, and cotensored over $`\mathrm{Ho}\mathrm{𝐒𝐒𝐞𝐭}`$ \[10, Chapter 5\]. This corollary is the first step to the assertion that the homotopy category of any stable (with respect to the suspension) model category is naturally enriched, tensored, and cotensored over the homotopy category of (simplicial) symmetric spectra. See for further results along these lines. Symmetric spectra are functorial in a natural way. ###### Theorem 8.3. Suppose the domains of the generating cofibrations of $`𝒟`$, $`𝒞`$, and the left proper cellular $`𝒞`$-model category $`𝒞^{}`$ are cofibrant. Then any $`𝒟`$-Quillen functor $`\mathrm{\Phi }:𝒞\stackrel{}{}𝒞^{}`$ extends naturally to a $`Sp^\mathrm{\Sigma }(𝒟,K)`$-Quillen functor $$Sp^\mathrm{\Sigma }(\mathrm{\Phi }):Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞^{},K).$$ Furthermore, if $`\mathrm{\Phi }`$ is a Quillen equivalence, so is $`Sp^\mathrm{\Sigma }(\mathrm{\Phi })`$. ###### Proof. The functor $`\mathrm{\Phi }`$ induces a $`𝒟^\mathrm{\Sigma }`$-functor $`𝒞^\mathrm{\Sigma }\stackrel{}{}(𝒞^{})^\mathrm{\Sigma }`$, which takes the symmetric sequence $`(X_n)`$ to the symmetric sequence $`(\mathrm{\Phi }X_n)`$. It follows that $`\mathrm{\Phi }`$ induces a $`Sp^\mathrm{\Sigma }(𝒟,K)`$-functor $`Sp^\mathrm{\Sigma }(\mathrm{\Phi }):Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞^{},K)`$, that takes the symmetric spectrum $`(X_n)`$ to the symmetric spectrum $`(\mathrm{\Phi }X_n)`$, with structure maps $$\mathrm{\Phi }X_nK\mathrm{\Phi }(X_nK)\stackrel{}{}\mathrm{\Phi }X_{n+1}.$$ Let $`\mathrm{\Gamma }`$ denote the right adjoint of $`\mathrm{\Phi }`$. Then the right adjoint of $`Sp^\mathrm{\Sigma }(\mathrm{\Phi })`$ is $`Sp^\mathrm{\Sigma }(\mathrm{\Gamma })`$, which takes the symmetric spectrum $`(Y_n)`$ to the symmetric spectrum $`(\mathrm{\Gamma }Y_n)`$, with structure maps adjoint to the composite $$\mathrm{\Phi }(\mathrm{\Gamma }Y_nK)\mathrm{\Phi }\mathrm{\Gamma }Y_nK\stackrel{}{}Y_nK\stackrel{}{}Y_{n+1}.$$ Since $`\mathrm{Ev}_nSp^\mathrm{\Sigma }(\mathrm{\Gamma })=\mathrm{\Gamma }\mathrm{Ev}_n`$, it follows that $`Sp^\mathrm{\Sigma }(\mathrm{\Phi })F_n=F_n\mathrm{\Phi }`$. It is clear that $`Sp^\mathrm{\Sigma }(\mathrm{\Gamma })`$ preserves level fibrations and level equivalences, so $`Sp^\mathrm{\Sigma }(\mathrm{\Phi })`$ is a Quillen functor with respect to the projective model structure. In view of Theorem 2.2, to see that $`Sp^\mathrm{\Sigma }(\mathrm{\Phi })`$ defines a Quillen functor with respect to the stable model structures, it suffices to show that $`Sp^\mathrm{\Sigma }(\mathrm{\Phi })(s_n^{QC})`$ is a stable equivalence for all domains and codomains of $`C`$ of the generating cofibrations of $`𝒞`$. But one can readily verify that $`Sp^\mathrm{\Sigma }(\mathrm{\Phi })(s_n^{QC})=s_n^{\mathrm{\Phi }QC}`$, which is a stable equivalence as required. Thus $`Sp^\mathrm{\Sigma }(\mathrm{\Phi })`$ is a Quillen functor with respect to the stable model structures. If $`\mathrm{\Phi }`$ is a Quillen equivalence, one can easily check that $`Sp^\mathrm{\Sigma }(\mathrm{\Phi })`$ is a Quillen equivalence with respect to the projective model structure. To see that it is still a Quillen equivalence with respect to the stable model structures, we need only show that $`Sp^\mathrm{\Sigma }(\mathrm{\Gamma })`$ reflects stably fibrant objects, in view of Proposition 2.3. But, if $`X`$ is level fibrant and $`Sp^\mathrm{\Sigma }(\mathrm{\Gamma })(X)`$ is an $`\mathrm{\Omega }`$-spectrum, this means that the map $`\mathrm{\Gamma }X_n\stackrel{}{}(\mathrm{\Gamma }X_{n+1})^K\mathrm{\Gamma }(X_{n+1}^K)`$ is a weak equivalence for all $`n`$. Since $`\mathrm{\Gamma }`$ reflects weak equivalences between fibrant objects, this means that $`X`$ is an $`\mathrm{\Omega }`$-spectrum, as required. ∎ Symmetric spectra are also functorial, in a limited sense, in the cofibrant object $`K`$. ###### Theorem 8.4. Suppose $`f:K\stackrel{}{}K^{}`$ is a weak equivalence of cofibrant objects of $`𝒟`$, and suppose the domains of the generating cofibrations of $`𝒟`$ and $`𝒞`$ are cofibrant. Then $`f`$ induces a Quillen equivalence $`Sp^\mathrm{\Sigma }(𝒞,f):Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K^{})`$ which is natural with respect to $`𝒟`$-Quillen functors of $`𝒞`$. ###### Proof. The map $`f`$ induces a map of commutative monoids $`\mathrm{Sym}(K)\stackrel{}{}\mathrm{Sym}(K^{})`$. This induces the usual induction and restriction adjunction $$Sp^\mathrm{\Sigma }(𝒞,f):Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K^{}).$$ That is, if $`X`$ is in $`Sp^\mathrm{\Sigma }(𝒞,K)`$, then $`Sp^\mathrm{\Sigma }(𝒞,f)(X)=X_{\mathrm{Sym}(K)}\mathrm{Sym}(K^{})`$. Restriction obviously preserves level fibrations and level equivalences, so is a Quillen functor with respect to the projective model structure. One can easily check that $`Sp^\mathrm{\Sigma }(𝒞,f)F_n=F_n`$. It follows that $`Sp^\mathrm{\Sigma }(𝒞,f)(s_n^{QC})`$ is the map $$F_{n+1}(QCK)\stackrel{}{}F_nQC$$ in $`Sp^\mathrm{\Sigma }(𝒞,K^{})`$. The weak equivalence $`QCK\stackrel{}{}QCK^{}`$ induces a level equivalence $`F_{n+1}(QCK)\stackrel{}{}F_{n+1}(QCK^{})`$. Since the map $`F_{n+1}(QCK^{})\stackrel{}{}F_nQC`$ is a stable equivalence, so is the given map. Thus induction is a Quillen functor with respect to the stable model structures. We now prove that induction is a Quillen equivalence between the projective model structures. The proof of this is similar to the proof of Theorem 5.5. That is, restriction certainly reflects level equivalences between level fibrant objects. It therefore suffices to show that the map $`X\stackrel{}{}X_{\mathrm{Sym}(K)}\mathrm{Sym}(K^{})`$ is a level equivalence for all cofibrant $`X`$. The argument of Theorem 5.5 will prove this without difficulty. To show that induction is a Quillen equivalence between the stable model structures, we need only check that restriction reflects stably fibrant objects. This follows from the fact that the map $`Z^K^{}\stackrel{}{}Z^K`$ is a weak equivalence for all fibrant $`Z`$. ∎ In particular, it does not matter, up to Quillen equivalence, what model of the simplicial circle one takes in forming the symmetric spectra of . ## 9. Comparison of spectra and symmetric spectra In this section, suppose $`𝒟`$ is a left proper cellular symmetric monoidal model category, $`K`$ is a cofibrant object of $`𝒟`$, and $`𝒞`$ is a left proper cellular $`𝒟`$-model category. Let $`G`$ denote the left Quillen endofunctor $`GX=XK`$ of $`𝒞`$. Then we have two different stabilizations of $`𝒞`$, namely $`Sp^{}(𝒞,G)`$ and $`Sp^\mathrm{\Sigma }(𝒞,K)`$. The object of this section is to compare them. We show that $`Sp^{}(𝒞,G)`$ and $`Sp^\mathrm{\Sigma }(𝒞,K)`$ are related by a chain of Quillen equivalences whenever the cyclic permutation self-map of $`KKK`$ is homotopic to the identity. This is not the ideal theorem; one might hope for a direct Quillen equivalence rather than a zigzag of Quillen equivalences, and one might hope for weaker hypotheses, or even no hypotheses. However, some hypotheses are necessary, as pointed out to the author by Jeff Smith. Indeed, the category $`\mathrm{Ho}Sp^\mathrm{\Sigma }(𝒟,K)`$ is symmetric monoidal, and therefore $`\mathrm{Ho}Sp^\mathrm{\Sigma }(𝒟,K)(F_0S,F_0S)`$, the self-maps of the unit, form a commutative monoid. If we have a chain of Quillen equivalences between $`Sp^{}(𝒟,K)`$ and $`Sp^\mathrm{\Sigma }(𝒟,K)`$ that preserves the functors $`F_0`$, then $`\mathrm{Ho}Sp^{}(𝒟,K)(F_0S,F_0S)`$ would also have to be a commutative monoid. With sufficiently many hypotheses on $`𝒟`$ and $`K`$, for example if $`𝒟`$ is the category of simplicial sets and $`K`$ is a finite simplicial set, we have seen in in Section 4 that this mapping set is the colimit $`\mathrm{colim}\mathrm{Ho}𝒟(K^n,K^n)`$. There are certainly examples where this monoid is not commutative; for example $`K`$ could be the mod $`p`$ Moore space, and then homology calculations show this colimit is not commutative. In fact, this monoid will be commutative if and only if the cyclic permutation map of $`KKK`$ becomes the identity in $`\mathrm{Ho}𝒟`$ after tensoring with sufficiently many copies of $`K`$. Hence we need some hypothesis on the cyclic permutation map. The heart of our argument is the following theorem. The argument of this theorem can be summarized by saying that commuting stabilization functors are equivalent, and as such, was suggested to the author by Mike Hopkins in a different context. Recall that there are two different ways to tensor with $`K`$ on $`Sp^{}(𝒞,G)`$; the functor $`XX\overline{}K`$ that is a Quillen equivalence but does not involve the twist map, and the functor $`XXK`$ that may not be a Quillen equivalence but does involve the twist map. ###### Theorem 9.1. Suppose that the functor $`XXK`$ is a Quillen equivalence of $`Sp^{}(𝒞,G)`$, and also that the domains of the generating cofibrations of $`𝒟`$ are cofibrant. Then there is a $`𝒟`$-model category $``$ together with $`𝒟`$-Quillen equivalences $`Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}\stackrel{}{}Sp^{}(𝒞,G)`$. Furthermore, we have a natural isomorphism $`[\mathrm{Ho}Sp^\mathrm{\Sigma }(𝒞,K)](F_0A,F_0B)[\mathrm{Ho}Sp^{}(𝒞,G)](F_0A,F_0B)`$ for $`A,B𝒞`$. ###### Proof. We take $`=Sp^{}(Sp^\mathrm{\Sigma }(𝒞,K),K)`$, where the functor $`G:Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}Sp^\mathrm{\Sigma }(𝒞,K)`$ used to form $``$ is defined by $`GX=XK`$ and is part of the $`𝒟`$-model structure of $`Sp^\mathrm{\Sigma }(𝒞,K)`$ (see the comment following Theorem 7.11). This means that the structure map of $`G`$ involves the twist map. By Theorem 8.1, $`F_0:Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}`$ is a $`𝒟`$-Quillen equivalence. On the other hand, consider $`Sp^\mathrm{\Sigma }(Sp^{}(𝒞,G),K)`$, where now we use the action of $`𝒟`$ on $`Sp^{}(𝒞,G)`$ that comes from Theorem 5.7. As pointed out in Remark 5.8, this means that we are using the functor $`XXK`$ to form $`Sp^\mathrm{\Sigma }(Sp^{}(𝒞,G),K)`$, not the functor $`XX\overline{}K`$. By hypothesis, this functor is already a Quillen equivalence, so Theorem 5.1 implies that $`F_0:Sp^{}(𝒞,G)\stackrel{}{}Sp^\mathrm{\Sigma }(Sp^{}(𝒞,G),K)`$ is a $`𝒟`$-Quillen equivalence. It remains to prove that $``$ and $`Sp^\mathrm{\Sigma }(Sp^{}(𝒞,G),K)`$ are isomorphic as model categories. This is mostly a matter of unwinding definitions. An object of $``$ is a set $`\{Y_{m,n}\}`$ of objects of $`𝒞`$, where $`m,n0`$. There is an action of $`\mathrm{\Sigma }_n`$ on $`Y_{m,n}`$, and there are $`\mathrm{\Sigma }_n`$-equivariant maps $`Y_{m,n}K\stackrel{𝜈}{}Y_{m+1,n}`$ and $`Y_{m,n}K\stackrel{𝜌}{}Y_{m,n+1}`$. In addition, the composite $`Y_{m,n}K^p\stackrel{}{}Y_{m,n+p}`$ is $`\mathrm{\Sigma }_n\times \mathrm{\Sigma }_p`$-equivariant, and there is a compatibility between $`\nu `$ and $`\rho `$, expressed in the commutativity of the following diagram. $$\begin{array}{ccccc}Y_{m,n}KK& \stackrel{1T}{}& Y_{m,n}KK& \stackrel{\rho 1}{}& Y_{m,n+1}K\\ \nu 1& & & & \nu & & \\ Y_{m+1,n}K& =& Y_{m+1,n}K& \underset{\rho }{}& Y_{m+1,n+1}\end{array}$$ An object $`\{Z_{m,n}\}`$ of the category $`Sp^\mathrm{\Sigma }(Sp^{}(𝒞,G),K)`$ has the same description if we switch $`m`$ and $`n`$. There is then an isomorphism of categories between $``$ and $`Sp^\mathrm{\Sigma }(Sp^{}(𝒞,G),K)`$ which simply switches $`m`$ and $`n`$. The model structures are also the same. Indeed, they are both the localization of the evident bigraded projective model structure with respect to the maps $`F_{m,n+1}(QCK)\stackrel{}{}F_{m,n}QC`$ and $`F_{m+1,n}(QCK)\stackrel{}{}F_{m,n}QC`$, where $`F_{m,n}`$ is left adjoint to the evaluation functor $`\mathrm{Ev}_{m,n}`$. The natural isomorphism $`[\mathrm{Ho}Sp^\mathrm{\Sigma }(𝒞,K)](F_0,F_0B)[\mathrm{Ho}Sp^{}(𝒞,G)](F_0A,F_0B)`$ follows from the fact that the composites $$𝒞\stackrel{F_0}{}Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{F_0}{}$$ and $$𝒞\stackrel{F_0}{}Sp^{}(𝒞,G)\stackrel{F_0}{}Sp^\mathrm{\Sigma }(Sp^{}(𝒞,G),K)$$ are equal. ∎ In particular, we have calculated $`[\mathrm{Ho}Sp^{}(𝒞,G)](F_0A,F_0B)`$ in Corollary 4.10; when both the hypotheses of that corollary and the hypotheses of Theorem 9.1 hold, we get the expected result $$[\mathrm{Ho}Sp^\mathrm{\Sigma }(𝒞,K)](F_0A,F_0B)=\mathrm{colim}\mathrm{Ho}𝒞(AK^n,BK^n)$$ for cofibrant $`A`$ and $`B`$. Theorem 9.1 indicates that we should try to prove that $`K`$ is a Quillen equivalence of $`Sp^{}(𝒞,G)`$. The only reasonable way to do this is by comparing this functor to $`\overline{}K`$, which we know is a Quillen equivalence. The basic idea of the proof is to compare $`XKK`$ to $`X\overline{}K\overline{}K`$. Both of these spectra have the same spaces, and their structure maps differ precisely by the cyclic permutation self-map of $`KKK`$. So if we knew that this map were the identity, they would be the same spectra. The hope is then that, if we know that the cyclic permutation map is homotopic to the identity, these two spectra are equivalent in $`\mathrm{Ho}Sp^{}(𝒞,G)`$. One can in fact construct a map of spectra $`X\overline{}K\overline{}K\stackrel{}{}R(XKK)`$, where $`R`$ is a level fibrant replacement functor and $`X`$ is cofibrant, by inductively modifying the identity map. Unfortunately, the author does not know how to do this modification in a natural way, so is unable to prove that the derived functors $`L(XKK)`$ and $`L(X\overline{}K\overline{}K)`$ are equivalent in this way. Instead, we will follow a suggestion of Dan Dugger and construct a new functor $`F`$ on cofibrant objects of $`Sp^{}(𝒞,G)`$ and natural level equivalences $`FX\stackrel{}{}X\overline{}K\overline{}K`$ and $`FX\stackrel{}{}XKK`$, for cofibrant $`X`$. It will follow immediately that $`L(X\overline{}K\overline{}K)`$ is naturally equivalent to $`L(XKK)`$, and so that $`XXK`$ is a Quillen equivalence on $`Sp^{}(𝒞,G)`$. Unfortunately, to make this construction we will need to make some unpleasant assumptions that ought to be unnecessary. ###### Definition 9.2. Given a symmetric monoidal model category $`𝒟`$ whose unit $`S`$ is cofibrant, a *unit interval* in $`𝒟`$ is a cylinder object $`I`$ for $`S`$ such that there exists a map $`H_I:II\stackrel{}{}I`$ satisfying $`H_I(1i_0)=H_I(i_01)=i_0\pi `$ and $`H_I(1i_1)`$ is the identity. Here $`i_0,i_1:S\stackrel{}{}I`$ and $`\pi :I\stackrel{}{}S`$ are the structure maps of $`I`$. Given a cofibrant object $`K`$ of $`𝒟`$, we say that $`K`$ is *symmetric* if there is a unit interval $`I`$ and a homotopy $$H:KKKI\stackrel{}{}KKK$$ from the cyclic permutation to the identity. Note that $`[0,1]`$ is a unit interval in the usual model structure on compactly generated topological spaces, and $`\mathrm{\Delta }[1]`$ is a unit interval in the category of topological spaces. Indeed, the required map $`H_1:\mathrm{\Delta }[1]\times \mathrm{\Delta }[1]`$ takes both of the nondegenerate $`2`$-simplices $`011\times 001`$ and $`001\times 011`$ to $`001`$. Similarly, the standard unit interval chain complex of abelian groups is a unit interval in the projective model structure on chain complexes. Also, any symmetric monoidal left Quillen functor preserves unit intervals. It follows, for example, that the unstable $`A^1`$-model category of Morel-Voevodksy has a unit interval. Our goal, then, is to prove the following theorem. ###### Theorem 9.3. Suppose $`𝒟`$ is a symmetric monoidal model category with cofibrant unit $`S`$, and $`𝒞`$ is a left proper cellular $`𝒟`$-model category. Suppose that $`K`$ is a cofibrant object of $`𝒟`$, and that either $`K`$ is itself symmetric or the domains of the generating cofibrations of $`𝒞`$ are cofibrant and $`K`$ is weakly equivalent to a symmetric object of $`𝒟`$. Then the functor $`XXK`$ is a Quillen equivalence of $`Sp^{}(𝒞,G)`$. This theorem is certainly not the best one can do. For example, by considering the analogous functors with more than three tensor factors of $`K`$, it should be possible to show that the same theorem holds if there is a left homotopy between some even permutation of $`K^n`$ and the identity. Also, it seems clear that one should only need the cyclic permutation, or more generally some even permutation, to be equal to the identity in $`\mathrm{Ho}𝒟`$. That is, we should not need a specific left homotopy. But the author does not know how to remove this hypothesis. In any case, we have the following corollary. ###### Corollary 9.4. Suppose $`𝒟`$ is a left proper cellular symmetric monoidal model category whose unit $`S`$ is cofibrant, and whose generating cofibrations can be taken to have cofibrant domains. Suppose $`𝒞`$ is a left proper $`𝒟`$-model category. Suppose $`K`$ is a cofibrant object of $`𝒟`$, and either that $`K`$ is itself symmetric, or else that the domains of the generating cofibrations of $`𝒞`$ are cofibrant and $`K`$ is weakly equivalent to a symmetric object of $`𝒟`$. Then there is a $`𝒟`$-model category $``$ and $`𝒟`$-Quillen equivalences $$Sp^\mathrm{\Sigma }(𝒞,K)\stackrel{}{}\stackrel{}{}Sp^{}(𝒞,G).$$ We will prove Theorem 9.3 in a series of lemmas. ###### Lemma 9.5. Suppose $`𝒟`$ is a symmetric monoidal model category whose unit $`S`$ is cofibrant. Suppose we have a square $$\begin{array}{ccc}A& \stackrel{f}{}& X\\ r& & s& & \\ B& \underset{g}{}& Y\end{array}$$ in a $`𝒟`$-model category $`𝒞`$, where $`A`$ and $`B`$ are cofibrant, and a left homotopy $`H:AI\stackrel{}{}Y`$ from $`gr`$ to $`sf`$, for some unit interval $`I`$. Then there is an object $`B^{}`$ of $`𝒞`$, a weak equivalence $`B^{}\stackrel{𝑞}{}B`$, a commutative square $$\begin{array}{ccc}A& \stackrel{f}{}& X\\ r^{}& & s& & \\ B^{}& \underset{g^{}}{}& Y\end{array}$$ such that $`qr^{}=r`$, and a left homotopy $`H^{}:B^{}I\stackrel{}{}Y`$ between $`gq`$ and $`g^{}`$. Furthermore, this construction is natural in an appropriate sense. Naturality means that, if we have a map of such homotopy commutative squares that preserves the homotopies, then we get a map of the resulting commutative squares that preserves the maps $`q`$ and $`H^{}`$. The precise statement is complicated, and we leave it to the reader. ###### Proof. We let $`B^{}`$ be the mapping cylinder of $`r`$. That is, we take $`B^{}`$ to be the pushout in the diagram below. $$\begin{array}{ccc}A& \stackrel{i_0}{}& AI\\ r& & h& & \\ B& \underset{j}{}& B^{}\end{array}$$ The map $`r^{}`$ is then the composite $`hi_1`$, and the map $`g^{}`$ is the map that is $`g`$ on $`B`$ and $`H`$ on $`AI`$. It follows that $`g^{}r^{}=Hi_1=sf`$, as required. The map $`q:B^{}\stackrel{}{}B`$ is defined to be the identity on $`B`$ and the composite $`AI\stackrel{𝜋}{}A\stackrel{𝑟}{}B`$ on $`AI`$. Since $`j`$ is a trivial cofibration (as a pushout of $`i_0`$), it follows that $`q`$ is a weak equivalence, and it is clear that $`qr^{}=r`$. We must now construct the homotopy $`H^{}`$. First note that $`B^{}`$ is cofibrant, since $`B`$ is so and $`j`$ is a trivial cofibration, and so $`B^{}I`$ is a cylinder object for $`B^{}`$. In fact, $`B^{}I`$ is the pushout of $`AII`$ and $`BI`$ over $`AI`$. Define $`H^{}`$ to be the constant homotopy $`BI\stackrel{𝜋}{}B\stackrel{𝑔}{}Y`$ on $`BI`$ and the homotopy $$AII\stackrel{1H_I}{}AI\stackrel{𝐻}{}Y$$ on $`AII`$. The fact that $`H_I(i_01)=i_0\pi `$ guarantees that $`H^{}`$ is well-defined, and the other conditions on $`H_I`$ guarantee that $`H^{}`$ is a left homotopy from $`qg`$ to $`g^{}`$. We leave the naturality of this construction to the reader. ∎ We also need the following lemma about the behavior of unit intervals. ###### Lemma 9.6. Suppose $`𝒟`$ is a symmetric monoidal model category whose unit $`S`$ is cofibrant. Let $`I`$ and $`I^{}`$ be unit intervals, and define $`J`$ by the pushout diagram below. $$\begin{array}{ccc}S& \stackrel{i_0}{}& I^{}\\ i_1& & j_1& & \\ I& \underset{j_0}{}& J\end{array}$$ Then $`J`$ is a unit interval. ###### Proof. The reader is well-advised to draw a picture in the topological or simplicial case, from which the proof should be clear. We think of $`J`$ as the interval whose left half is $`I`$ and whose right half is $`I^{}`$. In particular, $`J`$ is a cylinder object for $`S`$, where $`i_0^{}=j_0i_0`$ and $`i_1^{}=j_1i_1`$. Then, because the tensor product preserves pushouts, we can think of $`JJ`$ as a square consisting of a copy of $`II`$ in the lower left, a copy of $`II^{}`$ in the upper left, a copy of $`I^{}I`$ is the lower right, and a copy of $`I^{}I^{}`$ in the upper right. We define the necessary map $`G:JJ\stackrel{}{}J`$ by defining $`G`$ on each subsquare. On the lower left, we use the composite $`II\stackrel{𝐻}{}I\stackrel{j_0}{}J`$, where $`H`$ is the homotopy making $`I`$ into a unit interval. Similarly, on the upper right square, we use the composite $`j_1H^{}`$. On the upper left square we use the constant homotopy $`j_0(1\pi )`$, and on the lower right square we use the constant homotopy $`j_0(\pi 1)`$. We leave it to the reader to check that this makes $`J`$ into a unit interval. ∎ The importance of these two lemmas for spectra is indicated in the following consequence. ###### Lemma 9.7. Suppose $`𝒟`$ is a left proper cellular symmetric monoidal model category with a unit interval $`I`$, whose unit $`S`$ is cofibrant. Let $`K`$ be a cofibrant object of $`𝒟`$, and let $`𝒞`$ be a left proper cellular $`𝒟`$-model category. Suppose $`A,BSp^{}(𝒞,G)`$, where $`A`$ is cofibrant, and we have maps $`f_n:A_n\stackrel{}{}B_n`$ for all $`n`$ and a homotopy $`H_n:A_nKI\stackrel{}{}B_{n+1}`$ from $`f_{n+1}\sigma _A`$ to $`\sigma _B(f_n1)`$, where $`\sigma _{()}`$ is the structure map of the spectrum $`()`$. Then there is a spectrum $`C`$, a level equivalence $`C\stackrel{}{}A`$, and a map of spectra $`C\stackrel{𝑔}{}B`$ such that $`g_n`$ is homotopic to $`f_nh_n`$. Furthermore, this construction is natural in an appropriate sense. Once again, the naturality involves the homotopies $`H_n`$ as well as the maps $`f_n`$. We leave the precise statement to the reader. ###### Proof. We define $`C_n`$, $`h_n`$, $`g_n`$ and a homotopy $`H_n^{}:C_nI_n\stackrel{}{}B_n`$ from $`g_n`$ to $`f_nh_n`$, where $`I_n`$ is a unit interval, inductively on $`n`$, using Lemma 9.5. To get started, we take $`C_0=A_0`$, $`h_0`$ to be the identity, $`g_0`$ to be $`f_0`$, and $`H_0^{}`$ to be the constant homotopy (with $`I_0=I`$). For the inductive step, we apply Lemma 9.5 to the diagram $$\begin{array}{ccc}C_nK& \stackrel{g_n1}{}& B_nK\\ \sigma \left(h_n1\right)& & \sigma & & \\ A_{n+1}& \underset{f_{n+1}}{}& B_{n+1}\end{array}$$ and the homotopy obtained as follows. We have a homotopy $$C_nKI_n\stackrel{1T}{}C_nI_nK\stackrel{H_n^{}1}{}B_nK\stackrel{𝜎}{}B_{n+1}$$ from $`\sigma (f_n1)(h_n1)`$ to $`\sigma (g_n1)`$. On the other hand, we also have the homotopy $`H_n(h_n1)`$ from $`f_{n+1}\sigma (h_n1)`$ to $`\sigma (f_n1)(h_n1)`$. We can amalgamate these to get a homotopy $`G_n:C_nKI_{n+1}\stackrel{}{}B_{n+1}`$ from $`f_{n+1}\sigma (h_n1)`$ to $`\sigma (g_n1)`$, and $`I_{n+1}`$ is still a unit interval by Lemma 9.6. Hence Lemma 9.5 gives us an object $`C_{n+1}`$, a map $`\sigma :C_nK\stackrel{}{}C_{n+1}`$, and a map $`g_{n+1}:C_{n+1}\stackrel{}{}B_{n+1}`$ such that $`g_{n+1}\sigma =\sigma (g_n1)`$. It also gives us a map $`h_{n+1}:C_{n+1}\stackrel{}{}A_{n+1}`$ such that $`h_{n+1}\sigma =\sigma (h_n1)`$ and a homotopy $`H_{n+1}^{}:C_{n+1}I_{n+1}\stackrel{}{}B_{n+1}`$ from $`f_{n+1}h_{n+1}`$ to $`g_{n+1}`$. This completes the induction step and the proof (we leave naturality to the reader). ∎ With this lemma in had we can now give the proof of Theorem 9.3. ###### Proof of Theorem 9.3. We first reduce to the case where $`K`$ is itself symmetric. So suppose the generating cofibrations of $`𝒞`$ have cofibrant domains, and suppose $`K^{}`$ is symmetric and weakly equivalent to $`K`$; this means there are weak equivalences $`K\stackrel{}{}RK\stackrel{}{}RK^{}\stackrel{}{}K^{}`$, where $`R`$ denotes a fibrant replacement functor. This means the total left derived functors $`XX^LK`$ and $`XX^LK^{}`$ are naturally isomorphic on the homotopy category of any $`𝒟`$-model category. In particular, it suffices to show that $`XXK^{}`$ is a Quillen equivalence on $`Sp^{}(𝒞,K)`$. On the other hand, by Theorem 5.5, there are $`𝒟`$-Quillen equivalences $$Sp^{}(𝒞,K)\stackrel{}{}Sp^{}(𝒞,RK)\stackrel{}{}Sp^{}(𝒞,RK^{})\stackrel{}{}Sp^{}(𝒞,K^{}).$$ It therefore suffices to show that $`XXK^{}`$ is a Quillen equivalence of $`Sp^{}(𝒞,K^{})`$; that is, we can assume that $`K`$ itself is symmetric. Let $`H`$ denote the given homotopy from the cyclic permutation to the identity of $`KKK`$. Let $`X`$ be a cofibrant spectrum, let $`\stackrel{~}{\sigma }`$ denote the structure map of $`X\overline{}K\overline{}K`$, and let $`\sigma `$ denote the structure map of $`XKK`$. These two structure maps differ by the cyclic permutation, and therefore we are in the situation of Lemma 9.7, with $`A=X\overline{}K\overline{}K`$, $`B=XKK`$, $`f_n`$ equal to the identity map, and $`H_n=(\sigma _X11)(X_{n+1}H)`$. It follows that we get a functor $`F`$ defined on cofibrant objects of $`Sp^{}(𝒞,G)`$ and natural level equivalences $`FX\stackrel{}{}X\overline{}K\overline{}K`$ and $`FX\stackrel{𝑔}{}XKK`$, where the latter map is a level equivalence since $`g_n`$ is homotopic to $`h_n`$. Thus the total left derived functors of $`()\overline{}K\overline{}K`$ and $`()KK`$ are naturally isomorphic. Since we know already that $`()\overline{}K\overline{}K`$ is a Quillen equivalence, so is $`()KK`$, and hence so is $`()K`$. ∎ ## Appendix A Cellular model categories In this section we define cellular model categories and show that the projective model structures on $`Sp^{}(𝒞,G)`$ and $`Sp^\mathrm{\Sigma }(𝒞,K)`$ are cellular when $`𝒞`$ is so. This is necessary to be sure that the Bousfield localizations used in the paper do in fact exist. The definitions in this section are taken from . Throughout this section, then, $`G`$ will be a left Quillen endofunctor of $`𝒞`$; when we refer to $`Sp^\mathrm{\Sigma }(𝒞,K)`$, we will be thinking of $`𝒞`$ as a $`𝒟`$-model category, where $`𝒟`$ is some symmetric monoidal model category, and of $`K`$ as a cofibrant object of $`𝒟`$. A cellular model category is a special kind of cofibrantly generated model category. Three additional hypotheses are needed. ###### Definition A.1. A model category $`𝒞`$ is *cellular* if there is a set of cofibrations $`I`$ and a set of trivial cofibrations $`J`$ making $`𝒞`$ into a cofibrantly generated model category and also satisfying the following conditions. 1. The domains and codomains of $`I`$ are compact relative to $`I`$. 2. The domains of $`J`$ are small relative to the cofibrations. 3. Cofibrations are effective monomorphisms. The first hypothesis above requires considerable explanation, which we will provide below. We first point out that the second hypothesis will hold in the projective model structure on $`Sp^{}(𝒞,G)`$ or $`Sp^\mathrm{\Sigma }(𝒞,K)`$ when it holds in $`𝒞`$. ###### Lemma A.2. Suppose $`𝒞`$ is a cofibrantly generated model category with generating cofibrations $`I`$ and generating trivial cofibrations $`J`$, and $`G`$ is a left Quillen endofunctor of $`𝒞`$. If the domains of $`J`$ are small relative to the cofibrations in $`𝒞`$, then the domains of the generating trivial cofibrations $`J_G`$ of the projective model structure on $`Sp^{}(𝒞,G)`$ ($`Sp^\mathrm{\Sigma }(𝒞,K)`$) are small relative to the cofibrations in $`Sp^{}(𝒞,G)`$ ($`Sp^\mathrm{\Sigma }(𝒞,K)`$). ###### Proof. For $`Sp^{}(𝒞,G)`$, this follows immediately from the definition of $`J_G`$, Proposition 1.10, and Proposition 1.12. The proof for $`Sp^\mathrm{\Sigma }(𝒞,K)`$ is similar. ∎ We now discuss the third hypothesis. ###### Definition A.3. Suppose $`𝒞`$ is a category. A map $`f:X\stackrel{}{}Y`$ is an *effective monomorphism* if $`f`$ is the equalizer of the two obvious maps $`YY_XY`$. ###### Proposition A.4. Suppose $`𝒞`$ is a cofibrantly generated model category and $`G`$ is a left Quillen endofunctor of $`𝒞`$. If cofibrations are effective monomorphisms in $`𝒞`$, then level cofibrations, and in particular projective cofibrations, are effective monomorphisms in $`Sp^{}(𝒞,G)`$ and in $`Sp^\mathrm{\Sigma }(𝒞,K)`$. ###### Proof. This is immediate, since limits in $`Sp^{}(𝒞,G)`$ and $`Sp^\mathrm{\Sigma }(𝒞,K)`$ are taken levelwise. ∎ We must now define compactness. This will involve some preliminary definitions. These definitions are extremely technical; the reader is advised to keep the example of CW-complexes in mind. ###### Definition A.5. Suppose $`I`$ is a set of maps in a cocomplete category. A *relative $`I`$-cell complex* is a map which can be written as the transfinite composition of pushouts of coproducts of maps of $`I`$. That is, given a relative $`I`$-cell complex $`f`$, there is an ordinal $`\lambda `$ and a $`\lambda `$-sequence $`X:\lambda \stackrel{}{}𝒞`$ and a collection $`\{(T^\beta ,e^\beta ,h^\beta )_{\beta <\lambda }\}`$ satisfying the following properties. 1. $`f`$ is isomorphic to the transfinite composition of $`X`$. 2. Each $`T^\beta `$ is a set. 3. Each $`e^\beta `$ is a function $`e^\beta :T^\beta \stackrel{}{}I`$. 4. Given $`\beta <\lambda `$ and $`iT^\beta `$, if $`e_i^\beta :C_i\stackrel{}{}D_i`$ is the image of $`i`$ under $`e^\beta `$, then $`h_i^\beta `$ is a map $`h_i^\beta :C_i\stackrel{}{}X_\beta `$. 5. Each $`X_{\beta +1}`$ is the pushout in the diagram $$\begin{array}{ccc}_{T^\beta }C_i& & _{T^\beta }D_i\\ h_i^\beta & & & & \\ X_\beta & & X_{\beta +1}\end{array}$$ The ordinal $`\lambda `$ together with the $`\lambda `$-sequence $`X`$ and the collection $`\{(T^\beta ,e^\beta ,h^\beta )_{\beta <\lambda }\}`$ is called a *presentation* of $`f`$. The set $`_\beta T_\beta `$ is the *set of cells* of $`f`$, and given a cell $`e`$, its *presentation ordinal* is the ordinal $`\beta `$ such that $`eT^\beta `$. The *presentation ordinal* of $`f`$ is $`\lambda `$. We also need to define subcomplexes of relative $`I`$-cell complexes. ###### Definition A.6. Suppoe $`𝒞`$ is a cocomplete category and $`I`$ is a set of maps in $`𝒞`$. Given a presentation $`\lambda `$, $`X:\lambda \stackrel{}{}𝒞`$, and $`\{(T^\beta ,e^\beta ,h^\beta )_{\beta <\lambda }\}`$ of a map $`f`$ as a relative $`I`$-cell complex, a *subcomplex* of $`f`$ (or really of the presentation of $`f`$), is a collection $`\{(\stackrel{~}{T}^\beta ,\stackrel{~}{e}^\beta ,\stackrel{~}{h}^\beta )_{\beta <\lambda }\}`$ such that the following properties hold. 1. Every $`\stackrel{~}{T}^\beta `$ is a subset of $`T^\beta `$, and $`\stackrel{~}{e}^\beta `$ is the restriction of $`e^\beta `$ to $`\stackrel{~}{T}^\beta `$. 2. There is a $`\lambda `$-sequence $`\stackrel{~}{X}:\lambda \stackrel{}{}𝒞`$ such that $`\stackrel{~}{X}_0=X_0`$ and a map of $`\lambda `$-sequences $`\stackrel{~}{X}\stackrel{}{}X`$ such that, for every $`\beta <\lambda `$ and $`i\stackrel{~}{T}^\beta `$, the map $`\stackrel{~}{h}_i^\beta :C_i\stackrel{}{}\stackrel{~}{X}_\beta `$ is a factorization of $`h_i^\beta :C_i\stackrel{}{}X_\beta `$ through the map $`\stackrel{~}{X}_\beta \stackrel{}{}X_\beta `$. 3. Every $`X_{\beta +1}`$ is the pushout in the diagram $$\begin{array}{ccc}_{\stackrel{~}{T}^\beta }C_i& & _{\stackrel{~}{T}^\beta }D_i\\ \stackrel{~}{h}_i^\beta & & & & \\ \stackrel{~}{X}_\beta & & \stackrel{~}{X}_{\beta +1}\end{array}$$ Given a subcomplex of $`f`$, the *size* of that subcomplex is the cardinality of its set of cells $`_{\beta <\lambda }\stackrel{~}{T}^\beta `$. Usually, $`I`$ will be a set of cofibrations in a model category where the cofibrations are essential monomorphisms. This condition guarantees that a subcomplex is uniquely determined by its set of cells \[9, Proposition 12.5.9\]. Of course, every set of cells does not give rise to a subcomplex. We can now define compactness. ###### Definition A.7. Suppose $`𝒞`$ is a cocomplete category and $`I`$ is a set of maps in $`𝒞`$. 1. Given a cardinal $`\kappa `$, an object $`X`$ is *$`\kappa `$-compact relative to $`I`$* if, for every relative $`I`$-cell complex $`f:Y\stackrel{}{}Z`$ and for every presentation of $`f`$, every map $`X\stackrel{}{}Z`$ factors through a subcomplex of size at most $`\kappa `$. 2. An object $`X`$ is *compact relative to $`I`$* if $`X`$ is $`\kappa `$-compact relative to $`I`$ for some cardinal $`\kappa `$. The following proposition is adapted from an argument of Phil Hirschhorn’s. ###### Proposition A.8. Suppose $`𝒞`$ is a cellular model category with generating cofibrations $`I`$, and $`G`$ is a left Quillen endofunctor of $`𝒞`$. Let $`A`$ be a domain or codomain of $`I`$. Then $`F_nA`$ is compact relative to $`I_G`$ in $`Sp^{}(𝒞,G)`$ or $`Sp^\mathrm{\Sigma }(𝒞,K)`$. ###### Proof. We will prove the proposition only for $`Sp^{}(𝒞,G)`$, as the $`Sp^\mathrm{\Sigma }(𝒞,K)`$ case is similar. Throughout this proof we will use Proposition A.4, which guarantees that subcomplexes in $`𝒞_G`$ are determined by their cells. Choose an infinite cardinal $`\gamma `$ such that the domains and codomains of $`I`$ are all $`\gamma `$-compact relative to $`I`$. When dealing with relative $`I`$-cell complexes, we can assume that we have a presentation as a transfinite composition of pushouts of maps of $`I`$, rather than as a transfinite composition of pushouts of coproducts of maps of $`I`$, using \[10, Lemma 2.1.13\] or \[9, Section 12.2\]. A similar comment holds for relative $`I_G`$-cell complexes. We will proceed by transfinite induction on $`\beta `$, where the induction hypothesis is that for every presented relative $`I_G`$-cell complex $`f:X\stackrel{}{}Y`$ whose presentation ordinal is $`\beta `$, and for every map $`F_nA\stackrel{𝑓}{}Y`$ where $`n`$ is an integer and $`A`$ is a domain or codomain of $`I`$, $`f`$ factors through a subcomplex with at most $`\gamma `$ $`I_G`$-cells. Getting the induction started is easy. For the induction step, suppose the induction hypothesis holds for all ordinals $`\alpha <\beta `$, and suppose we have a presentation $$X=X^0\stackrel{}{}X^1\stackrel{}{}\mathrm{}\stackrel{}{}X^\alpha \stackrel{}{}\mathrm{}X^\beta =Y$$ of $`f:X\stackrel{}{}Y`$ as a transfinite composition of pushouts of maps of $`I_G`$. Then the boundary of each $`I_G`$-cell of this presentation is represented by a map $`F_mC\stackrel{}{}X^\alpha `$, for some $`\alpha <\beta `$, some $`m0`$, and some domain $`C`$ of a map of $`I`$. This map factors through a subcomplex with at most $`\gamma `$ $`I_G`$-cells, by induction. It follows that the $`I_G`$-cell itself is contained in a subcomplex of at most $`\gamma `$ $`I_G`$-cells, since we can just attach the interior of the $`I_G`$-cell to the given subcomplex. Now suppose we have an arbitrary map $`F_nA\stackrel{}{}Y`$, where $`A`$ is a domain or codomain of a map of $`I`$. Such a map is determined by a map $`A\stackrel{}{}Y_n`$ in $`𝒞`$. The map $`f_n:X_n\stackrel{}{}Y_n`$ is the transfinite composition of the cofibrations $`X_n^\alpha \stackrel{}{}X_n^{\alpha +1}`$. For each $`\alpha `$, there is an $`m`$ and a map $`h`$ of $`I`$ such that $`X_n^\alpha \stackrel{}{}X_n^{\alpha +1}`$ is the pushout of $`G^mh`$, where we interpret $`G^mh`$ as the identity map if $`m`$ is negative. Apply \[9, Lemma 12.4.19\] to write the $`\beta `$-sequence $`X_n^\alpha `$ as a retract of a $`\beta `$-sequence $$X_n=Z^0\stackrel{}{}Z^1\stackrel{}{}\mathrm{}\stackrel{}{}Z^\alpha \stackrel{}{}\mathrm{}Z^\beta =Z$$ where each map $`Z^\alpha \stackrel{}{}Z^{\alpha +1}`$ is a relative $`I`$-cell complex. We denote the retraction by $`r:Z\stackrel{}{}Y_n`$, noting that the restriction of $`r`$ to $`Z^\alpha `$ factors (uniquely) through $`X_n^\alpha `$. We can think of the entire map $`X_n\stackrel{}{}Z`$ as a relative $`I`$-cell complex, each cell $`e`$ of which appears in the relative $`I`$-cell complex $`Z^{t(e)}\stackrel{}{}Z^{t(e)+1}`$ for some unique ordinal $`t(e)`$, and so has associated to it the $`I_G`$-cell $`c(e)`$ of $`f`$ used to form $`X^{t(e)}\stackrel{}{}X^{t(e)+1}`$. The composite $`A\stackrel{}{}Y_n\stackrel{}{}Z`$ then factors through a subcomplex $`V`$ with at most $`\gamma `$ $`I`$-cells. The proof will be completed if we can find a subcomplex $`W`$ of $`Y`$ with at most $`\gamma `$ $`I_G`$-cells such that the restriction of $`r`$ to $`V`$ factors through $`W_n`$. Take $`W`$ to be a subcomplex of $`Y`$ containing the $`I_G`$-cells $`c(e)`$ as $`e`$ runs through the cells of $`V`$. Then $`W_n`$ contains $`r(e)`$ for every cell of $`V`$, so $`W_n`$ contains $`rV`$, as required. Furthermore, since each $`I_G`$-cell $`c(e)`$ lies in a subcomplex with no more that $`\gamma `$ $`I_G`$-cells, and $`V`$ has no more than $`\gamma `$ cells, there is a choice for $`W`$ which has no more than $`\gamma ^2=\gamma `$ cells. This completes the induction step and the proof. ∎ Altogether then, we have the following theorem. ###### Theorem A.9. Suppose $`𝒞`$ is a left proper cellular model category, and $`G`$ is a left Quillen endofunctor on $`𝒞`$. Then the category $`Sp^{}(𝒞,G)`$ of $`G`$-spectra and the category $`Sp^\mathrm{\Sigma }(𝒞,K)`$ of symmetric spectra, with the projective model structures, are left proper cellular model categories.
warning/0004/physics0004062.html
ar5iv
text
# Testing cosmological variability of fundamental constants ## Introduction Contemporary theories (SUSY GUT, superstring and others) not only predict the dependence of fundamental physical constants on energy<sup>1</sup><sup>1</sup>1 The prediction of the theory that the fundamental constants depend on the energy of interaction has been confirmed in experiment. In this paper, we consider only the space-time variability of their low-energy limits., but also have cosmological solutions in which low-energy values of these constants vary with the cosmological time. The predicted variation at the present epoch is small but non-zero, and it depends on theoretical model. In particular, Damour and Polyakov Damour have developed a modern version of the string theory, whose parameters could be determined from cosmological variations of the coupling constants and hadron-to-electron mass ratios. Clearly, a discovery of these variations would be a great step in our understanding of Nature. Even a reliable upper bound on a possible variation rate of a fundamental constant presents a valuable tool for selecting viable theoretical models. Historically, a hypothesis that the fundamental constants may depend on the cosmological time $`t`$ (that is the age of the Universe) was first discussed by Milne Milne and Dirac Dirac . The latter author proposed his famous “large-number hypothesis” and suggested that the gravitational constant was directly proportional to $`t`$. Later the variability of fundamental constants was analyzed, using different arguments, by Gamow Gamow , Dyson Dyson , and others. The interest in the problem has been revived due to recent major achievements in GUT and Superstring models (e.g., Damour ). Presently, the fundamental constants are being measured with a relative error of $`10^8`$. These measurements obviously rule out considerable variations of the constants on a short time scale, but do not exclude their changes over the lifetime of the Universe, $`1.5\times 10^{10}`$ years. Moreover, one cannot rule out the possibility that the constants differ in widely separated regions of the Universe; this could be disproved only by astrophysical observations and different kinds of experiments. Laboratory experiments cannot trace possible variation of a fundamental constant during the entire history of the Universe. Fortunately, Nature has provided us with a tool for direct measuring the physical constants in the early epochs. This tool is based on observations of quasars, the most powerful sources of radiation. Many quasars belong to most distant objects we can observe. Light from the distant quasars travels to us about $`10^{10}`$ years. This means that the quasar spectra registered now were formed $`10^{10}`$ years ago. The wavelengths of the lines observed in these spectra ($`\lambda _{\mathrm{obs}}`$) increase compared to their laboratory values ($`\lambda _{\mathrm{lab}}`$) in proportion $`\lambda _{\mathrm{obs}}`$ = $`\lambda _{\mathrm{lab}}(1+z)`$, where the cosmological redshift $`z`$ can be used to determine the age of the Universe at the line-formation epoch. In some cases, the redshift is as high as $`z35`$, so that the intrinsically far-ultraviolet lines are registered in the visible range. The examples are demonstrated in Fig. 1. Analysing these spectra we may study the epoch when the Universe was several times younger than now. Here we review briefly the studies of the space-time variability of the fine-structure constant $`\alpha `$ and the proton-to-electron mass ratio $`\mu `$. ## Fine-structure constant Various tests of the fundamental constant variability differ in space-time regions of the Universe which they cover. Local tests relate to the values of constants on the Earth and in the Solar system. In particular, laboratory tests infer the possible variation of certain combinations of constants “here and now” from comparison of different frequency standards. Geophysical tests impose constraints on combinations of fundamental constants over the past history of the Solar system, although most of these constraints are very indirect. In contrast, astrophysical tests allows one to “measure” the values of fundamental constants in distant areas of the early Universe. ### Local tests #### Laboratory experiments There were a number of laboratory experimets aimed at detection of trends of the fundamental constants with time by comparison of frequency standards which have different dependences on the constants. We mention only two of the published experiments. Comparison of H-masers with Cs-clocks during 427 days revealed a relative (H–Cs) frequency drift with a rate $`1.5\times 10^{16}`$ per day, while the rates of (H-H) and (Cs–Cs) drifts (i.e., the drifts between identical standards, used to control their stability) were less than $`1\times 10^{16}`$ per day Demidov . A similar result was found in comparison of a Hg<sup>+</sup>-clock with a H-maser during 140 days Prestage : the rate of the relative frequency drift was less than $`(2\pm 1)\times 10^{16}`$ per day. Such a drift is treated as a consequence of a difference in the long-term stability of different atomic clocks. In principle, however, it may be caused by variation of $`\alpha `$. That is why it gives an upper limit to the $`\alpha `$ variation Prestage : $`|\dot{\alpha }/\alpha |3.7\times 10^{14}\mathrm{yr}^1`$. #### Geophysical tests The strongest bound on the possible time-variation rate of $`\alpha `$ was derived in 1976 by Shlyakhter Shlyakhter , and recently, from a more detailed analysis, by Damour and Dyson DD , who obtained $`|\dot{\alpha }/\alpha |<0.7\times 10^{16}\mathrm{yr}^1`$, The analysis was based on measurements of isotope ratios in the Oklo site in Gabon, where a unique natural uranium nuclear fission reactor had operated 1.8 billion years ago. The isotope ratios of samarium produced in this reactor by the neutron capture reaction <sup>149</sup>Sm$`+n^{150}`$Sm$`+\gamma `$ would be completely different, if the energy of the nuclear resonance responsible for this capture were shifted at least by 0.1 eV. Another strong bound, $`|\dot{\alpha }/\alpha |<5\times 10^{15}\mathrm{yr}^1`$, was obtained by Dyson Dyson from an isotopic analysis of natural radioactive decay products in meteorites. A weak point of these tests is their dependence on the model of the phenomenon, fairly complex, involving many physical effects. For instance, Damour and Dyson DD estimated possible shift of the above-mentioned resonance due to the $`\alpha `$ variation, assuming that the Coulomb energy of the excited state of <sup>150</sup>Sm, responsible for the resonance, is not less than the Coulomb energy of the ground state of <sup>150</sup>Sm. In absence of experimental data on the nuclear state in question, this assumption is not justified, since heavy excited nuclei often have Coulomb energies smaller than those for their ground states Kalvius . Furthermore, a correlation between the constants of strong and electroweak interactions (which is likely in the frame of modern theory) might lead to further softening of the mentioned bounds by 100-fold, to $`|\dot{\alpha }/\alpha |<5\times 10^{15}\mathrm{yr}^1`$, as noted by Sisterna and Vucetich Sisterna . In addition, the local tests cannot be extended to distant space regions and to the early Universe, since the law of possible space-time variation of $`\alpha `$ is unknown a priory. It is the extragalactic astronomy that allows us to study these remote regions of spacetime, in particular the regions which were causally disconnected at the epoch of formation of the observed absorption spectra. ### Astrophysical tests To find out whether $`\alpha `$ changed over the cosmological time, we have studied the fine splitting of the doublet lines of Si iv, C iv, Mg ii and other ions, observed in the spectra of distant quasars. According to quantum electrodynamics, the relative splitting of these lines $`\delta \lambda /\lambda `$ is proportional to $`\alpha ^2`$ (neglecting very small corrections). Consequently, if $`\alpha `$ changed with time, then $`\delta \lambda /\lambda `$ would depend on the cosmological redshift $`z`$. This method of measuring $`\alpha `$ in distant regions of the Universe had been first suggested by Savedoff Savedoff and was used later by other authors. For instance, Wolfe et al. Wolfe derived an estimate $`|\dot{\alpha }/\alpha |<4\times 10^{12}\mathrm{yr}^1`$ from an observation of the Mg ii absorption doublet at $`z=0.524`$. An approximate formula which relates a deviation of $`\alpha `$ at redshift $`z`$ from its current value, $`\mathrm{\Delta }\alpha _z`$, with measured $`\delta \lambda /\lambda `$ in the extragalactic spectra and in laboratory reads $$\mathrm{\Delta }\alpha _z\frac{c_r}{2}\left[\frac{(\delta \lambda /\lambda )_z}{(\delta \lambda /\lambda )_0}1\right],$$ (1) where $`c_r1`$ takes into account radiation corrections Dzuba : for instance, for Si iv $`c_r0.9`$. Many high-quality quasar spectra measured in the last decade have enabled us to significantly increase the accuracy of determination of $`\delta \lambda /\lambda `$ at large $`z`$. An example of the spectra observed is shown in Fig. 1. For the present report, we have selected the results of high-resolution observations Petitjean ; VPI ; Outram , most suitable for an analysis of $`\alpha `$ variation. The values of $`\mathrm{\Delta }\alpha /\alpha `$ calculated from these data according to Eq. (1) are given in Table 1. As a result, we obtain a new estimate of the possible deviation of the fine-structure constant at $`z=2`$–4 from its present ($`z=0`$) value: $$\mathrm{\Delta }\alpha /\alpha =(4.6\pm 4.3[\mathrm{stat}]\pm 1.4[\mathrm{syst}])\times 10^5,$$ (2) where the statistical error is obtained from the scatter of astronomical data (at large $`z`$) and the systematic one is estimated from the uncertainty of the fine splitting measurement in the laboratory Morton ; Kelly (at $`z=0`$, which serves as the reference point for the estimation of $`\mathrm{\Delta }\alpha `$). The corresponding upper limit of the $`\alpha `$ variation rate averaged over $`10^{10}`$ yr is $$|\dot{\alpha }/\alpha |<1.4\times 10^{14}\mathrm{yr}^1$$ (3) (at the 95% confidence level). This constraint is much more stringent than those obtained from all but one previous astronomical observations. The notable exception is presented by Webb et al. Webb , who have analysed spectroscopic data of similar quality, but estimated $`\alpha `$ from comparison of Fe ii and Mg ii fine-splitted walelengths in extragalactic spectra and in the laboratory. Their result indicates a tentative time-variation of $`\alpha `$: $`\mathrm{\Delta }\alpha /\alpha =(1.9\pm 0.5)\times 10^5`$ at $`z=1.0`$–1.6. Note, however, two important sources of a possible systematic error which could mimic the effect: (a) Fe,ii and Mg ii lines used are situated in different orders of the echelle-spectra, so relative shifts in calibration of the different orders can simulate the effect of $`\alpha `$-variation, and (b) were the relative abundances of Mg isotopes changing during the cosmological evolution, the Mg ii lines would be subjected to an additional $`z`$-dependent shift relative to the Fe ii lines, quite sufficient to simulate the variation of $`\alpha `$ (this shift can be easily estimated from recent laboratory measurements Pickering ). In contrast, the method based on the fine splitting of a line of the same ion species (Si iv in the above example) is not affected by these two uncertainty sources. Thus we believe that the restriction (3) is the most reliable at present for the long-term history of the Universe. According to our analysis, some theoretical models are inconsistent with observations. For example, power laws $`\alpha t^n`$ with $`n=1`$, $`1/4`$, and $`4/3`$, published by various authors in 1980s, are excluded. Moreover, the Teller–Dyson’s hypothesis on the logarithmic dependence of $`\alpha `$ on $`t`$ Teller ; Dyson has also been shown to be inconsistent with observations. Many regions of formation of the spectral lines, observed at large redshifts in different directions in the sky, had been causally disconnected at the epochs of line formation. Thus, no information could have been exchanged between these regions of the Universe and, in principle, the fundamental constants could be different there. However, a separate analysis SpSciRev has shown that $`\alpha `$ value is the same in different directions in the sky within the $`3\sigma `$ relative error $`|\mathrm{\Delta }\alpha /\alpha |<3\times 10^4`$. ## Proton-to-electron mass ratio The dimensionless constant $`\mu =m_\mathrm{p}/m_\mathrm{e}`$ approximately equals the ratio of the constants of strong interaction $`g^2/(\mathrm{}c)14`$ and electromagnetic interaction $`\alpha 1/137.036`$, where $`g`$ is the effective coupling constant calculated from the amplitude of nucleon–$`\pi `$-meson scattering at low energy. In order to check the cosmological variability of $`\mu `$ we have used high-redshift absorption lines of molecular hydrogen H<sub>2</sub> in the spectrum of the quasar PKS 0528–250. This is the first (and, in a sense, unique) high-redshift system of H<sub>2</sub> absorption lines discovered in 1985 LV85 . A study of these objects yields information of paramount importance on the physical conditions $`10^{10}`$ years ago. A possibility of distinguishing between the cosmological redshift of spectral wavelengths and shifts due to a variation of $`\mu `$ arises from the fact that the electronic, vibrational, and rotational energies of H<sub>2</sub> each undergo a different dependence on the reduced mass of the molecule. Hence comparing ratios of wavelengths $`\lambda _i`$ of various H<sub>2</sub> electron-vibration-rotational lines in a quasar spectrum at some redshift $`z`$ and in laboratory (at $`z=0`$), we can trace variation of $`\mu `$. The method had been used previously by Foltz et al. FCB , whose analysis was corrected later in our papers VL93 ; SpSciRev ; VP96 . In the latter papers, we calculated the sensitivity coefficients $`K_i`$ of the wavelengths $`\lambda _i`$ with respect to possible variation of $`\mu `$ and applied a linear regression analysis to the measured redshifts of individual lines $`z_i`$ as function of $`K_i`$. An illustration of the wavelength dependences on the mass of the nucleus is given in Table 2, where a few resonance wavelengths of hydrogen, deuterium, and tritium molecules are listed. One can see that, as the nuclear mass increases, different wavelengths shift in different directions. More complete tables, as well as two algorithms of $`K_i`$ calculation, are given in Refs. SpSciRev ; Lanzetta . Thus, if the proton mass in the epoch of line formation were different from the present value, the measured $`z_i`$ and $`K_i`$ values would correlate: $$\frac{z_i}{z_k}=\frac{(\lambda _i/\lambda _k)_z}{(\lambda _i/\lambda _k)_0}1+(K_iK_k)\left(\frac{\mathrm{\Delta }\mu }{\mu }\right).$$ (4) We have performed a $`z`$-to-$`K`$ regression analysis using a modern high-resolution spectrum of PKS 0528$``$250 Lanzetta . Several tens of the H<sub>2</sub> lines have been identified; a portion of the spectrum which reveales some of the lines is shown in Fig. 1. The redshift estimates for individual absorption lines with their individual errorbars are plotted in Fig. 2 against their sensitivity coefficients. The resulting parameter estimate and $`1\sigma `$ uncertainty is $$\mathrm{\Delta }\mu /\mu =(11.5\pm 7.6[\mathrm{stat}]\pm 1.9[\mathrm{syst}])\times 10^5.$$ (5) The $`2\sigma `$ confidence interval to $`\mathrm{\Delta }\mu /\mu `$ is $$|\mathrm{\Delta }\mu /\mu |<2.0\times 10^4.$$ (6) Assuming that the age of the Universe is $`15`$ Gyr the redshift of the H<sub>2</sub> absorption system $`z=2.81080`$ corresponds to the elapsed time of 13 Gyr (in the standard cosmological model). Therefore we arrive at the restriction $$|\dot{\mu }/\mu |<1.5\times 10^{14}\mathrm{yr}^1$$ (7) on the variation rate of $`\mu `$, averaged over 90% of the lifetime of the Universe. ## Conclusions Despite the theoretical prediction of the time-dependences of fundamental constants, a statistically significant variation of any of the constants have not been reliably detected up to date, according to our point of view substantiated above. The upper limits obtained indicate that the constants of electroweak and strong interactions did not significantly change over the last 90% of the history of the Universe. This shows that more precise measurements and observations and their accurate statistical analyses are required in order to detect the expected variations of the fundamental constants. Acknowledgements. This work was performed in frames of the Project 1.25 of the Russian State Program “Fundamental Metrology” and supported by the grant RFBR 99-02-18232.
warning/0004/astro-ph0004407.html
ar5iv
text
# Magnetospheric Scattering and Emission in Millisecond Pulsars ## 1 Introduction and Light Curve Model With the launch of CXO and XMM-Newton, there has been a dramatic increase in sensitivity for high resolution X-ray studies of compact objects. Rapidly rotating neutron stars are targets of particular interest, since measurements of the stellar surface can probe the neutron star evolution and the equation of state at high density. Several relativistic effects may also be visible, in some cases. A number of millisecond pulsars (MSP) have been observed in the X-rays, both as isolated rotation powered objects (Becker & Trümper, 1999) and as accretion powered LMXBs (e.g. Wijnands & van der Klis, 1998). In Braje, Romani, & Rauch (2000), we explored the distortions of the surface emission imparted by rapid rotation, including the important effects of Doppler boosting, aberration, and gravitational focusing and time delays. More subtle effects induced by frame dragging were also considered. For rotation-powered pulsars, the surface radiation must traverse the magnetosphere, where resonant scattering can introduce significant distortions in the pulse profile (Rajagopal & Romani, 1997). Further, several X-ray MSP have narrow, non-thermal pulse components, suggesting a direct origin in the magnetosphere. In this paper, we extend the treatment of rapid rotation effects to include scattering and emission in a surrounding dipole magnetosphere. Our model assumes a spherical star, which we take to be sufficiently centrally condensed that a Schwarzschild (or Kerr) metric is an adequate model of the external spacetime (the ‘Roche approximation’). Because we wish to extend the modeling to small spin periods $`P_{}`$, the effects of ‘sweep-back’ on the magnetic field, even for $`r`$ a few times the stellar mass $`M`$, can be substantial. In addition, the rapid rotation develops large electro-motive forces. We follow the common assumption that this EMF causes pair production such that the closed zone of the magnetosphere is filled with a charge-separated pair plasma, whose charge distribution cancels the rotational EMF. For the magnetic field, we assume a point dipole located at the stellar center and generalize the computation of the field structure for the non-rotating dipole in the Schwarzschild spacetime (e.g., Bardeen & Press, 1973; Prasanna & Gupta, 1997; Muslimov & Harding, 1997). Our result passes smoothly to the Schwarzschild magnetic field $`𝑩`$ as the stellar angular velocity $`\mathrm{\Omega }_{}`$ vanishes and to the flat space rotating dipole solution as $`r\mathrm{}`$. We assume that the plasma co-rotates (is stationary) in the closed zone, which is defined by the ‘last closed’ field lines traced from tangent approach to the light cylinder at $`r_{\mathrm{LC}}=cP_{}/2\pi `$ to the stellar surface. Our light curve modeling follows the procedure described in Braje et al. (2000). This Monte Carlo code starts with photons randomly drawn from the surface emission zone, with initial directions drawn from a model limb-darkened distribution. We aberrate these surface photons and propagate to infinity through curved space, using the Schwarzschild or Kerr metric, as appropriate. As the photon passes through the magnetosphere, we monitor for local cyclotron resonance. At the resonance position photons are re-emitted with the proper boosted scattering angular distribution. We also follow photons arising directly from the magnetosphere — from acceleration gaps or other non-thermal sources. After including the gravitational and time-of-flight delays, the photons are assigned to energy and rotational phase bins to produce maps of the radiation on the sky. Slices through these maps at the Earth’s viewing angle provide pulsar light curves and phase-resolved spectra. ## 2 Magnetic Field Structure We have developed an approximate, retarded, dipolar magnetic field expression that links the exact non-rotating Schwarzschild result at small $`r`$ with the flat space ‘swept-back’ field structure at large $`r`$. We expect the magnetic field lines to have approximately the same shape as a static dipole near the star, but to curve back near the light cylinder, as Figure 1 displays. To derive our magnetic field expression, we write Maxwell’s Stress Tensor as derivatives of the vector potential in covariant form as $$F_{\mu \nu }=\frac{A_\nu }{x^\mu }\frac{A_\mu }{x^\nu }.$$ (1) We use the Schwarzschild metric in geometrized units ($`G=c=1`$): $$\mathrm{d}s^2=\eta ^2\mathrm{d}t^2\frac{\mathrm{d}r^2}{\eta ^2}r^2\mathrm{d}\theta ^2r^2\mathrm{sin}^2\theta \mathrm{d}\varphi ^2$$ (2) where $`\eta =\sqrt{12M/r}`$. If we solve Maxwell’s Equations in vacuum $$F_{;\mu }^{\mu \nu }=0$$ (3) for a dipole moment $`𝝁=\mu \widehat{z}`$ we arrive at the solution in the Schwarzschild spacetime $$A_\varphi (r,\theta )=f(r)\frac{\mu \mathrm{sin}^2\theta }{r}$$ (4) where we have written the expression as the flat-space result weighted by the multiplicative function: $$f(r)=\frac{3r^3}{8M^3}\left[\mathrm{log}\eta ^2+\frac{2M}{r}\left(1+\frac{M}{r}\right)\right]$$ (5) (e.g., Wasserman & Shapiro, 1983). For the more general rotating case, we multiply the flat space time dependent vector potential by the same function $`f(r)`$: $$𝑨=f(r)\left\{\frac{𝝁(t)\mathbf{\times }\text{ }𝒓}{r^3}+\frac{\dot{𝝁}(t)\mathbf{\times }\text{ }𝒓}{r^2}\right\}$$ (6) where we now have a time dependent dipole moment, $`𝝁(t)=\mu (\mathrm{sin}\alpha \mathrm{cos}\mathrm{\Omega }_{}(tr),\mathrm{sin}\alpha \mathrm{sin}\mathrm{\Omega }_{}(tr),\mathrm{cos}\alpha )`$, tilted with respect to the rotational ($`z`$) axis by an angle $`\alpha `$. We use {$`𝒆_𝒕`$,$`𝒆_𝒓`$,$`𝒆_𝜽`$,$`𝒆_\mathit{\varphi }`$} to denote the coordinate basis vectors and {$`𝒆_{\widehat{𝒕}}`$,$`𝒆_{\widehat{𝒓}}`$,$`𝒆_{\widehat{𝜽}}`$,$`𝒆_{\widehat{\mathit{\varphi }}}`$} for the orthonormal basis vectors. By taking the dot product of the vector potential with the coordinate basis vectors, we obtain the vector potential components $`(A_t,A_r,A_\theta ,A_\varphi )`$. Using $$\begin{array}{ccccc}B_{\widehat{r}}& =& F_{\widehat{\varphi }\widehat{\theta }}& =& \frac{1}{r^2\mathrm{sin}\theta }F_{\varphi \theta }\hfill \\ B_{\widehat{\theta }}& =& F_{\widehat{r}\widehat{\varphi }}& =& \frac{\eta }{r\mathrm{sin}\theta }F_{r\varphi }\hfill \\ B_{\widehat{\varphi }}& =& F_{\widehat{\theta }\widehat{r}}& =& \frac{\eta }{r}F_{\theta r}\hfill \end{array}$$ (7) (c.f. Muslimov & Harding, 1997) and the definition of Maxwell’s Stress Tensor (eqn. 1), we derive an expression for the field: $$\begin{array}{ccc}B_{\widehat{r}}& =& \frac{6\mu }{(2M)^3}\chi [\mathrm{cos}\alpha \mathrm{cos}\theta \hfill \\ & & +\mathrm{sin}\alpha \mathrm{sin}\theta (\mathrm{cos}\stackrel{~}{\varphi }+r\mathrm{\Omega }_{}\mathrm{sin}\stackrel{~}{\varphi })]\hfill \\ B_{\widehat{\theta }}& =& \frac{3\mu \eta }{(2M)^3}[r^2\mathrm{\Omega }_{}^2\chi \mathrm{cos}\theta \mathrm{cos}\stackrel{~}{\varphi }\mathrm{sin}\alpha \hfill \\ & & +2(\mathrm{log}\eta ^2+\frac{x(1x/2)}{\eta ^2})\{\mathrm{cos}\alpha \mathrm{sin}\theta \hfill \\ & & +\mathrm{cos}\theta \mathrm{sin}\alpha (\mathrm{cos}\stackrel{~}{\varphi }+r\mathrm{\Omega }_{}\mathrm{sin}\stackrel{~}{\varphi })\}]\hfill \\ B_{\widehat{\varphi }}& =& \frac{3\mu \eta \mathrm{sin}\alpha }{(2M)^3}[r^2\mathrm{\Omega }_{}^2\chi \mathrm{sin}\stackrel{~}{\varphi }\hfill \\ & & +(\frac{x(2x)}{\eta ^2}+2\mathrm{log}\eta ^2)(r\mathrm{\Omega }_{}\mathrm{cos}\stackrel{~}{\varphi }\mathrm{sin}\stackrel{~}{\varphi })]\hfill \end{array}$$ (8) with $`x=2M/r`$, $`\chi =\mathrm{log}\eta ^2+x(1+x/2)`$, and $`\stackrel{~}{\varphi }=\varphi \mathrm{\Omega }_{}(tr)`$. The above expressions recover the familiar results in the appropriate limits: for $`M0`$ we obtain the flat space retarded fields and for $`\mathrm{\Omega }_{}0`$ we get the Schwarzschild expression for a dipole tilted with an angle $`\alpha `$ with respect to the $`z`$ axis. Checking Maxwell’s Equations (3), we find that two are exactly satisfied ($`\nu =1,4`$). The other two are nonzero with leading terms of order $`𝒪[(\mathrm{\Omega }_{}r/c)^2(GM/rc^2)]`$. In Figure 1, we display the field line differences induced at small $`r`$ in curved space coordinates. We have drawn field lines from the stellar surface in the $`𝛀_{\mathbf{}}`$-$`𝝁`$ plane. Note that the field lines are pulled closer to the star by curved space distortions. The open field lines are nearly radial and thus display smaller curved space distortions. ## 3 Magnetospheric Scattering Surface photons (e.g. from a thermal polar cap) must propagate through the magnetosphere surrounding the pulsar. Photons having an energy in the correct range pass through a local cyclotron resonance in the magnetosphere. For each photon energy, this defines a re-emission shell in the closed zone (Figure 1) in which strong scattering off the $`e^\pm `$ re-directs flux along the local magnetic field line direction. For millisecond pulsars with $`B10^810^9`$ gauss, closed zone resonance typically occurs for photons of energy $`110`$eV. With typical surface temperatures of a heated polar cap at $`10^6`$K, this radiation falls in the Rayleigh-Jeans part of the blackbody spectrum. In our Monte Carlo code, each photon trajectory represents a range of emitted photon energies. For each photon energy $`E_\gamma `$, we find the local cyclotron resonance position along the curved space path as the photon redshifts to lower energy. For rapidly rotating stars, we must include Doppler shifts by boosting the redshifted photon energy into the local plasma rest frame, checking for resonance, re-emitting in this locally co-rotating frame, and then boosting back to the lab frame where we continue to integrate the photon trajectory. We do not scatter at open zone resonance positions, since these are assumed to have no stationary plasma. In principle, photons can undergo multiple scatterings in the magnetosphere, for sufficiently large optical depth $`\tau `$. Following Rajagopal & Romani (1997), we estimate $`\tau `$ assuming a stationary, flat space dipolar magnetic field. The cyclotron energy as a function of distance from the star is $$E_\mathrm{c}=11.6B_{12}^{}(1+3\mathrm{cos}^2\theta _\mathrm{B})^{1/2}(R_{}/r)^3\mathrm{keV}$$ (9) where $`B_{12}^{}`$ is the equatorial surface dipole field strength in units of $`10^{12}`$ gauss, $`\theta _\mathrm{B}`$ is the polar angle between the dipole moment and the direction of the local magnetic field, and $`R_{}`$ is the neutron star radius. The resonant cyclotron cross section is $$\sigma _{\mathrm{res}}=\frac{\alpha _\mathrm{F}h^2}{m_\mathrm{e}}|e_{}|^2\delta (E_\gamma E_\mathrm{c}),$$ (10) where $`e_{}`$ contains the dependence on the photon direction and polarization and $`\alpha _\mathrm{F}`$ is the fine structure constant (e.g., Mèszàros, 1992). We assume a co-rotation (Goldreich-Julian) charge density $`n_{\mathrm{GJ}}=7\times 10^{13}B_{12}^z/P_{}(\mathrm{ms})\mathrm{cm}^3`$ (Goldreich & Julian, 1969). For rapid magnetospheric pair production the total $`e^\pm `$ density may, of course, be higher. Integrating the resonant cross section through the co-rotation charge density we get the optical depth for scattering $`\tau _{\mathrm{GJ}}(E_\gamma )`$ $`=`$ $`{\displaystyle _R_{}^{\mathrm{}}}n(r)\sigma _{\mathrm{res}}dr`$ (11) $`=`$ $`{\displaystyle \frac{2.3}{P_{}(\mathrm{ms})}}B_{12}^{1/3}\left({\displaystyle \frac{3\mathrm{e}\mathrm{V}}{E_\gamma }}\right)^{1/3}{\displaystyle \frac{B_z}{B}}{\displaystyle \frac{R_{}}{10\mathrm{k}\mathrm{m}}}`$ where we have taken $`|e_{}|^2=1/3`$. For concreteness, we adopt fiducial parameters for a millisecond pulsar ($`P_{}=3`$ms, $`R_{}=10`$km, and $`B_{}=10^9`$ gauss); these give a characteristic optical depth of $`\tau =7.7\times 10^2B_z/B`$ for $`E_\gamma =3`$eV. For polar cap emission, the average optical depth is typically somewhat smaller as significant flux travels through the scattering-free open zone. Thus, for co-rotation charge densities, we neglect multiple scatterings. In Figure 2, we plot light curves for thermal emission from a single polar cap, including closed zone scattering. The asymmetry in the scattered flux for the simple Schwarzschild propagations is due to the field ‘sweep-back’. Doppler boosts, aberration, and time delay effects are quite important, producing asymmetry in the direct flux and further phase shifting the scattered components. The ‘inter-pulse’ peaks in the scattered flux are due to radiation scattered back past the star and gravitationally focused to produce peaked pulse components. The strong energy dependence of the location of the scattering screen can been seen in the shifts of this pulse component. Note that the scattering amplitude decreases with $`E_\gamma `$ as is apparent from equation (11). The energy dependence of the phase shifts is larger for rapid rotators (small $`P_{}`$), as the high altitude scattering acquires additional aberration and Doppler boosting. Formally, these scattered radiation components provide a precision probe of the near star gravitational field, with field geometry and other systematics calibrated through the energy dependence of the component phase. In principle, precision tests of strong gravity are possible. We have run the code with Kerr metric propagations to check the detectability of these effects. At $`P_{}=1.5`$ms we find frame dragging shifts the positions of the scattered photon peaks (Figure 2) by $`\mathrm{\Delta }\varphi 0.01`$. We conclude that unless neutron stars are very compact at short periods, or unless closed zone plasmas substantially exceed the co-rotation density, frame dragging effects will be quite difficult to measure in scattered photons from surface thermal emission. ## 4 Outer Gap Emission As a straightforward extension of these models, we consider the effects of spacetime curvature on the light curves from outer gap emission, updating the work of Romani & Yadigaroglu (1995) to include curved space propagations. The outer gap region follows the edge of the closed zone extending from the ‘null charge surface’ (where $`n_{\mathrm{GJ}}`$ changes sign, i.e. $`B_z=0`$) to $`r_{\mathrm{LC}}`$. In outer gap models this region remains charge-starved, large potential drops develop, particles accelerate, and non-thermal radiation is produced tangent to the local $`B`$ (e.g., Cheng, Ho, & Ruderman, 1986; Romani, 1996). For MSP, this emission zone comes very close to the neutron star surface where spacetime curvature becomes important. To find the null charge boundary, we make a bicubic spline of the closed zone surface and step out along the grid until $`B_\mathrm{z}=0`$. We emit tangentially from the ‘last closed’ field line surface with an emissivity $`Fr^{3/2}`$. As described in Romani & Yadigaroglu (1995) aberration, boosts, and time delays are all essential for any outer gap light curve computation. Accordingly, here we check the additional distortions induced by curved space effects, as might be important for millisecond pulsars. In Figure 3, we show sample light curves for outer gap emission. The phase of the magnetic axis may be inferred from the structure of the radio polarization data and so the relative phasing of the high energy (outer gap) pulse is significant. Of course even classical pulsar radio emission, if arising from more than a few $`R_{}`$, will show similar phase shifts, so comparison of light curve phases must account for the altitude of the radio emission zone. The effect of neglecting Schwarzschild propagations is to both compress the pulse and shift it to later phase. For our fiducial parameters the first pulse component arises at high altitude. The second pulse arises from small $`r`$ where the cap field lines (and photon directions) are nearly radial, allowing only small gravitational bending. The curved space differences are thus dominated by the phase delays induced by the retardation in the strong near surface field. The curved space pulse expansion (relative to flat space) can similarly be attributed to gravitational time delay, since the first peak arises at larger $`r`$ where these effects are small. ## 5 Conclusions and Observational Prospects We have improved the modeling of light curves of rapidly rotating neutron stars (MSP) by extending the description of the surrounding vacuum dipole magnetosphere to include curved space and rapid rotation effects up to corrections of order $`𝒪[(\mathrm{\Omega }_{}r/c)^2(GM/rc^2)]`$. These effects produce substantial changes in the pulse components introduced by magnetospheric resonant scattering (for thermal surface emission) or direct emission from the rotating magnetosphere (high altitude gap radiation). We have developed a Monte Carlo code to compute light curves illustrating these effects including the energy dependence of the scattered photon pulse shapes. Higher order effects from frame dragging have also been computed, but these are likely too subtle to be discernible in magnetospheric pulse components unless the S/N is very high. Scattering perturbations, though small, will be greatly enhanced if the magnetosphere can support a plasma with $`\tau >\tau _{\mathrm{GJ}}`$. In this paper, we have only considered the minimum possible scattering perturbation. With new detector technologies (e.g. Romani et al., 1999), we should see a marked increase in the quality of optical light curves. With enhanced sensitivity and spectral resolution (and possibly higher scattering $`\tau `$), light curve perturbations due to rapid rotation effects may be promoted to important probes of pulsar physics. This work was supported in part by NASA grant NAG5-3263. Roger W. Romani is a Cottrell Scholar of Research Corporation.
warning/0004/hep-ph0004056.html
ar5iv
text
# 1 Introduction ## 1 Introduction Future high energy experiments at the LHC and possibly a linear collider (TESLA or the NLC for instance), will probe the full non-Abelian nature of the electroweak Standard Model (SM). Thus one has to view the photon in particular as a particle with non-Abelian character. At energies much larger than the weak scale, Sudakov logarithms, originating from vector boson exchange, can lead to significant radiative corrections. The double logarithms (DL) can be of order $`𝒪(20`$ %$`)`$ at one loop in the TeV range and a few % at the two loop level. In addition, subleading corrections can also be significant, especially if the experimental accuracy is of the order of $`𝒪(1`$ %$`)`$. As of this writing, there is no complete two loop calculation in the electroweak theory due to the complexity of the number and nature of processes involved. It is thus of considerable interest to investigate terms which are potentially large and which can be resummed to all orders. In Ref. the leading DL corrections were calculated and found to exponentiate. The results were obtained by using the infrared evolution equation method calculated with the massless fields of the unbroken theory. The equation has a different kernel depending on the value of the infrared cutoff. There are, however, some important differences of the electroweak theory with respect to an unbroken gauge theory. Since the physical cutoff of the massive gauge bosons is the weak scale $`MM_\mathrm{w}`$, pure virtual correction lead to physical cross sections depending on the infrared “cutoff”. Only the photon needs to be treated in a semi-inclusive way. Additional complications arise due to the mixing involved to make the mass eigenstates and the fact that at high energies, the longitudinal degrees of freedom are not suppressed. Furthermore, since the asymptotic states are not group singlets, it is expected that fully inclusive cross sections contain Bloch-Nordsieck violating electroweak corrections . In this paper we extend the method of Ref. to the next to leading order for the case of virtual corrections at high energies where we can neglect particle masses. In addition, we show that the results of Ref. are also valid for longitudinal degrees of freedom, which at first sight, is far from obvious. The connection between the calculation performed in the massless theory and the longitudinal degrees of freedom is provided by the Goldstone boson equivalence theorem. Another complication in comparison with the unbroken non-Abelian case is the mixing of the mass-eigenstates. Especially for the external $`Z`$-boson and the photon states, the corresponding corrections in general don’t factorize with respect to the original amplitude. We indicate how the corrections are given to subleading level in terms of the fields of the unbroken theory. Finally we compare our results with existing one-loop corrections in the high energy approximation for on-shell $`W`$-pair production in electron-positron scattering. Although this comparison constitutes a strong test of our approach it should be mentioned that it would be extremely helpful to compare the results obtained in this approach with a general method in terms of the physical fields. In this context a two loop DL-calculation would help clarify the situation with contradicting results in the literature and a subleading one loop approach would give further support to the results presented in this work. ## 2 Logarithmic corrections in non-Abelian theories In this section we are concerned with virtual double and single logarithmic corrections to scattering amplitudes in massless non-Abelian theories at fixed angle with all invariants large with respect to an infrared cutoff $`\mu `$, i.e. $`\mu ^2s_{j,l}2p_jp_ls`$. It must be emphasized that in high energy collider experiments there are also contributions depending on angular variables (i.e. u/t etc.) which can be of genuine subleading nature . The philosophy adopted here is that terms of the type $`\mathrm{log}\frac{s}{\mu ^2}\mathrm{log}\frac{u}{t}`$ etc. should be calculated exactly at least at the one loop level. For the higher order terms below, we are only concerned with the $`\mathrm{log}\frac{s}{\mu ^2}`$ behavior with $`\mu ^2s|t||u|`$. All mass terms are neglected, i.e. we assume $`m_i<\mu `$. We begin by reviewing the general method for virtual corrections in the DL-approximation following the approach of Ref. . ### 2.1 Double logarithmic corrections Sudakov effects have been widely discussed for non-Abelian gauge theories, such as $`SU(N)`$ and can be calculated in various ways (see, for instance, ). A general method of finding the DL asymptotics (not only of the Sudakov type) is based on the infrared evolution equations describing the dependence of the amplitudes on the infrared cutoff $`\mu `$ of the virtual particle transverse momenta . This cutoff plays the same role as the fictitious photon mass $`\lambda `$ in QED, but, unlike $`\lambda `$, it is not necessary that it vanishes and it may take an arbitrary value. It can be introduced in a gauge invariant way by working, for instance, in a finite phase space volume in the transverse direction with linear size $`l1/\mu `$. Instead of calculating asymptotics of particular Feynman diagrams and summing these asymptotics for a process with $`n`$ external lines it is convenient to extract the virtual particle with the smallest value of $`|𝒌_{}|`$ in such a way, that the transverse momenta $`|𝒌_{}^{}|`$ of the other virtual particles are much bigger $$𝒌_{}^^2𝒌_{}^2\mu ^2.$$ (1) For the other particles $`𝒌_{}^2`$ plays the role of the initial infrared cut-off $`\mu ^2`$. In particular, the Sudakov DL corrections are related to the exchange of soft gauge bosons. For this case the integral over the momentum $`k`$ of the soft (i.e. $`|k^0|\sqrt{s}`$) virtual boson with the smallest $`𝒌_{}`$ can be factored off, which leads to the following infrared evolution equation: $`(p_1,\mathrm{},p_n;\mu ^2)`$ $`=`$ $`_{\mathrm{Born}}(p_1,\mathrm{},p_n){\displaystyle \frac{i}{2}}{\displaystyle \frac{g_s^2}{(2\pi )^4}}{\displaystyle \underset{j,l=1,jl}{\overset{n}{}}}{\displaystyle _{s𝒌_{}^2\mu ^2}}{\displaystyle \frac{d^4k}{k^2+iϵ}}{\displaystyle \frac{p_jp_l}{(kp_j)(kp_l)}}`$ (2) $`\times T^a(j)T^a(l)(p_1,\mathrm{},p_n;𝒌_{}^2),`$ where the amplitude $`(p_1,\mathrm{},p_n;𝒌_{}^2)`$ on the right hand side is to be taken on the mass shell, but with the substituted infrared cutoff: $`\mu ^2𝒌_{}^2`$. The generator $`T^a(l)(a=1,\mathrm{},N)`$ acts on the color indices of the particle with momentum $`p_l`$. The non-Abelian gauge coupling is $`g`$. In Eq. (2), and below, $`𝒌_{}`$ denotes the component of the gauge boson momentum $`𝒌`$ transverse to the particle emitting this boson. It can be expressed in invariant form as $`𝒌_{}^2\mathrm{min}((kp_l)(kp_j)/(p_lp_j))`$ for all $`jl`$. The above factorization is related to a non-Abelian generalization of the Gribov theorem<sup>1</sup><sup>1</sup>1The non-Abelian generalization of Gribov’s theorem is given in Ref. , together with a description of its essential content. for the amplitude of the Bremsstrahlung of a photon with small transverse momentum $`𝒌_{}`$ in high energy hadron scattering . The form in which we present Eq. (2) corresponds to a covariant gauge for the gluon with momentum $`k`$. Formally this expression can be written in a gauge invariant way if we include in the sum the term with $`j=l`$ (which does not give a DL contribution). Indeed, in this case we can substitute $`p_ip_j`$ by $`p_i^\mu p_j^\nu d_{\mu \nu }(k)`$, where the polarization matrices of the boson $`d_{\mu \nu }(k)`$ in the various gauges differ by the terms proportional to $`k^\mu `$ or $`k^\nu `$ giving a vanishing contribution due to the conservation of the total color charge $`_aT^a=0`$. Thus we have the possibility of choosing appropriate gauges for each kinematical region of quasi-collinearity of $`k`$ and $`p_l`$. We can, however, use (2) as well, noting that in this region for $`jl`$ we have $`p_jp_l/kp_jE_l/\omega `$, where $`E_l`$ is the energy of the particle with momentum $`p_l`$ and $`\omega `$ the frequency of the emitted gauge boson, so that: $`(p_1,\mathrm{},p_n;\mu ^2)`$ $`=`$ $`_{\mathrm{Born}}(p_1,\mathrm{},p_n){\displaystyle \frac{2g_s^2}{(4\pi )^2}}{\displaystyle \underset{l=1}{\overset{n}{}}}{\displaystyle _{\mu ^2}^s}{\displaystyle \frac{d𝒌_{}^2}{𝒌_{}^2}}{\displaystyle _{|𝒌_{}|/\sqrt{s}}^1}{\displaystyle \frac{dv}{v}}`$ (3) $`\times C_l(p_1,\mathrm{},p_n;𝒌_{}^2),`$ where $`C_l`$ is the eigenvalue of the Casimir operator $`T^a(l)T^a(l)`$ ($`C_l=C_A`$ for gauge bosons in the adjoint representation of the gauge group $`SU(N)`$ and $`C_l=C_F`$ for fermions in the fundamental representation). In this last step we also used the identity $`_{j=1}^nT^a(j)(p_1,\mathrm{},p_n;𝒌_{}^2)=0`$, corresponding to the conservation of the total group charge. The integral over $`d^4k`$ was written in terms of the Sudakov components according to the discussion in section 2.2 upon replacing the longitudinal component $`u`$ with the boson on-shell expression $`suv=\mu ^2+𝒌_{}^2`$. Thus, in Sudakov DL corrections there are no interference effects, so that we can talk about the emission (and absorption) of a gauge boson by a definite (external) particle, namely by a particle with momentum almost collinear to $`𝒌`$. The differential form of the infrared evolution equation follows immediately from (3): $$\frac{(p_1,\mathrm{},p_n;\mu ^2)}{\mathrm{log}(\mu ^2)}=K(\mu ^2)(p_1,\mathrm{},p_n;\mu ^2),$$ (4) where $$K(\mu ^2)\frac{1}{2}\underset{l=1}{\overset{n}{}}\frac{W_l(s,\mu ^2)}{\mathrm{log}(\mu ^2)}$$ (5) with $$W_l(s,\mu ^2)=\frac{g_s^2}{(4\pi )^2}C_l\mathrm{log}^2\frac{s}{\mu ^2}.$$ (6) $`W_l`$ is the probability to emit a soft and almost collinear gauge boson from the particle $`l`$, subject to the infrared cut-off $`\mu `$ on the transverse momentum . Note again that the cut-off $`\mu `$ is not taken to zero. To logarithmic accuracy, we obtain directly from (6): $$\frac{W_l(s,\mu ^2)}{\mathrm{log}(\mu ^2)}=\frac{g_s^2}{8\pi ^2}C_l\mathrm{log}\frac{s}{\mu ^2}.$$ (7) The infrared evolution equation (4) should be solved with an appropriate initial condition. In the case of large scattering angles, if we choose the cut-off to be the large scale $`s`$ then clearly there are no Sudakov corrections. The initial condition is therefore $$(p_1,\mathrm{},p_n;s)=_{\mathrm{Born}}(p_1,\mathrm{},p_n),$$ (8) and the solution of (4) is thus given by the product of the Born amplitude and the Sudakov form factors: $$(p_1,\mathrm{},p_n;\mu ^2)=_{\mathrm{Born}}(p_1,\mathrm{},p_n)\mathrm{exp}\left(\frac{1}{2}\underset{l=1}{\overset{n}{}}W_l(s,\mu ^2)\right)$$ (9) Therefore we obtain an exactly analogous Sudakov exponentiation for the gauge group $`SU(N)`$ to that for the Abelian case . ### 2.2 Soft divergences in the massless theory In this section we briefly review the types of soft, i.e. $`|k^0|\sqrt{s}`$, divergences in loop corrections with massless particles. In general, those contributions, unlike the collinear logarithms, can be obtained by setting all $`k`$ dependent terms in the numerator of tensor integrals to zero (since the terms left are of the order of the hard scale $`s`$). Thus it is clear that the tensor structure which emerges is that of the inner scattering amplitude in Fig. 2 taken on the mass-shell, times a scalar function of the given loop correction. In the Feynman gauge, for instance, we find for the well known vertex corrections the familiar three-point function $`C_0`$ and for higher point functions we note that in the considered case all infrared divergent scalar integrals reduce to $`C_0`$ multiplied by factors of $`\frac{1}{s}`$ etc.. The only infrared divergent three point function is given by $$C_0(s/\mu ^2)_{𝒌_{}^2>\mu ^2}\frac{d^4k}{(2\pi )^4}\frac{1}{(k^2+i\epsilon )(k^2+2p_jk+i\epsilon )(k^22p_lk+i\epsilon )}$$ (10) It is now convenient to use the Sudakov parametrization for the exchanged virtual boson: $$k=vp_j+up_l+k_{}$$ (11) For the boson propagator we use the identity $$\frac{i}{suv𝒌_{}^2+i\epsilon }=𝒫\frac{i}{suv𝒌_{}^2}+\pi \delta (suv𝒌_{}^2)$$ (12) writing it in form of the real and imaginary parts (the principle value is indicated by $`𝒫`$). The latter does not contribute to the DL asymptotics and at higher orders gives subsubleading contributions. Rewriting the measure as $`d^4k=d^2k_{}d^2k_{}`$ with $`d^2k_{}`$ $`=`$ $`|𝒌_{}|d|𝒌_{}|d\varphi ={\displaystyle \frac{1}{2}}d𝒌_{}^2d\varphi =\pi d𝒌_{}^2`$ (13) $`d^2k_{}`$ $`=`$ $`|(k^0,k^x)/(u,v)|dudv{\displaystyle \frac{s}{2}}dudv`$ (14) where we turn the coordinate system such that the $`p_j,p_l`$ plane corresponds to $`0,x`$ and the $`y,z`$ coordinates to the $`k_{}`$ direction so that it is purely spacelike. The last equation follows from $`p_l^2=0`$, i.e. $`p_{l_x}^2p_{l_0}^2`$ and $$(p_{j_0}p_{l_x}p_{l_0}p_{j_x})^2(p_{j_0}p_{l_0}p_{l_x}p_{j_x})^2=(p_jp_l)^2=s/2$$ (15) The function $`C_0(s/\mu ^2)`$ is fastly converging for large $`𝒌_{}^2`$ and we are interested here in the region $`\mu ^2s`$ in order to obtain large logarithms. Then logarithmic corrections come from the region $`𝒌_{}^2s|u|,s|v|s`$ (the strong inequalities give DL, the simple inequalities single ones) and we can write to logarithmic accuracy: $`C_0(s/\mu ^2)`$ $`=`$ $`{\displaystyle \frac{s\pi }{2(2\pi )^4}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}du{\displaystyle _{\mathrm{}}^{\mathrm{}}}dv{\displaystyle _{\mu ^2}^{\mathrm{}}}d𝒌_{}^2\times `$ (16) $`{\displaystyle \frac{1}{(suv𝒌_{}^2+i\epsilon )(suv𝒌_{}^2+su+i\epsilon )(suv𝒌_{}^2sv+i\epsilon )}}`$ $``$ $`{\displaystyle \frac{si\pi ^2}{2(2\pi )^4}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{du}{su}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}{\displaystyle \frac{dv}{sv}}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑𝒌_{}^2\mathrm{\Theta }(𝒌_{}^2\mu ^2)\delta (suv𝒌_{}^2)`$ $``$ $`{\displaystyle \frac{i}{2(4\pi )^2s}}{\displaystyle _1^1}{\displaystyle \frac{du}{u}}{\displaystyle _1^1}{\displaystyle \frac{dv}{v}}\mathrm{\Theta }(suv\mu ^2)`$ $`=`$ $`{\displaystyle \frac{i}{(4\pi )^2s}}{\displaystyle _0^1}{\displaystyle \frac{du}{u}}{\displaystyle _0^1}{\displaystyle \frac{dv}{v}}\mathrm{\Theta }(suv\mu ^2)`$ $`=`$ $`{\displaystyle \frac{i}{(4\pi )^2s}}{\displaystyle _{\frac{\mu ^2}{s}}^1}{\displaystyle \frac{du}{u}}{\displaystyle _{\frac{\mu ^2}{su}}^1}{\displaystyle \frac{dv}{v}}`$ $`=`$ $`{\displaystyle \frac{i}{2(4\pi )^2s}}\mathrm{log}^2{\displaystyle \frac{s}{\mu ^2}}`$ Thus, no single soft logarithmic corrections are present in $`C_0(s/\mu ^2)`$. In order to see that this result is not just a consequence of our regulator, we repeat the calculation for a fictitious gluon mass<sup>2</sup><sup>2</sup>2Note that this regulator spoils gauge invariance and leads to possible inconsistencies at higher orders. Great care must be taken for instance when a three gluon vertex is regulated inside a loop integral.. In this case we have $$C_0(s/\lambda ^2)\frac{d^4k}{(2\pi )^4}\frac{1}{(k^2\lambda ^2+i\epsilon )(k^2+2p_jk+i\epsilon )(k^22p_lk+i\epsilon )}$$ (17) It is clear that $`C_0(s/\lambda ^2)`$ contains soft and collinear divergences ($`kp_{j,l}`$) and is regulated with the cutoff $`\lambda `$, which plays the role of $`\mu `$ in this case. Integrating over Feynman parameters we find: $$C_0(s/\lambda ^2)=\frac{i}{(4\pi )^2s}\left(\frac{1}{2}\mathrm{log}^2\frac{s}{\lambda ^2+i\epsilon }+\frac{\pi ^2}{3}\right)$$ (18) We are only interested here in the real part of loop corrections of scattering amplitudes since they are multiplied by the Born amplitude and the imaginary pieces contribute to cross sections at the next to next to leading level as mentioned above. In fact, the minus sign inside the double logarithm corresponds precisely to the omitted principle value contribution of Eq. (12) in the previous calculation. Thus, no single soft logarithmic correction is present in the case when particle masses can be neglected. This feature prevails to higher orders as well since it has been shown that also in non-Abelian gauge theories the one-loop Sudakov form factor exponentiates . In case we would keep mass-terms, even two point functions, which in our scheme can only yield collinear logarithms, would contain a soft logarithm due to the mass-renormalization which introduces a derivative contribution . In conclusion, all leading soft corrections are contained in double logarithms (soft and collinear) and subleading logarithmic corrections in a massless theory, with all invariants large ($`s_{j,l}=2p_jp_l𝒪(s)`$) compared to the infrared cutoff, are of the collinear type or renormalization group logarithms. ### 2.3 Virtual logarithmic corrections from the Altarelli-Parisi splitting functions In an axial gauge, collinear logarithms are related to corrections on a particular external leg depending on the choice of the four vector $`n_\nu `$ . A typical diagram is depicted in Fig. 1. In a general covariant gauge this corresponds (using Ward identities) to a sum over insertions in all $`n`$ external legs . We can therefore adopt the strategy to extract the gauge invariant contribution from the external line corrections on the invariant matrix element at the subleading level. The results of the previous section are thus important in that they allow the use of the Altarelli-Parisi approach to calculate the subleading contribution to the evolution kernel of Eq. (4). We are here only concerned with virtual corrections and use the universality of the splitting functions to calculate the subleading terms. For this purpose we use the virtual quark and gluon contributions to the splitting functions $`P_{qq}^V(z)`$ and $`P_{gg}^V(z)`$ describing the probability to emit a soft and/or collinear virtual particle with energy fraction $`z`$ of the original external line four momentum. The infinite momentum frame corresponds to the Sudakov parametrization with lightlike vectors. In general, the splitting functions $`P_{BA}`$ describe the probability of finding a particle $`B`$ inside a particle $`A`$ with fraction $`z`$ of the longitudinal momentum of $`A`$ with probability $`𝒫_{BA}`$ to first order : $$d𝒫_{BA}(z)=\frac{\alpha _s}{2\pi }P_{BA}dt$$ (19) where the variable $`t=\mathrm{log}\frac{s}{\mu ^2}`$ for our purposes. It then follows that $$d𝒫_{BA}(z)=\frac{\alpha _s}{2\pi }\frac{z(1z)}{2}\overline{\underset{spins}{}}\frac{|V_{AB+C}|^2}{𝒌_{}^2}d\mathrm{log}𝒌_{}^2$$ (20) where $`V_{AB+C}`$ denotes the elementary vertices and $$P_{BA}(z)=\frac{z(1z)}{2}\overline{\underset{spins}{}}\frac{|V_{AB+C}|^2}{𝒌_{}^2}$$ (21) The upper bound on the integral over $`d𝒌_{}^2`$ in Eq. (20) is $`s`$ and it is thus directly related to $`dt`$. Regulating the virtual infrared divergences with the transverse momentum cutoff as described above, we find the virtual contributions to the splitting functions for external quark and gluon lines: $`P_{qq}^V(z)`$ $`=`$ $`C_F\left(2\mathrm{log}{\displaystyle \frac{s}{\mu ^2}}+3\right)\delta (1z)`$ (22) $`P_{gg}^V(z)`$ $`=`$ $`C_A\left(2\mathrm{log}{\displaystyle \frac{s}{\mu ^2}}+{\displaystyle \frac{4}{C_A}}\beta _0^{\mathrm{QCD}}\right)\delta (1z)`$ (23) The functions can be calculated directly from loop corrections to the elementary processes and the logarithmic term corresponds to the leading kernel of section 2.1. We introduce virtual distribution functions which include only the effects of loop computations. These fulfill the Altarelli-Parisi equations<sup>3</sup><sup>3</sup>3Note that the off diagonal splitting functions $`P_{qg}`$ and $`P_{gq}`$ do not contribute to the virtual probabilities to the order we are working here. In fact, for virtual corrections there is no need to introduce off-diagonal terms as the corrections factorize with respect to the Born amplitude. The normalization of the Eqs. (22) and (23) corresponds to calculations in two to two processes on the cross section level with the gluon symmetry factor $`\frac{1}{2}`$ included. The results, properly normalized, are process independent. $`{\displaystyle \frac{q(z,t)}{t}}`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{2\pi }}{\displaystyle _z^1}{\displaystyle \frac{dy}{y}}q(z/y,t)P_{qq}^V(y)`$ (24) $`{\displaystyle \frac{g(z,t)}{t}}`$ $`=`$ $`{\displaystyle \frac{\alpha _s}{2\pi }}{\displaystyle _z^1}{\displaystyle \frac{dy}{y}}g(z/y,t)P_{gg}^V(y)`$ (25) The splitting functions are related by $`P_{BA}=P_{BA}^R+P_{BA}^V`$, where $`R`$ denotes the contribution from real gauge boson emission<sup>4</sup><sup>4</sup>4$`P_{qq}`$ was first calculated by V.N. Gribov and L.N. Lipatov in the context of QED .. $`P_{BA}`$ is free of logarithmic corrections and positive definite. The subleading term in Eq. (23) indicates that the only subleading corrections in the pure glue sector are related to a shift in the scale of the coupling. These corrections enter with a different sign compared to the conventional running coupling effects. The renormalizations with respect to the Born amplitude as well as the ones belonging to the next to leading terms at higher orders will be indicated below by writing $`\alpha _s(\overline{\mu }^2)\alpha _s(s)`$. For fermion lines there is an additional subleading correction from collinear terms which is not related to a change in the scale of the coupling. Inserting the virtual probabilities of Eqs. (22) and (23) into the Eqs. (24) and (25) we find: $`q(1,t)`$ $`=`$ $`q_0\mathrm{exp}\left[{\displaystyle \frac{\alpha _s(s)C_F}{2\pi }}\left(\mathrm{log}^2{\displaystyle \frac{s}{\mu ^2}}3\mathrm{log}{\displaystyle \frac{s}{\mu ^2}}\right)\right]`$ (26) $`g(1,t)`$ $`=`$ $`g_0\mathrm{exp}\left[{\displaystyle \frac{\alpha _s(s)C_A}{2\pi }}\left(\mathrm{log}^2{\displaystyle \frac{s}{\mu ^2}}{\displaystyle \frac{4}{C_A}}\beta _0^{\mathrm{QCD}}\mathrm{log}{\displaystyle \frac{s}{\mu ^2}}\right)\right]`$ (27) where $`\beta _0^{\mathrm{QCD}}=\frac{11}{12}C_A\frac{1}{3}T_Fn_f`$ with $`C_A=3`$ and $`T_F=\frac{1}{2}`$. These functions describe the total contribution for the emission of virtual particles (i.e. $`z=1`$), with all invariants large compared to the cutoff $`\mu `$, to the densities $`q(z,t)`$ and $`g(z,t)`$. The normalization is on the level of the cross section. For the invariant matrix element we thus find at the subleading level: $`(p_1,\mathrm{},p_n,g_s,\mu )`$ $`=`$ $`(p_1,\mathrm{},p_n,g_s(s))\times `$ (28) $`\mathrm{exp}\left({\displaystyle \frac{1}{2}}{\displaystyle \underset{j=1}{\overset{n_q}{}}}W_j^q(s,\mu ^2){\displaystyle \frac{1}{2}}{\displaystyle \underset{l=1}{\overset{n_g}{}}}W_l^g(s,\mu ^2)\right)`$ with $`n_q+n_g=n`$, and $`W^q(s,\mu ^2)`$ $`=`$ $`{\displaystyle \frac{\alpha _s(s)C_F}{4\pi }}\left(\mathrm{log}^2{\displaystyle \frac{s}{\mu ^2}}3\mathrm{log}{\displaystyle \frac{s}{\mu ^2}}\right)`$ (29) $`W^g(s,\mu ^2)`$ $`=`$ $`{\displaystyle \frac{\alpha _s(s)C_A}{4\pi }}\left(\mathrm{log}^2{\displaystyle \frac{s}{\mu ^2}}{\displaystyle \frac{4}{C_A}}\beta _0^{\mathrm{QCD}}\mathrm{log}{\displaystyle \frac{s}{\mu ^2}}\right)`$ (30) Again we note that the running coupling notation in the Born-amplitude of Eq. (28) denotes the renormalization corrections of the Born amplitude and higher order corrections. The functions $`W^q`$, $`W^g`$ correspond to the probability of emitting a virtual soft and/or collinear gauge boson from the particle $`q`$, $`g`$ subject to the infrared cutoff $`\mu `$. Typical diagrams contributing to Eq. (28) in a covariant gauge are depicted in Fig. 2. In massless QCD there is no need for the label $`W_j^q`$ or $`W_l^g`$, however, we write it for later convenience. The universality of the splitting functions is crucial in obtaining the above result. ### 2.4 Renormalization group equation The solution presented in Eq. (28) determines the evolution of the virtual scattering amplitude $`(p_1,\mathrm{},p_n,g_s,\mu )`$ for large energies at fixed angles and subject to the infrared regulator $`\mu `$. In the massless case there is a one to one correspondence between the high energy limit and the infrared limit as only the ratio $`s/\mu ^2`$ enters as a dimensionless variable . Thus, we can generalize the Altarelli-Parisi equations (24) and (25) to the invariant matrix element in the language of the renormalization group. For this purpose, we define the infrared singular (logarithmic) anomalous dimensions $$\mathrm{\Gamma }_q(t)\frac{C_F\alpha _s}{4\pi }t;\mathrm{\Gamma }_g(t)\frac{C_A\alpha _s}{4\pi }t$$ (31) Infrared divergent anomalous dimensions have been derived in the context of renormalization properties of gauge invariant Wilson loop functionals . In this context they are related to undifferentiable cusps of the path integration and the cusp angle $`p_jp_l/\mu ^2`$ gives rise to the logarithmic nature of the anomalous dimension. In case we use off-shell amplitudes, one also has contributions from end points of the integration . The leading terms in the equation below have also been discussed in Refs. , and in the context of QCD. With these notations we find that Eq. (28) satisfies $`\left({\displaystyle \frac{}{t}}+\beta ^{\mathrm{QCD}}{\displaystyle \frac{}{g_s}}+n_g\left(\mathrm{\Gamma }_g(t){\displaystyle \frac{1}{2}}{\displaystyle \frac{\alpha _s}{\pi }}\beta _0^{\mathrm{QCD}}\right)+n_q\left(\mathrm{\Gamma }_q(t)+{\displaystyle \frac{1}{2}}\gamma _{q\overline{q}}\right)\right)`$ $`\times (p_1,\mathrm{},p_n,g_s,\mu )=0`$ (32) to the order we are working here and where $`(p_1,\mathrm{},p_n,g_s,\mu )`$ is taken on the mass-shell. The difference in the sign of the derivative term compared to Eq. (4) is due to the fact that instead of differentiating with respect to $`\mathrm{log}\mu ^2`$ we use $`\mathrm{log}s/\mu ^2`$. The quark-antiquark operator anomalous dimension $`\gamma _{q\overline{q}}=C_F\frac{3}{4}\frac{\alpha }{\pi }`$ enters even for massless theories as the quark antiquark operator leads to scaling violations through loop effects since the quark masslessness is not protected by gauge invariance and a dimensionful infrared cutoff needs to be introduced. Thus, although the Lagrangian contains no $`mq\overline{q}`$ term, quantum corrections lead to the anomalous scaling violations in the form of $`\gamma _{q\overline{q}}`$. The factor $`\frac{1}{2}`$ occurs since we write Eq. (32) in terms of each external line separately<sup>5</sup><sup>5</sup>5In case of a massive theory, we could, for instance avoid the anomalous dimension term $`\gamma _{q\overline{q}}`$ by adopting the pole mass definition. In this case, however, we would obtain terms in the wave function renormalization, and in any case, the one to one correspondence between UV and IR scaling, crucial for the validity of Eq. (32), is violated.. For the gluon, the scaling violation due to the infrared cutoff are manifest in terms of an anomalous dimension proportional to the $`\beta `$-function since the gluon mass is protected by gauge invariance from loop corrections. Thus, in the bosonic sector the subleading terms correspond effectively to a scale change of the coupling. Fig. 3 illustrates the corrections to the external quark-antiquark lines from loop effects. Except for the infrared singular anomalous dimension (Eq. (31)), all other terms in Eq. (32) are the standard contributions to the renormalization group equation for S-matrix elements . In QCD, observables with infrared singular anomalous dimensions, regulated with a fictitious gluon mass, are ill defined due to the masslessness of gluons. In the electroweak theory, however, we can legitimately investigate only virtual corrections since the gauge bosons will require a mass. Eq. (32) will thus be very useful in the following sections. ## 3 Logarithmic corrections in broken gauge theories In the following we will apply the results obtained in the previous sections to the case of spontaneously broken gauge theories. It will be necessary, at least at the subleading level, to distinguish between transverse and longitudinal degrees of freedom. The physical motivation in this approach is that for very large energies, $`sM_W^2M^2`$, the electroweak theory is in the unbroken phase, with an exact $`SU(2)\times U(1)`$ gauge symmetry. We will calculate the corrections to this theory and use the high energy solution as a matching condition for the regime for values of $`\mu <M`$. We begin by considering some simple kinematic arguments for massive vector bosons. A vector boson at rest has momentum $`k^\nu =(M,0,0,0)`$ and a polarization vector that is a linear combination of the three orthogonal unit vectors $$e_1(0,1,0,0),e_1(0,0,1,0),e_3(0,0,0,1).$$ (33) After boosting this particle along the $`3`$-axis, its momentum will be $`k^\nu =(E_k,0,0,k)`$. The three possible polarization vectors are now still satisfying: $$k_\nu e_j^\nu =0,e_j^2=1.$$ (34) Two of these vectors correspond to $`e_1`$ and $`e_2`$ and describe the transverse polarizations. The third vector satisfying (34) is the longitudinal polarization vector $$e_L^\nu (k)=(k/M,0,0,E_k/M)$$ (35) i.e. $`e_L^\nu (k)=k^\nu /M+𝒪(M/E_k)`$ for large energies. These considerations illustrate that the transversely polarized degrees of freedom at high energies are related to the massless theory, while the longitudinal degrees of freedom need to be considered separately. Another manifestation of the different high energy nature of the two polarization states is contained in the Goldstone boson equivalence theorem. It states that the unphysical Goldstone boson that is “eaten up” by a massive gauge boson still controls its high energy asymptotics. A more precise formulation is given below in section 3.2. Thus we can legitimately use the results obtained in the massless non-Abelian theory if we restrict ourselves to the transverse degrees of freedom at high energies. We will, however, show that to DL accuracy the results of Ref. can be used in connection with the Goldstone boson equivalence theorem. Another difference to the situation in an unbroken non-Abelian theory is the mixing of the physical fields with the fields in the unbroken phase. These complications are especially relevant for the $`Z`$-boson and the photon. ### 3.1 Results for transverse degrees of freedom The results we obtain in this section are generally valid for spontaneously broken gauge theories, however, for definiteness we discuss only the electroweak Standard Model. The physical gauge bosons are thus a massless photon (described by the field $`A_\nu `$) and massive $`W^\pm `$ and $`Z`$ bosons (described correspondingly by fields $`W_\nu ^\pm `$ and $`Z_\nu `$).: $`W_\nu ^\pm `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{2}}}\left(W_\nu ^1\pm iW_\nu ^2\right)`$ (36) $`Z_\nu `$ $`=`$ $`\mathrm{cos}\theta _\mathrm{w}W_\nu ^3+\mathrm{sin}\theta _\mathrm{w}B_\nu `$ (37) $`A_\nu `$ $`=`$ $`\mathrm{sin}\theta _\mathrm{w}W_\nu ^3+\mathrm{cos}\theta _\mathrm{w}B_\nu `$ (38) Thus, amplitudes containing physical fields will correspond to a linear combination of the massless fields in the unbroken phase. The situation is illustrated schematically for a single gauge boson external leg in Fig. 4. In case of the $`W^\pm `$ bosons, the corrections factorize with respect to the physical amplitude. To logarithmic accuracy, all masses can be set equal: $$M_ZM_WM_{\mathrm{Higgs}}M$$ and the energy considered to be much larger, $`\sqrt{s}M`$. The physical fields are given in terms of the unbroken fields according to Eqs. (36), (37) and (38). The left and right handed fermions are correspondingly doublets ($`T=1/2`$) and singlets ($`T=0`$) of the $`SU`$(2) weak isospin group and have hypercharge $`Y`$ related to the electric charge $`Q`$, measured in units of the proton charge, by the Gell-Mann-Nishijima formula $`Q=T^3+Y/2`$. The value for the infrared cutoff $`\mu `$ can be chosen in two different regimes: 1) $`\sqrt{s}\mu M`$ and 2) $`\mu M`$. The second case is universal in the sense that it does not depend on the details of the electroweak theory and will be discussed below. In the first region we can neglect spontaneous symmetry breaking effects (in particular gauge boson masses) and consider the theory with fields $`B_\nu `$ and $`W_\nu ^a`$. One could of course also calculate everything in terms of the physical fields, however, we emphasize again that in this case we need to consider the photon also in region 1). The omission of the photon would lead to the violation of gauge invariance since the photon contains a mixture of the $`B_\nu `$ and $`W_\nu ^3`$ fields. In region 1), the renormalization group equation (or generalized infrared evolution equation) (32) in the case of all $`m_i<M`$ reads<sup>6</sup><sup>6</sup>6We exclude here top-Yukawa couplings which couple proportional to $`\frac{m_t^2}{M^2}`$ since they don’t have an analogue in QCD. It is, however, not unlikely that those terms can also be included in the splitting functions fulfilling Altarelli-Parisi equations. Note also, that the amplitude on the right hand side is in general a linear combination of fields in the unbroken phase according to Eqs. (36), (37) and (38). In addition, in the electroweak theory matching will be required at the scale $`M`$ and often on-shell renormalization of the couplings $`e`$ and $`\mathrm{sin}\theta _\mathrm{w}`$ is used. In this case one has additional complications in the running coupling terms due to the different mass scales involved below $`M`$. Details are presented in section 4. $`\left({\displaystyle \frac{}{t}}+\beta {\displaystyle \frac{}{g}}+\beta ^{}{\displaystyle \frac{}{g^{}}}+{\displaystyle \underset{i=1}{\overset{n_g}{}}}\mathrm{\Gamma }_g^i(t)n_\mathrm{W}{\displaystyle \frac{1}{2}}{\displaystyle \frac{\alpha }{\pi }}\beta _0n_\mathrm{B}{\displaystyle \frac{1}{2}}{\displaystyle \frac{\alpha ^{}}{\pi }}\beta _0^{}+{\displaystyle \underset{k=1}{\overset{n_f}{}}}\left(\mathrm{\Gamma }_f^k(t)+{\displaystyle \frac{1}{2}}\gamma _{q\overline{q}}^k\right)\right)`$ $`\times ^{}(p_1,\mathrm{},p_n,g,g^{},\mu )=0`$ (39) where the index $``$ indicates that we consider only $`n_g`$ transversely polarized external gauge bosons and $`n_\mathrm{W}+n_\mathrm{B}=n_g`$. The two $`\beta `$-functions are given by: $`\beta (g(\overline{\mu }^2))`$ $`=`$ $`{\displaystyle \frac{g(\overline{\mu }^2)}{\mathrm{log}\overline{\mu }^2}}\beta _0{\displaystyle \frac{g^3(\overline{\mu }^2)}{8\pi ^2}}`$ (40) $`\beta ^{}(g^{}(\overline{\mu }^2))`$ $`=`$ $`{\displaystyle \frac{g^{}(\overline{\mu }^2)}{\mathrm{log}\overline{\mu }^2}}\beta _0^{}{\displaystyle \frac{g_{}^{}{}_{}{}^{3}(\overline{\mu }^2)}{8\pi ^2}}`$ (41) with the one-loop terms given by: $$\beta _0=\frac{11}{12}C_A\frac{1}{3}n_{gen}\frac{1}{24}n_h,\beta _0^{}=\frac{5}{9}n_{gen}\frac{1}{24}n_h$$ (42) where $`n_{gen}`$ denotes the number of fermion generations and $`n_h`$ the number of Higgs doublets. The infrared singular anomalous dimensions read $$\mathrm{\Gamma }_{f,g}^i(t)=\left(\frac{\alpha }{4\pi }T_i(T_i+1)+\frac{\alpha ^{}}{4\pi }\left(\frac{Y_i}{2}\right)^2\right)t$$ (43) where $`T_i`$ and $`Y_i`$ are the total weak isospin and hypercharge respectively of the particle emitting the soft and collinear gauge boson. Analogously, $$\gamma _{q\overline{q}}^i=3\left(\frac{\alpha }{4\pi }T_i(T_i+1)+\frac{\alpha ^{}}{4\pi }\left(\frac{Y_i}{2}\right)^2\right)$$ (44) The initial condition for Eq. (39) is given by the requirement that for the infrared cutoff $`\mu ^2=s`$ we obtain the Born amplitude. The solution of (39) is thus given by $`^{}(p_1,\mathrm{},p_n,g,g^{},\mu )=_{\mathrm{Born}}^{}(p_1,\mathrm{},p_n,g(s),g^{}(s))`$ $`\times \mathrm{exp}\{{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{n_g}{}}}({\displaystyle \frac{\alpha (s)}{4\pi }}T_i(T_i+1)+{\displaystyle \frac{\alpha ^{}(s)}{4\pi }}\left({\displaystyle \frac{Y_i}{2}}\right)^2)\mathrm{log}^2{\displaystyle \frac{s}{\mu ^2}}`$ $`+\left(n_\mathrm{W}{\displaystyle \frac{\alpha (s)}{2\pi }}\beta _0+n_\mathrm{B}{\displaystyle \frac{\alpha ^{}(s)}{2\pi }}\beta _0^{}\right)\mathrm{log}{\displaystyle \frac{s}{\mu ^2}}`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{k=1}{\overset{n_f}{}}}({\displaystyle \frac{\alpha (s)}{4\pi }}T_k(T_k+1)+{\displaystyle \frac{\alpha ^{}(s)}{4\pi }}\left({\displaystyle \frac{Y_k}{2}}\right)^2)[\mathrm{log}^2{\displaystyle \frac{s}{\mu ^2}}3\mathrm{log}{\displaystyle \frac{s}{\mu ^2}}]\}`$ (45) where $`n_\mathrm{W}`$ and $`n_\mathrm{B}`$ denote the number of external $`W`$ and $`B`$ fields respectively. The $`SU(2)\times U(1)`$ group factors in the exponential can be written in terms of the parameters of the broken theory as follows: $$g^2T_i(T_i+1)+g_{}^{}{}_{}{}^{2}\left(\frac{Y_i}{2}\right)^2=e_i^2+g^2\left(T_i(T_i+1)(T_i^3)^2\right)+\frac{g^2}{\mathrm{cos}^2\theta _w}\left(T_i^3\mathrm{sin}^2\theta _wQ_i\right)^2,$$ where the three terms on the r.h.s. correspond to the contributions of the soft photon (interacting with the electric charge $`e_i=Q_ig\mathrm{sin}\theta _w`$), the $`W^\pm `$ and the $`Z`$ bosons, respectively. Although we may rewrite solution (45) in terms of the parameters of the broken theory in the form of a product of three exponents corresponding to the exchanges of photons, $`W^\pm `$ and $`Z`$ bosons, it would be wrong to identify the contributions of the diagrams without virtual photons with this expression for the particular case $`e_i^2=0`$. This becomes evident when we note that if we were to omit photon lines then the result would depend on the choice of gauge, and therefore be unphysical. Only for $`\theta _w=0`$, where the photon coincides with the $`B`$ gauge boson, would the identification of the $`e_i^2`$ term with the contribution of the diagrams with photons be correct. We now need to discuss the solution in the general case. In region 1) we calculated the scattering amplitude for the theory in the unbroken phase in the massless limit. Choosing the cutoff $`\mu `$ in region 2), $`\mu M`$, we have to only consider the photon contribution. In this region we cannot necessarily neglect all mass terms, so we need to discuss the subleading terms for QED with mass effects. If $`m_i\mu `$, the results from the massless QCD discussion of section 2.3 can be used directly by using the Abelian limit $`C_F=1`$. In case $`\mu m_i`$ we must use the well known next to leading order QED results, e.g. , and the virtual probabilities take the following form for fermions: $$w_i^f(s,\mu ^2)=\{\begin{array}{cc}\frac{e_i^2}{(4\pi )^2}\left(\mathrm{log}^2\frac{s}{\mu ^2}3\mathrm{log}\frac{s}{\mu ^2}\right)\hfill & ,m_i\mu \\ \frac{e_i^2}{(4\pi )^2}[(\mathrm{log}\frac{s}{m_i^2}1)2\mathrm{log}\frac{m_i^2}{\mu ^2}\hfill & \\ +\mathrm{log}^2\frac{s}{m_i^2}3\mathrm{log}\frac{s}{m_i^2}]\hfill & ,\mu m_i\end{array}$$ (46) Note, that in the last equation the full subleading collinear logarithmic term is used in distinction to Ref. . In the explicit two loop calculation presented in Ref. it can be seen that the full collinear term also exponentiates at the subleading level in massive QED. For $`W^\pm `$ bosons we have analogously: $$w_i^\mathrm{w}(s,\mu ^2)=\frac{e_i^2}{(4\pi )^2}\left[\left(\mathrm{log}\frac{s}{M^2}1\right)2\mathrm{log}\frac{M^2}{\mu ^2}+\mathrm{log}^2\frac{s}{M^2}\right]$$ (47) In addition we have collinear terms for external on-shell photon lines <sup>7</sup><sup>7</sup>7I thank the authors of Ref. for clarifying this point. from fermions with mass $`m_j`$ and electromagnetic charge $`e_j`$ up to scale $`M`$: $$w_i^\gamma (M^2,\mu ^2)=\{\begin{array}{cc}\frac{n_f}{3}\frac{e_j^2}{4\pi ^2}N_C^j\mathrm{log}\frac{M^2}{\mu ^2}\hfill & ,m_j\mu \\ \frac{1}{3}_{j=1}^{n_f}\frac{e_j^2}{4\pi ^2}N_C^j\mathrm{log}\frac{M^2}{m_j^2}\hfill & ,\mu m_j\end{array}$$ (48) Note that automatically, $`w_i^\gamma (M^2,M^2)=0`$. At one loop order, this contribution cancels against terms from the renormalization of the QED coupling up to scale $`M`$. For external $`Z`$-bosons, however, there are no such collinear terms since the mass is large compared to the $`m_i`$. Thus, the corresponding RG-logarithms up to scale $`M`$ remain uncanceled. The appropriate initial condition is given by Eq. (45) evaluated at the matching point $`\mu =M`$. Thus we find for the general solution in region 2): $`^{}(p_1,\mathrm{},p_n,g,g^{},\mu )=_{\mathrm{Born}}^{}(p_1,\mathrm{},p_n,g(s),g^{}(s))`$ $`\times \mathrm{exp}\{{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{n_g}{}}}({\displaystyle \frac{\alpha (s)}{4\pi }}T_i(T_i+1)+{\displaystyle \frac{\alpha ^{}(s)}{4\pi }}\left({\displaystyle \frac{Y_i}{2}}\right)^2)\mathrm{log}^2{\displaystyle \frac{s}{M^2}}`$ $`+\left(n_\mathrm{W}{\displaystyle \frac{\alpha (s)}{2\pi }}\beta _0+n_\mathrm{B}{\displaystyle \frac{\alpha ^{}(s)}{2\pi }}\beta _0^{}\right)\mathrm{log}{\displaystyle \frac{s}{M^2}}`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{k=1}{\overset{n_f}{}}}({\displaystyle \frac{\alpha (s)}{4\pi }}T_k(T_k+1)+{\displaystyle \frac{\alpha ^{}(s)}{4\pi }}\left({\displaystyle \frac{Y_k}{2}}\right)^2)[\mathrm{log}^2{\displaystyle \frac{s}{M^2}}3\mathrm{log}{\displaystyle \frac{s}{M^2}}]\}`$ $`\times \mathrm{exp}[{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{n_f}{}}}(w_i^f(s,\mu ^2)w_i^f(s,M^2)){\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{n_w}{}}}(w_i^\mathrm{w}(s,\mu ^2)w_i^\mathrm{w}(s,M^2))`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{n_\gamma }{}}}w_i^\gamma (M^2,\mu ^2)]`$ $`=_{\mathrm{Born}}^{}(p_1,\mathrm{},p_n,g(s),g^{}(s))`$ $`\times \mathrm{exp}\{{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{n_g}{}}}({\displaystyle \frac{\alpha (s)}{4\pi }}T_i(T_i+1)+{\displaystyle \frac{\alpha ^{}(s)}{4\pi }}\left({\displaystyle \frac{Y_i}{2}}\right)^2)\mathrm{log}^2{\displaystyle \frac{s}{M^2}}`$ $`+\left(n_\mathrm{W}{\displaystyle \frac{\alpha (s)}{2\pi }}\beta _0+n_\mathrm{B}{\displaystyle \frac{\alpha ^{}(s)}{2\pi }}\beta _0^{}\right)\mathrm{log}{\displaystyle \frac{s}{M^2}}`$ $`{\displaystyle \frac{1}{2}}{\displaystyle \underset{k=1}{\overset{n_f}{}}}({\displaystyle \frac{\alpha (s)}{4\pi }}T_k(T_k+1)+{\displaystyle \frac{\alpha ^{}(s)}{4\pi }}\left({\displaystyle \frac{Y_k}{2}}\right)^2)[\mathrm{log}^2{\displaystyle \frac{s}{M^2}}3\mathrm{log}{\displaystyle \frac{s}{M^2}}]\}`$ $`\times \mathrm{exp}[{\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{n}{}}}({\displaystyle \frac{e_i^2(s)}{(4\pi )^2}}(2\mathrm{log}{\displaystyle \frac{s}{m_iM}}\mathrm{log}{\displaystyle \frac{M^2}{m_i^2}}+2\mathrm{log}{\displaystyle \frac{s}{m_i^2}}\mathrm{log}{\displaystyle \frac{m_i^2}{\mu ^2}}+3\mathrm{log}{\displaystyle \frac{m_i^2}{M^2}}`$ $`2\mathrm{log}{\displaystyle \frac{m_i^2}{\mu ^2}})){\displaystyle \frac{1}{2}}{\displaystyle \underset{i=1}{\overset{n_\gamma }{}}}{\displaystyle \frac{1}{3}}{\displaystyle \frac{e_i^2}{4\pi ^2}}N_C^i\mathrm{log}{\displaystyle \frac{M^2}{m_i^2}}]`$ (49) The last equality holds for $`\mu m_iM`$ and we have absorbed all $`\beta `$-function terms not related to external lines into redefinitions of the scales of the couplings. It is important to note again that, unlike the situation in QCD, in the electroweak theory we have in general different mass scales determining the running of the couplings of the physical on-shell renormalization scheme quantities. We have written the above result in such a way that it holds for arbitrary chiral fermions and transversely polarized gauge bosons. In order to include physical external photon states in the on-shell scheme, the renormalization condition is given by the requirement that the physical photon does not mix with the Z-boson. This leads to the condition that the Weinberg rotations in Fig. 4 at one loop receive no RG-corrections. Thus, above the scale $`M`$ the subleading collinear and RG-corrections cancel for physical photon and Z-boson states. For physical observables, soft real photon emission must be taken into account in an inclusive (or semi inclusive) way and the parameter $`\mu ^2`$ in (49) will be replaced by parameters depending on the experimental requirements. This will be briefly discussed in the following section. #### 3.1.1 Semi inclusive cross sections In order to make predictions for observable cross sections, the unphysical infrared cutoff $`\mu ^2`$ has to be replaced with a cutoff $`\mu _{exp}^2`$, related to the lower bound of $`𝒌_{}^2`$ of the other virtual particles of those gauge bosons emitted in the process which are not included in the cross section. We assume that $`\mu _{exp}^2<M^2`$, so that the non-Abelian component of the photon is not essential. The case $`\mu _{exp}^2>M^2`$ is much more complicated and is discussed in Ref. through two loops at the DL level. We again only discuss transversely polarized external gauge boson in the Born process and can write the expression for the semi-inclusive cross section: $`d\sigma ^{}(p_1,\mathrm{},p_n,g,g^{},\mu _{exp})`$ $`=`$ $`d\sigma _{\mathrm{elastic}}^{}(p_1,\mathrm{},p_n,g(s),g^{}(s),\mu )`$ (50) $`\times \mathrm{exp}(w_{\mathrm{exp}}^\gamma (s,m_i,\mu ,\mu _{exp}))`$ In the soft photon approximation we have: $`w_{\mathrm{exp}}^\gamma (s,m_i,\mu ,\mu _{exp})`$ $`=`$ $`\{\begin{array}{cc}\frac{e_i^2}{(4\pi )^2}_{i=1}^n\left[\mathrm{log}^2\frac{s}{\mu _{exp}^2}+\mathrm{log}^2\frac{s}{\mu ^2}3\mathrm{log}\frac{s}{\mu ^2}\right]\hfill & ,m_i\mu \\ \frac{e_i^2}{(4\pi )^2}_{i=1}^n[(\mathrm{log}\frac{s}{m_i^2}1)2\mathrm{log}\frac{m_i^2}{\mu ^2}+\mathrm{log}^2\frac{s}{m_i^2}\hfill & \\ 2\mathrm{log}\frac{s}{\mu _{exp}^2}(\mathrm{log}\frac{s}{m_i^2}1)]\hfill & ,\mu m_i\end{array}`$ (54) where the upper case applies only to fermions since for $`W^\pm `$ we have $`\mu <M`$ in region 2). Since the upper bound on $`𝒌_{}^2`$ of the photons which are allowed to be radiated is less than $`M^2`$, we must use the cut-off $`\mu ^2<M^2`$ and, consequently, (49) for the matrix element of the non-radiative process. Therefore, we obtain<sup>8</sup><sup>8</sup>8The notation here is again simplified in the sense that for $`Z`$-boson and $`\gamma `$ final states one has to include the mixing correctly as described above. $`d\sigma ^{}(p_1,\mathrm{},p_n,g,g^{},\mu _{exp})=d\sigma _{\mathrm{Born}}^{}(p_1,\mathrm{},p_n,g(s),g^{}(s))`$ $`\times \mathrm{exp}\{{\displaystyle \underset{i=1}{\overset{n_g}{}}}({\displaystyle \frac{\alpha (s)}{4\pi }}T_i(T_i+1)+{\displaystyle \frac{\alpha ^{}(s)}{4\pi }}\left({\displaystyle \frac{Y_i}{2}}\right)^2)\mathrm{log}^2{\displaystyle \frac{s}{M^2}}`$ $`+\left(n_\mathrm{W}{\displaystyle \frac{\alpha (s)}{\pi }}\beta _0+n_\mathrm{B}{\displaystyle \frac{\alpha ^{}(s)}{\pi }}\beta _0^{}\right)\mathrm{log}{\displaystyle \frac{s}{M^2}}`$ $`{\displaystyle \underset{k=1}{\overset{n_f}{}}}({\displaystyle \frac{\alpha (s)}{4\pi }}T_k(T_k+1)+{\displaystyle \frac{\alpha ^{}(s)}{4\pi }}\left({\displaystyle \frac{Y_k}{2}}\right)^2)[\mathrm{log}^2{\displaystyle \frac{s}{M^2}}3\mathrm{log}{\displaystyle \frac{s}{M^2}}]\}`$ $`\times \mathrm{exp}[{\displaystyle \underset{i=1}{\overset{n_f}{}}}(w_i^f(s,\mu ^2)w_i^f(s,M^2)){\displaystyle \underset{i=1}{\overset{n_w}{}}}(w_i^\mathrm{w}(s,\mu ^2)w_i^\mathrm{w}(s,M^2))`$ $`{\displaystyle \underset{i=1}{\overset{n_\gamma }{}}}w_i^\gamma (M^2,m_j^2)]`$ $`\times \mathrm{exp}({\displaystyle \frac{e_i^2}{(4\pi )^2}}{\displaystyle \underset{i=1}{\overset{n}{}}}[(\mathrm{log}{\displaystyle \frac{s}{m_i^2}}1)2\mathrm{log}{\displaystyle \frac{m_i^2}{\mu ^2}}2\mathrm{log}{\displaystyle \frac{s}{\mu _{exp}^2}}(\mathrm{log}{\displaystyle \frac{s}{m_i^2}}1)`$ $`+\mathrm{log}^2{\displaystyle \frac{s}{m_i^2}}])`$ (56) where we use $`\mu m_i`$. The $`\mu `$ dependence in this expression cancels and the semi-inclusive cross section depends only on the parameters of the experimental requirements. Eq. (56) contains all leading double and single logarithms to cross sections<sup>9</sup><sup>9</sup>9We emphasize again that we did not consider angular logarithms which can be sizable and should be calculated at least to one loop order. containing arbitrary numbers of external fermions and transversely polarized gauge bosons. We have only assumed that all masses are not larger than the electroweak scale $`M`$ and impose a cut on the the allowed values of emitted real gauge bosons $`𝒌_{}^2\mu _{exp}^2<M^2`$, i.e. up to the weak scale we only need to consider real QED effects. ### 3.2 Longitudinal degrees of freedom In this section we discuss if results obtained from the massless unbroken phase of the $`SU(2)\times U(1)`$ theory, where due to gauge invariance we have only transverse physical degrees of freedom, can be extended to the full theory including longitudinal vector bosons. This point of discussion is necessary and important since the longitudinal degrees of freedom don’t decouple at high energies and could give crucial clues to potentially strong dynamical effects for large Higgs masses $`m_H1TeV`$ . The connection between the strategy pursued for the transverse degrees of freedom and the corrections to longitudinally polarized vector bosons at high energies is provided by the Goldstone boson equivalence theorem . It states that at tree level for S-matrix elements for longitudinal bosons at the high energy limit $`M^2/s0`$ can be expressed through matrix elements involving their associated would be Goldstone bosons. We write schematically in case of a single gauge boson: $`(W_L^\pm ,\psi _{\mathrm{phys}})`$ $`=`$ $`(\varphi ^\pm ,\psi _{\mathrm{phys}})+𝒪\left({\displaystyle \frac{M_\mathrm{w}}{\sqrt{s}}}\right)`$ (57) $`(Z_L,\psi _{\mathrm{phys}})`$ $`=`$ $`(\varphi ,\psi _{\mathrm{phys}})+𝒪\left({\displaystyle \frac{M_\mathrm{z}}{\sqrt{s}}}\right)`$ (58) The problem with this statement of the equivalence theorem is that it holds only at tree level . For calculations at higher orders, additional terms enter which change Eqs. (57) and (58). Because of the gauge invariance of the physical theory and the associated BRST invariance, a modified version of Eqs. (57) and (58) can be derived which reads $`k^\nu (W_\nu ^\pm (k),\psi _{\mathrm{phys}})`$ $`=`$ $`C_\mathrm{w}M_\mathrm{w}(\varphi ^\pm (k),\psi _{\mathrm{phys}})+𝒪\left({\displaystyle \frac{M_\mathrm{w}}{\sqrt{s}}}\right)`$ (59) $`k^\nu (Z_\nu (k),\psi _{\mathrm{phys}})`$ $`=`$ $`C_\mathrm{z}M_\mathrm{z}(\varphi (k),\psi _{\mathrm{phys}})+𝒪\left({\displaystyle \frac{M_\mathrm{z}}{\sqrt{s}}}\right)`$ (60) where the multiplicative factors $`C_\mathrm{w}`$ and $`C_\mathrm{z}`$ depend only on wave function renormalization constants and mass counterterms. Thus, using the form of the longitudinal polarization vector of Eq. (35) we can write $`(W_L^\pm (k),\psi _{\mathrm{phys}})`$ $`=`$ $`C_\mathrm{w}(\varphi ^\pm (k),\psi _{\mathrm{phys}})+𝒪\left({\displaystyle \frac{M_\mathrm{w}}{\sqrt{s}}}\right)`$ (61) $`(Z_L(k),\psi _{\mathrm{phys}})`$ $`=`$ $`C_\mathrm{z}(\varphi (k),\psi _{\mathrm{phys}})+𝒪\left({\displaystyle \frac{M_\mathrm{z}}{\sqrt{s}}}\right)`$ (62) Thus we see that in principle, there are logarithmic loop corrections to the tree level equivalence theorem<sup>10</sup><sup>10</sup>10 An exception is the background field gauge where the Ward-identities guarantee that the factors $`C_\mathrm{w}=1`$ and $`C_\mathrm{z}=1`$ to all orders . It should thus be investigated if subleading corrections can also be obtained from the Goldstone boson equivalence theorem.. In addition, for longitudinal gauge bosons we also have logarithmic corrections with Yukawa terms . On the one hand, this means that the method of section 3.1 should be used with caution to obtain the relevant subleading terms. Thus we should consider these corrections separately. On the other, since the corrections are at most logarithmic, it means that the results of Ref. can be extended to the longitudinal sector as well. Thus we find<sup>11</sup><sup>11</sup>11For longitudinally polarized $`Z`$-boson final states there are no mixing terms since the photon has only transverse polarization states. Thus one needs to only include the associated Goldstone boson $`\varphi `$ at the DL level. for $`\mu _{exp}<M`$: $`d\sigma ^{}(p_1,\mathrm{},p_n,g,g^{},\mu _{exp})=d\sigma _{\mathrm{Born}}^\varphi (p_1,\mathrm{},p_n,g,g^{})`$ $`\times \mathrm{exp}\left\{{\displaystyle \underset{i=1}{\overset{n}{}}}\left({\displaystyle \frac{\alpha }{4\pi }}T_i(T_i+1)+{\displaystyle \frac{\alpha ^{}}{4\pi }}\left({\displaystyle \frac{Y_i}{2}}\right)^2\right)\mathrm{log}^2{\displaystyle \frac{s}{M^2}}\right\}`$ $`\times \mathrm{exp}\left[{\displaystyle \underset{i=1}{\overset{n}{}}}\left(w_i^{\mathrm{DL}}(s,\mu ^2)w_i^{\mathrm{DL}}(s,M^2)\right)+w_{exp}^{\mathrm{DL}}\right]`$ (63) where the index $``$ indicates the cross section for longitudinally polarized gauge bosons, while the field $`\varphi `$ indicates that the appropriate fields and quantum numbers on the r.h.s. in Eq. (63) are those of the associated would be Goldstone bosons. Thus, we have shown that all DL corrections can be summed to all orders by employing the evolution equation approach of Ref. in connection with the Goldstone boson equivalence theorem. ## 4 Comparison with explicit results In this section we compare our results obtained in the previous sections with known results in special cases and one loop calculations. In Ref. , QCD-results for the Sudakov form factor were generalized to the high energy electroweak theory<sup>12</sup><sup>12</sup>12In addition, angular terms at the one loop level were calculated which we do not consider in this work.. Since the general strategy pursued is the same as in Ref. , we of course agree with their result for leading and subleading electroweak corrections to $`e^+e^{}f\overline{f}`$ to all orders. A very important check is provided by the explicit one-loop corrections of Ref. for high energy on-shell $`W`$-pair production in the soft photon approximation. In the following, the lower index on the cross section indicates the helicity of the electron, where $`e_{}^{}`$ denotes the left handed electron. We summarize the relevant results for $`e_+^+e_{}^{}W_{}^+W_{}^{}`$, $`e_+^+e_{}^{}W_{}^+W_{}^{}`$ and $`e_{}^+e_+^{}W_{}^+W_{}^{}`$ for convenience as follows: $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_,`$ $``$ $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_,^{\mathrm{Born}}\{1+{\displaystyle \frac{e^2}{8\pi ^2}}[{\displaystyle \frac{1+2c_\mathrm{w}^2+8c_\mathrm{w}^4}{4c_\mathrm{w}^2s_\mathrm{w}^2}}\mathrm{log}^2{\displaystyle \frac{s}{M^2}}+3{\displaystyle \frac{12c_\mathrm{w}^2+4c_\mathrm{w}^4}{4c_\mathrm{w}^2s_\mathrm{w}^2}}\mathrm{log}{\displaystyle \frac{s}{M^2}}`$ (64) $`+3\mathrm{log}{\displaystyle \frac{s}{m_e^2}}+2\mathrm{log}{\displaystyle \frac{4\mathrm{\Delta }E^2}{s}}\left(\mathrm{log}{\displaystyle \frac{s}{m_e^2}}+\mathrm{log}{\displaystyle \frac{s}{M^2}}2\right)`$ $`{\displaystyle \frac{4}{3}}{\displaystyle \underset{j=1}{\overset{n_f}{}}}Q_j^2N_C^j\mathrm{log}{\displaystyle \frac{m_j^2}{M^2}}]\}`$ $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_,`$ $``$ $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_,^{\mathrm{Born}}\{1+{\displaystyle \frac{e^2}{8\pi ^2}}[{\displaystyle \frac{12c_\mathrm{w}^2+4c_\mathrm{w}^4}{2c_\mathrm{w}^2s_\mathrm{w}^2}}\mathrm{log}^2{\displaystyle \frac{s}{M^2}}`$ (65) $`+2\mathrm{log}{\displaystyle \frac{4\mathrm{\Delta }E^2}{s}}(\mathrm{log}{\displaystyle \frac{s}{m_e^2}}+\mathrm{log}{\displaystyle \frac{s}{M^2}})]\}`$ $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_{+,}`$ $``$ $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_{+,}^{\mathrm{Born}}\{1+{\displaystyle \frac{e^2}{8\pi ^2}}[{\displaystyle \frac{510c_\mathrm{w}^2+8c_\mathrm{w}^4}{4c_\mathrm{w}^2s_\mathrm{w}^2}}\mathrm{log}^2{\displaystyle \frac{s}{M^2}}`$ (66) $`+2\mathrm{log}{\displaystyle \frac{4\mathrm{\Delta }E^2}{s}}(\mathrm{log}{\displaystyle \frac{s}{m_e^2}}+\mathrm{log}{\displaystyle \frac{s}{M^2}})]\}`$ where the last line in Eq. (64) corresponds to a sum over all fermions contributing to the coupling renormalization (with multiplicity $`N_C=3`$ for quarks and $`N_C=1`$ for leptons). This contribution can be included in the scale of the running on-shell charge $`\alpha _{\mathrm{eff}}(M^2)`$ . For the longitudinal cross sections we are only concerned with DL corrections. The Born cross sections are given by: $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_,^{\mathrm{Born}}`$ $`=`$ $`{\displaystyle \frac{e^4}{64\pi ^2s}}{\displaystyle \frac{1}{4s_\mathrm{w}^4}}{\displaystyle \frac{u^2+t^2}{t^2}}\mathrm{sin}^2\theta `$ (67) $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_,^{\mathrm{Born}}`$ $`=`$ $`{\displaystyle \frac{e^4}{64\pi ^2s}}{\displaystyle \frac{1}{16s_\mathrm{w}^4c_\mathrm{w}^4}}\mathrm{sin}^2\theta `$ (68) $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_{+,}^{\mathrm{Born}}`$ $`=`$ $`{\displaystyle \frac{e^4}{64\pi ^2s}}{\displaystyle \frac{1}{4c_\mathrm{w}^4}}\mathrm{sin}^2\theta `$ (69) where we keep the angular dependence. In Eq. (67) a sum over the two transverse polarizations of the $`W^\pm `$ ($`++`$ and $``$) is implicit. These expressions demonstrate that the longitudinal cross sections in Eqs. (68) and (69) are not suppressed with respect to Eq. (67). On the other hand, $`\left(\frac{d\sigma }{d\mathrm{\Omega }}\right)_{+,}^{\mathrm{Born}}`$ is mass suppressed . Eqs. (64), (65) and (66) were of course calculated in terms of the physical fields of the broken theory and in the on-shell scheme. We denote $`c_\mathrm{w}=\mathrm{cos}\theta _\mathrm{w}`$ and $`s_\mathrm{w}=\mathrm{sin}\theta _\mathrm{w}`$ respectively. In order to compare with the results of section 3 we listed the relevant quantum numbers in Table 1. For comparison with Eq. (64) and to logarithmic accuracy, we can absorb the running coupling effects from our massless scheme to the on-shell scheme as follows. The Born cross section in our approach is proportional to $`g^4`$ (see Eq. (67)). The coupling renormalization above the scale $`M`$ is given by: $$g^2(s)=g^2(M^2)\left(1\frac{g^2(M^2)}{4\pi ^2}\left(\frac{11}{12}C_A\frac{1}{24}n_h\frac{n_{gen}}{3}\right)\mathrm{log}\frac{s}{M^2}\right)$$ (70) Below the scale where non-Abelian effects enter, the running is only due to the electromagnetic coupling and we write $`g^2(M^2)=\frac{e_{\mathrm{eff}}^2(M^2)}{s_\mathrm{w}^2}`$ with $$e_{\mathrm{eff}}^2(M^2)=e^2\left(1+\frac{1}{3}\frac{e^2}{4\pi ^2}\underset{j=1}{\overset{n_f}{}}Q_j^2N_C^j\mathrm{log}\frac{M^2}{m_j^2}\right)$$ (71) We therefore observe that the running coupling terms proportional to $`\mathrm{log}\frac{s}{M^2}`$ cancel for this process with the subleading contributions from the virtual splitting functions (see Eq. (49)) and what remains are just the Abelian terms up to scale $`M`$. Thus for Eq. (64) we obtain from Eq. (56) at the one loop level: $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_,`$ $`=`$ $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_,^{\mathrm{Born}}\{1({\displaystyle \frac{g^2}{8\pi ^2}}T_\mathrm{w}(T_\mathrm{w}+1)+{\displaystyle \frac{g_{}^{}{}_{}{}^{2}}{8\pi ^2}}{\displaystyle \frac{Y_\mathrm{w}^2}{4}})\mathrm{log}^2{\displaystyle \frac{s}{M^2}}`$ (72) $`({\displaystyle \frac{g^2}{8\pi ^2}}T_e_{}^{}(T_e_{}^{}+1)+{\displaystyle \frac{g_{}^{}{}_{}{}^{2}}{8\pi ^2}}{\displaystyle \frac{Y_e_{}^{}^2}{4}})(\mathrm{log}^2{\displaystyle \frac{s}{M^2}}3\mathrm{log}{\displaystyle \frac{s}{M^2}}){\displaystyle \frac{e^2}{8\pi ^2}}\times `$ $`[(\mathrm{log}{\displaystyle \frac{s}{m_e^2}}1)2\mathrm{log}{\displaystyle \frac{m_e^2}{\mu ^2}}+\mathrm{log}^2{\displaystyle \frac{s}{m_e^2}}3\mathrm{log}{\displaystyle \frac{s}{m_e^2}}\mathrm{log}^2{\displaystyle \frac{s}{M^2}}+3\mathrm{log}{\displaystyle \frac{s}{M^2}}`$ $`+2\left(\mathrm{log}{\displaystyle \frac{s}{M^2}}1\right)\mathrm{log}{\displaystyle \frac{M^2}{\mu ^2}}\left(\mathrm{log}{\displaystyle \frac{s}{m_e^2}}1\right)\left(2\mathrm{log}{\displaystyle \frac{m_e^2}{\mu ^2}}2\mathrm{log}{\displaystyle \frac{s}{\mu _{exp}^2}}\right)`$ $`2(\mathrm{log}{\displaystyle \frac{s}{M^2}}1)(\mathrm{log}{\displaystyle \frac{M^2}{\mu ^2}}\mathrm{log}{\displaystyle \frac{s}{\mu _{exp}^2}})\mathrm{log}^2{\displaystyle \frac{s}{m_e^2}}\mathrm{log}^2{\displaystyle \frac{s}{M^2}}]`$ $`+{\displaystyle \frac{2}{3}}{\displaystyle \frac{e^2}{4\pi ^2}}{\displaystyle \underset{j=1}{\overset{n_f}{}}}Q_j^2N_C^j\mathrm{log}{\displaystyle \frac{M^2}{m_j^2}}\}`$ $`=`$ $`\left({\displaystyle \frac{d\sigma }{d\mathrm{\Omega }}}\right)_,^{\mathrm{Born}}\{1{\displaystyle \frac{e^2}{8\pi ^2}}({\displaystyle \frac{1+10c_\mathrm{w}^2}{4s_\mathrm{w}^2c_\mathrm{w}^2}}\mathrm{log}^2{\displaystyle \frac{s}{M^2}}3{\displaystyle \frac{1+2c_\mathrm{w}^2}{4s_\mathrm{w}^2c_\mathrm{w}^2}}\mathrm{log}{\displaystyle \frac{s}{M^2}})+{\displaystyle \frac{e^2}{8\pi ^2}}\times `$ $`\left[2\mathrm{log}^2{\displaystyle \frac{s}{M^2}}3\mathrm{log}{\displaystyle \frac{m_e^2}{M^2}}4\mathrm{log}{\displaystyle \frac{s}{\mu _{exp}^2}}\left(\mathrm{log}{\displaystyle \frac{s}{m_eM}}1\right)\right]`$ $`+{\displaystyle \frac{2}{3}}{\displaystyle \frac{e^2}{4\pi ^2}}{\displaystyle \underset{j=1}{\overset{n_f}{}}}Q_j^2N_C^j\mathrm{log}{\displaystyle \frac{M^2}{m_j^2}}\}`$ Eq. (72) agrees with Eq. (64), which are both valid in the soft photon approximation. Here and below we assume that $`\mathrm{\Delta }E<M`$ and $`\mu _{exp}<M`$. Analogously in the DL approximation, it is straightforward to check the validity of our results for Eqs. (65) and (66), emphasizing again that in this case we need to use the quantum numbers of the associated Goldstone bosons, see Fig. 5 and Tab. 1. Thus we have verified that our results, calculated in terms of the unbroken massless fields, give the correct leading and subleading logarithms in transversely polarized $`W`$-pair production at the one loop level. For longitudinally polarized $`W`$-pairs the correct DL asymptotics is reproduced. ## 5 Conclusions In this paper we considered the calculation of virtual next to leading electroweak corrections at energies much larger than the electroweak scale when all particle masses can be neglected. We follow the same approach as in Ref. which consists of using the fields of the unbroken theory to obtain logarithmic corrections with the infrared evolution equation method in different regions of the infrared cutoff. When particle masses can be neglected there is a one to one correspondence between the high and low energy scaling behavior and the evolution equation can be formulated in terms of the renormalization group with infrared singular anomalous dimension. The next to leading kernel can then be obtained from the virtual contribution to the Altarelli-Parisi splitting functions. For external gauge boson emission one can use the above approach for transverse degrees of freedom. For fermions and $`W^\pm `$ external states, the next to leading corrections exponentiate with respect to the physical Born amplitudes. For $`Z`$ boson and $`\gamma `$ final states, one needs to include the effect of mixing appropriately. For these final states we have exponentiation with respect to the amplitudes containing the fields of the unbroken theory but not with respect to the physical Born amplitude. For longitudinal degrees of freedom, one can use the Goldstone-boson equivalence theorem to obtain the correct DL asymptotic behavior. These terms are found to exponentiate as well. Loop corrections, however, lead to additional corrections including Yukawa terms at the subleading level and should be considered separately. We also compared our results with exact one loop calculations in the high energy approximation and could reproduce the leading and subleading terms for transversely polarized $`W`$-pair production in $`e^+e^{}`$ collisions and the DL corrections for the longitudinal degrees of freedom. Finally we note that there are of course terms which we have neglected in this analysis. As mentioned above, there are angular logarithms of the form $`\mathrm{log}\frac{s}{M^2}\mathrm{log}\frac{t}{u}`$, which in general could be significant and should be computed separately. We also omitted all mass-logarithms of the form $`\mathrm{log}\frac{s}{M^2}\mathrm{log}\frac{M_Z}{M_W}`$ and top-Yukawa terms. For the latter, it might be possible to include them consistently into the virtual splitting functions. For longitudinal degrees of freedom it would also be very helpful to calculate subleading corrections. In conclusion, for future high energy experiments in the multi-TeV energy regime, the leading high energy behavior of general scattering amplitudes can be an important ingredient to study the effect of new physics expected in precisely this range. ## Acknowledgements We would like to thank A. Denner and S. Pozzorini for valuable discussions, especially concerning mixing effects and corrections to the Goldstone boson equivalence theorem. We would also like to thank W. Beenakker, V.S. Fadin and D. Graudenz for consultations.
warning/0004/quant-ph0004047.html
ar5iv
text
# Quantum Noise and Superluminal Propagation ## 1 Introduction Chiao et al. ,, have shown that certain “superluminal” effects are possible without violation of standard notions of Einstein causality, i.e., without conveying information faster than the velocity $`c`$ of light in vacuum. Such effects have been demonstrated experimentally in optical tunneling ,, and in an electric circuit . It has been suggested by Chiao, Kozhekin, and Kurizki (CKK) that an optical pulse can propagate superluminally in an amplifier whose relaxation times are long compared with the pulse duration. The dispersion relation they derive can be obtained directly, as follows, starting from the formula for the refractive index of a monatomic gas: $$n(\omega )=1+\frac{2\pi e^2}{m}\underset{i}{}\underset{j}{}\frac{N_if(i,j)}{\omega _{ji}^2\omega ^2}$$ (1) for $`n(\omega )1`$, where $`N_i`$ is the number density of atoms in state $`i`$ and $`f(i,j)`$ is the oscillator strength for absorption on the $`ij`$ transition of frequency $`\omega _{ji}`$. Near a two–level resonance this becomes $$n(\omega )=1+\frac{2\pi e^2f}{m}\frac{N_1N_2}{\omega _o^2\omega ^2},$$ (2) where 1 and 2 designate the lower– and upper–energy levels, respectively, and $`\omega _o=\omega _{21}`$. Close to the transition resonance frequency $`\omega _o`$, $$n(\omega )1+\frac{\pi e^2f}{m\omega _o}\frac{N_1N_2}{\omega _o\omega i\beta }$$ (3) when we include a dipole damping rate $`\beta `$. The (real) refractive index near a resonance is then $$n_R(\omega )=1+\frac{\pi e^2f}{m\omega _o}\frac{\omega _o\omega }{(\omega _o\omega )^2+\beta ^2}(N_1N_2).$$ (4) Introducing the inversion $`w=(N_2N_1)/N`$, where $`N`$ is the number density of atoms, and assuming a field sufficiently far from resonance that $`(\omega _o\omega )^2>>\beta ^2`$, we have $$n_R(\omega )1\frac{\pi e^2Nwf}{m\omega _o}\frac{1}{\omega _o\omega },$$ (5) $$k=n_R(\omega )\frac{\omega }{c}=\frac{\omega }{c}\left[1\frac{\pi e^2Nwf}{m\omega _o}\frac{1}{\omega _o\omega }\right]=\frac{\omega }{c}\left[1\frac{\omega _p^2w/4\omega _o}{\omega _o\omega }\right],$$ (6) $$kk_o=\frac{1}{c}(\omega \omega _o)\frac{\omega }{c}\frac{\omega _p^2w/4\omega _o}{\omega _o\omega }\frac{1}{c}(\omega \omega _o)\frac{\omega _p^2w/4c}{\omega _o\omega },$$ (7) and $$\mathrm{\Omega }^2Kc\mathrm{\Omega }+\frac{1}{4}w\omega _p^2=0,$$ (8) where $`K=kk_o`$, $`\mathrm{\Omega }=\omega \omega _o`$, and the “plasma frequency” $`\omega _p`$ is defined by $$\omega _p^2=4\pi Ne^2f/m=8\pi Nd^2\omega _o/\mathrm{},$$ (9) with $`d`$ the electric dipole transition moment. Equation (8) is the dispersion relation obtained by CKK. We refer the reader to the CKK paper for a discussion of this dispersion relation. Here we simply note that (7) implies the group velocity $$v_g=\frac{d\omega }{dk}=c\left[1\frac{\omega _p^2w/4}{(\omega _o\omega )^2}\right]^1,$$ (10) so that, in the case of an amplifier ($`w>0`$), a short off-resonant pulse can propagate with a group velocity $`v_g>c`$. Questions have been raised about the validity of the latter prediction at the one-photon level, which would correspond to what CKK call an “optical tachyon” . Aharonov, Reznik, and Stern (ARS) have presented general arguments, based on the unitary evolution of the state vector, that “strongly questions the possibility that these systems may have tachyonlike quasiparticle excitations made up of a small number of photons.” They also consider a particular model as an analog of the CKK system. In this paper we address the question of superluminal propagation at the one- or few-photon level, and in particular the role played by quantum noise in the propagation of such extremely weak pulses. We begin in the following section with some physical considerations about the observability of superluminal propagation, and we briefly compare the ARS and CKK models. In Section 3 we formulate the Heisenberg equations of motion for the propagation of a short optical pulse in an inverted medium, and briefly review some relevant results from the theory of superfluorescence (SF). In Section 4 we derive a signal-to-noise ratio for the case where an incident, Gaussian signal pulse made up of $`q`$ photons is very short compared with the radiative lifetime and has a central frequency far removed from the resonance frequency of the medium. If we impose the ARS criterion for distinguishing between superluminal propagation and propagation at the speed of light, we find, consistent with their conclusions, that the signal must be “exponentially large” in order to distinguish it from quantum noise. If the ARS criterion is replaced by a much weaker one, however, the signal-to-noise ratio can exceed unity even for a one-photon signal pulse, as suggested by CKK. We relate the amplified quantum field fluctuations of ARS to quantum fluctuations of the atomic dipoles in the case of the optical amplifier. In Section 5, following the ideas of ARS, we present some general considerations based on the premises of unitarity and superluminal propagation. ARS show that, when the group velocity exceeds the speed of light, the superluminal signal is reconstructed from a truncated initial wavepacket, and that this truncated wavepacket has unstable modes. We show that the truncated wavepacket introduced by ARS propagates with both luminal and superluminal parts, and that, while the superluminal part is the reconstructed signal, the luminal part has the exponentially growing parts corresponding to the unstable modes. In addition, we study the residual wavepacket formed by the difference of the complete and truncated wavepackets. We show that contributions from the truncated and residual wavepackets cancel in the luminal region, but that, unlike the signal, the noise does not cancel, leading to the conclusion that the quantum noise is mostly luminal rather than superluminal. In the limit of a very weak incident signal pulse the signal-to-noise ratio will be very small, consistent with the conclusions reached by ARS. It may be worth recalling that a primary reason for rejecting the possibility of superluminal transmission of information is the requirement that causality be maintained when Lorentz transformations are made: superluminal transmission of information would allow an event A causing an event B in one reference frame to occur after event B in a different frame. Considerations of superluminal propagation therefore often raise questions relating to Lorentz invariance. When and how should one include relativistic effects in order to ensure that physically meaningful results are obtained? As in all previous treatments of pulse propagation in an inverted medium that we know of, we choose the reference frame in which the atoms are at rest. The Lorentz invariance of the fundamental, fully relativistic theory implies, of course, that our conclusions do not depend on this specific choice of a reference frame. Working in this frame, we treat the response of the atoms to the field in the approximation of nonrelativistic quantum mechanics. The electromagnetic field in this frame is also treated approximately, namely in the slowly varying envelope approximation that is used practically universally in the theory of resonant atom-field interactions. A different choice of reference frame would require us to start with the fully Lorentz-invariant equations and then make the slowly-varying-envelope and other approximations as appropriate. These approximations are known to be very accurate unless, for instance, the light pulse is extremely short, and to the extent that they are valid our results and conclusions are Lorentz-invariant. ## 2 Preliminary Considerations The quantum noise limitations to superluminal propagation discussed by ARS were associated physically with spontaneous emission in the case of an optical amplifier, and could invalidate the CKK results in two ways. First, CKK assume that the atoms stay in their excited states as the pulse propagates through the amplifier. Radiative decay of the excited state will modify their “tachyonic dispersion relation” and, if the decay is rapid enough, can lead to a subluminal rather than superluminal group velocity, since $`w`$ in equation (10) can become negative. This can be avoided by using a sufficiently short pulse. Second, spontaneously emitted radiation might interfere with the measurement of the superluminal group velocity by introducing substantial noise. It is this possibility that is addressed by ARS. Although the ARS arguments are certainly compelling, they are based in part on an analog of an optical amplifier rather than a theory involving the interaction of the electromagnetic field with an atomic medium. In particular, theirs is a model of a single quantum field rather than coupled atomic and electromagnetic quantum fields. The dispersion relation associated with this model, and the criteria assumed by ARS for the observability of superluminal propagation, lead to the conclusion, by analogy to an optical amplifier, that spontaneous emission noise cannot be avoided no matter how short the pulse or the transit time through the amplifier. Specifically, the unstable modes appearing in their model – which “are analogous to spontaneous emission in the optical model of an inverted medium of two-level systems” – will preclude the observation of superluminal group velocity when the pulse is made up of a small number of photons; the quantum noise will be larger than the signal. In this section we present some physical considerations, motivated by the CKK and ARS analyses, for the observability of superluminal group velocity. Following their equation (11), ARS state two necessary conditions for the observability of superluminal propagation \[$`c=1`$ in their units\]: (1) $`v_gT>>1/\delta k`$, where $`v_g`$ is the group velocity, $`T`$ is the time at which the wavepacket is observed, and $`\delta k`$ is the spectral width of their initial pulse. (2) $`(v_g1)T>>1/\delta k`$ . The first condition ensures that “the point of observation \[is\] far outside the initial spread of the wavepacket.” The second allows us to “distinguish between superluminal propagation and propagation at the speed of light.” In the ARS model, where the field $`\varphi `$ satisfies $$\frac{^2\varphi }{t^2}\frac{^2\varphi }{z^2}m^2\varphi =0,$$ (11) the group velocity is $$v_g=\frac{k_o}{\sqrt{k_o^2m^2}},$$ (12) where $`k_o`$ is the central value of the spatial frequency $`k`$ for the initial pulse. For $`m<k_o`$ we can approximate $`v_g`$ by $`1+m^2/2k_o^2`$, so that condition 2 \[and also condition 1\] is satisfied if $$m^2T>>k_o^2/\delta k>>k_o.$$ (13) $`k_o>>1/T`$ – the condition that the observation time should be much larger than the central frequency of the pulse – then implies $$mT>>1.$$ (14) Since for $`mT>>1`$ the amplified quantum noise grows exponentially \[see Section 3\], ARS conclude that the “signal amplitude should be exponentially large” in order to distinguish it from noise. Thus, according to ARS, the observability of superluminality for an input pulse consisting of only a few photons would be clouded by spontaneous emission noise. Consider now the implications of conditions 1 and 2 for the actual system of interest, namely a very short optical pulse in an inverted medium. Can we satisfy these conditions for observation times short compared with the radiative lifetime? For a short optical pulse of central frequency $`\omega `$ propagating in an inverted medium ($`w=1`$) with resonance frequency $`\omega _o`$, the refractive index is \[equation (6)\] $$n(\omega )1+\frac{2\pi Nd^2/\mathrm{}}{\omega \omega _o}1\frac{\omega _p^2}{4\omega _o\mathrm{\Delta }}$$ (15) for $`\omega _p^2/(4\omega _o)<<|\omega _o\omega ||\mathrm{\Delta }|`$. We are assuming that $`|\mathrm{\Delta }|`$ is large compared with the absorption width, which in our case is the radiative decay rate. Equation (15) implies $$\frac{v_g}{c}=\left(\frac{d}{d\omega }[\omega n(\omega )]\right)^1=\frac{1}{1\omega _p^2/4\mathrm{\Delta }^2}$$ (16) and $$\frac{v_g}{c}1=\frac{\omega _p^2/4\mathrm{\Delta }^2}{1\omega _p^2/4\mathrm{\Delta }^2}=\frac{\omega _p^2}{4\mathrm{\Delta }^2}\frac{v_g}{c}.$$ (17) Then conditions 1 and 2 of ARS become, respectively, $$\frac{T}{1\omega _p^2/4\mathrm{\Delta }^2}>>\frac{1}{c\delta k}\tau _p,$$ (18) $$\frac{(\omega _p^2/4\mathrm{\Delta }^2)T}{1\omega _p^2/4\mathrm{\Delta }^2}>>\frac{1}{c\delta k}\tau _p,$$ (19) with $`\tau _p`$ the pulse duration. Both conditions can be satisfied if, for instance, $`T>>\tau _p`$ and $`\omega _p^2/4\mathrm{\Delta }^2`$ is not too small. To avoid spontaneous emission during the observation time $`T`$, take $`T<<\tau _{RAD}`$, where $`\tau _{RAD}`$ is the radiative lifetime of a single inverted atom. Then the ARS conditions require that $$\tau _{RAD}>>T>>\tau _p.$$ (20) As noted by CKK, there is another aspect of an inverted atomic medium that must be addressed, namely superfluorescence (SF). SF is a collective phenomenon of the sample as a whole. We shall denote by $`N_T`$, $`S`$, and $`L`$ the number of atoms, the cross–sectional area, and the length of the sample, respectively, so that the density of atoms is given by $`N=N_T/SL`$. If collisional and other dephasing mechanisms are sufficiently weak, an inverted medium of $`N_T`$ atoms can emit SF radiation at the rate $$\tau _R=\tau _{RAD}/N_T,$$ (21) i.e., the radiative decay time can in effect be smaller by a factor of $`N_T`$ than the single–atom radiative lifetime $`\tau _{RAD}`$ assumed in the discussion thus far. The peak of the SF pulse occurs at a time $$\tau _D\tau _R\left[\frac{1}{4}\mathrm{ln}(2\pi N_T)\right]^2$$ (22) following the excitation of the atoms. It would appear then that the quantum noise associated with SF will be small if $$\tau _p,L/c<\tau _R<\tau _D.$$ (23) We note for later purposes that $$\omega _p^2=\frac{8\pi Nd^2\omega _o}{\mathrm{}}=\frac{1}{\tau _{RAD}}\frac{N_T}{SL}Sc=\frac{4}{\tau _R}\frac{c}{L},$$ (24) where we have used equation (135) of Appendix A for the single-atom radiative lifetime $`\tau _{RAD}`$. This brief summary lends support to the CKK suggestion, but obviously a more quantitative analysis is called for. To this end we now formulate, in the Heisenberg picture, the quantum theory of pulse propagation in an amplifier. ## 3 Formalism for Pulse Propagation We begin with the Hamiltonian for $`N_T`$ two–level atoms (TLAs) interacting with the quantized electromagnetic field via electric dipole transitions: $$\widehat{H}=\frac{1}{2}\mathrm{}\omega _o\underset{j=1}{\overset{N_T}{}}\widehat{\sigma }_{zj}d\underset{j=1}{\overset{N_T}{}}\widehat{\sigma }_{xj}\widehat{E}(z_j)+\underset{k}{}\mathrm{}\omega _k\widehat{a}_k^{}\widehat{a}_k,$$ (25) where $`\omega _o`$ and $`d`$ have the same meaning as before and $`z_j`$ is the $`z`$–coordinate of atom $`j`$. The carets ( ^) are used to denote operators. We consider a one–dimensional model in which the atoms occupy the region from $`z=0`$ to $`z=L`$ and the field is a superposition of plane waves propagating in the $`z`$ direction. The electric field operator is given by $`\widehat{E}(z)=\widehat{E}^{(+)}(z)+\widehat{E}^{()}(z)`$, where $$\widehat{E}^{(+)}(z)=i\underset{k}{}\left(\frac{2\pi \mathrm{}\omega _k}{S\mathrm{}}\right)^{1/2}\widehat{a}_ke^{ikz}(k=\omega _k/c)$$ (26) and $`\widehat{E}^{()}(z)=\widehat{E}^{(+)}(z)^{}`$. $`S\mathrm{}`$, where $`S`$, as before, is a cross–sectional area and $`\mathrm{}`$ a length, is the quantization volume. For simplicity we consider only a single field polarization, namely linear polarization along the direction of the transition dipole moment of the TLAs. $`\widehat{a}_k`$ and $`\widehat{a}_k^{}`$ are the photon annihilation and creation operators, respectively, for mode $`k`$, and the $`\widehat{\sigma }`$’s are the Pauli two–state operators in the standard notation. We will work in the Heisenberg picture, in which the time–dependent electric field operator satisfies $$\left(\frac{^2}{z^2}\frac{1}{c^2}\frac{^2}{t^2}\right)\widehat{E}=\frac{4\pi }{c^2}\frac{^2\widehat{P}}{t^2}=\frac{4\pi d}{c^2S}\underset{j=1}{\overset{N_T}{}}\frac{^2\widehat{\sigma }_{xj}}{t^2}\delta (zz_j)\frac{4\pi }{c^2}Nd\frac{^2}{t^2}\widehat{\sigma }_x(z,t),$$ (27) where in the last step we have made the continuum approximation for the polarization density $`\widehat{P}`$, assuming a uniform atomic density $`N`$. We now write $$\widehat{E}^{(+)}(z,t)=\widehat{F}(z,t)e^{i\omega (tz/c)}$$ (28) and assume $`\widehat{F}(z,t)`$ is slowly varying in $`z`$ and $`t`$ compared with $`\mathrm{exp}[i\omega (tz/c)]`$. In this approximation $$2i\frac{\omega }{c}(\frac{\widehat{F}}{z}+\frac{1}{c}\frac{\widehat{F}}{t})+h.c.=\frac{4\pi }{c^2}Nd\frac{^2\widehat{\sigma }_x}{t^2}e^{i\omega (tz/c)}.$$ (29) It will be convenient to use the atomic lowering and raising operators $`\widehat{\sigma }=\frac{1}{2}(\widehat{\sigma }_xi\widehat{\sigma }_y)`$ and $`\widehat{\sigma }^{}=\frac{1}{2}(\widehat{\sigma }_x+i\widehat{\sigma }_y)`$, respectively, such that $`[\widehat{\sigma },\widehat{\sigma }^{}]=\widehat{\sigma }_z`$, and to write $$\widehat{\sigma }(z,t)=\widehat{s}(z,t)e^{i\omega (tz/c)},$$ (30) where the operator $`\widehat{s}(z,t)`$ is assumed to be slowly varying in the same sense as $`\widehat{F}(z,t)`$. Then, in the rotating–wave approximation (RWA), we can replace (29) with $$\frac{\widehat{F}}{z}+\frac{1}{c}\frac{\widehat{F}}{t}=(2\pi iNd\frac{\omega _o}{c})\widehat{s},$$ (31) where on the right-hand side we have approximated $`\omega `$ by $`\omega _o`$. This equation and the TLA Heisenberg equations $$\frac{\widehat{s}}{t}=i(\mathrm{\Delta }i\beta )\widehat{s}\frac{id}{\mathrm{}}\widehat{\sigma }_z\widehat{F},$$ (32) $$\frac{\widehat{\sigma }_z}{t}=2\beta (1+\widehat{\sigma }_z)\frac{2id}{\mathrm{}}(\widehat{F}^{}\widehat{s}\widehat{s}^{}\widehat{F}),$$ (33) derived in Appendix A form a closed set of operator equations. They provide the basis for a quantum theory of propagation in either amplifying or absorbing media. In the semiclasical approximation in which the atom and field operators are replaced by their expectation values, equations (31)-(33) reduce to well known Maxwell-Bloch equations. Otherwise, different limits can apply: * The limit of $`\beta 0,\mathrm{\Delta }=0`$ and $`\widehat{\sigma }_z1`$ considered below gives equations (35)-(37) implying superfluorescence when the initial state of the field is the vacuum. * The limit of $`\omega \omega _0`$ gives the ARS field equation, as discussed below. * Finally, in Section 4 the CKK case of large detuning, $`\widehat{\sigma }_z1`$, and the initial state of a very short incoming pulse, is studied. If the field central frequency $`\omega `$ is assumed to match exactly the atomic resonance frequency $`\omega _o`$, so that $`\mathrm{\Delta }=0`$, and if we restrict ourselves to times short compared with the single–atom radiative lifetime $`[\tau _{RAD}=(2\beta )^1]`$ and assume that the atoms remain with probability $`1`$ in their excited states over times of interest, we can ignore (33) and replace $`\widehat{\sigma }_z(z,t)`$ by 1 and equation (32) by $$\frac{\widehat{s}}{t}=\frac{id}{\mathrm{}}\widehat{F}.$$ (34) In terms of the independent variables $`\zeta =tz/c`$ and $`\eta =z`$, $$\frac{\widehat{s}}{\zeta }=\frac{id}{\mathrm{}}\widehat{F},$$ (35) $$\frac{\widehat{F}}{\eta }=\left(2\pi iNd\frac{\omega _o}{c}\right)\widehat{s},$$ (36) implying $$\frac{^2\widehat{s}}{\eta \zeta }=\left(\frac{\omega _p^2}{4c}\right)\widehat{s},\frac{^2\widehat{F}}{\eta \zeta }=\left(\frac{\omega _p^2}{4c}\right)\widehat{F}.$$ (37) Equations (35)-(37) have been used in studies of the buildup of superfluorescent radiation . It will be useful for the discussion in Section 4 to briefly rederive here one of the most important results of those studies. Equation (31) has the formal solution $`\widehat{F}(z,t)`$ $`=`$ $`\widehat{F}_o(z,t)+\left(2\pi iNd{\displaystyle \frac{\omega _o}{c}}\right){\displaystyle _0^z}𝑑z^{}\widehat{s}(z^{},t{\displaystyle \frac{zz^{}}{c}})\theta (t{\displaystyle \frac{zz^{}}{c}})`$ (38) $`=`$ $`\widehat{F}_o(z,t)+\left(2\pi iNd{\displaystyle \frac{\omega _o}{c}}\right){\displaystyle _0^z}𝑑z^{}\widehat{s}(zz^{},tz^{}/c)\theta (tz^{}/c),`$ where we have chosen the retarded Green function over the advanced Green function in order to ensure causality. Here $`\theta `$ is the unit step function and $`\widehat{F}_o(z,t)`$ is a solution of the homogeneous equation. We are interested here in the expectation value $`\widehat{F}^{}(L,t)\widehat{F}(L,t)`$ at the end ($`z=L`$) of the medium. For SF the expectation value is taken over the vacuum state of the field, in which case the first term on the right-hand side of (38) does not contribute to normally ordered expectation values. We may therefore ignore this term for practical purposes. Defining $`y=2\sqrt{\zeta \eta }`$ we find from (37) that $`\widehat{s}`$ satisfies the differential equation for $`I_0(y)`$, the modified Bessel function of order zero . The solution of interest for $`\widehat{F}(L,t)`$ is then $$\widehat{F}(L,t)=\left(2\pi iNd\frac{\omega _o}{c}\right)_0^L𝑑z^{}\widehat{s}(Lz^{},0)I_0\left(\omega _p\sqrt{(z^{}/c)(tz^{}/c)}\right)\theta (tz^{}/c).$$ (39) In order to calculate $`\widehat{F}^{}(L,t)\widehat{F}(L,t)`$ we require $`\widehat{s}^{}(z^{},0)\widehat{s}(z,0)`$, which we evaluate in Appendix B. We obtain $$\widehat{F}^{}(L,t)\widehat{F}(L,t)=\left(2\pi d\frac{\omega _o}{c}\right)^2\frac{N}{S}_0^L𝑑x\theta (tx/c)I_0^2\left(\omega _p\sqrt{(x/c)(tx/c)}\right).$$ (40) For times large enough that $`I_0`$ may be replaced by its asymptotic form, $$\widehat{F}^{}(L,t)\widehat{F}(L,t)\frac{1}{8\pi }\frac{2\pi \mathrm{}\omega _o}{Sct}e^{\sqrt{t/\tau _R}}.$$ (41) Equating the intensity expectation value $`(c/2\pi )\widehat{F}^{}(L,t)\widehat{F}(L,t)`$ to the maximum expected SF intensity $`N_T\mathrm{}\omega _o/S\tau _R`$, we arrive at the expression (22) for the time at which the SF pulse reaches its peak intensity. In the short-time limit, on the other hand, $$\widehat{F}^{}(L,t)\widehat{F}(L,t)\left(2\pi d\frac{\omega _o}{c}\right)^2\frac{N}{S}ct,$$ (42) a result we will return to in Section 4. ### Approximation Leading to ARS Field Equation Our considerations thus far assume that the field central frequency lies in the vicinity of the atomic resonance in the sense that the detuning $`\mathrm{\Delta }`$ is small in magnitude compared with $`\omega `$ and $`\omega _o`$. Let us now suppose instead that the field frequency $`\omega `$ is very large compared with $`\omega _o`$. In this case we must work with the atomic operators $`\widehat{\sigma }_x`$, $`\widehat{\sigma }_y`$ instead of the slowly varying $`\widehat{s}`$. ¿From equations (124) and (125) of Appendix A we have $$\ddot{\widehat{\sigma }}_x+\omega _o^2\widehat{\sigma }_x=\frac{2d\omega _o}{\mathrm{}}\widehat{\sigma }_z\widehat{E}\frac{2d\omega _o}{\mathrm{}}\widehat{E}$$ (43) in the approximation $`\widehat{\sigma }_z1`$. The assumption $`\omega >>\omega _o`$ implies $$\ddot{\widehat{\sigma }}_x\frac{2d\omega _o}{\mathrm{}}\widehat{E},$$ (44) so that, from equation (27), $$\left(\frac{^2}{t^2}c^2\frac{^2}{z^2}\omega _p^2\right)\widehat{E}=0.$$ (45) This is identical to the equation of motion for the quantum field in the ARS model when we equate $`\omega _p^2`$ to their $`m^2`$. From this perspective the ARS equation of motion describes the interaction of the electromagnetic field with $`N`$ unbound electrons ($`\omega >>\omega _o`$) per unit volume. However, the usual plasma dispersion formula $`n^2=1\omega _p^2/\omega ^2`$ for the refractive index $`n`$ is replaced in this case by $$n^2=1+\omega _p^2/\omega ^2.$$ (46) This is a consequence of the assumption $`\widehat{\sigma }_z1`$; had we assumed $`\widehat{\sigma }_z1`$ we would have obtained the familiar plasma dispersion formula. To describe the growth of the quantum noise with time in this model, we write (45) in the form $$\frac{^2\widehat{E}}{\tau _1\tau _2}\frac{m^2}{4}\widehat{E}=0,$$ (47) where $`\tau _1=tz/c`$, $`\tau _2=t+z/c`$. In terms of the independent variable $`y=m\sqrt{\tau _1\tau _2}`$, equation(47) has solutions that are linear combinations of the zero-order modified Bessel functions $`I_0(y),K_0(y)`$. For large $`t`$, the vacuum expectation value $$\widehat{E}^2(z,t)I_0^2(y)\frac{e^{2mt}}{2\pi mt},$$ (48) so that the quantum noise grows exponentially in time from the initial fluctuations of the vacuum field, the fluctuations present before the medium in the ARS model is “inverted.” ## 4 Signal and Noise We wish to determine to what extent the observation of the superluminal group velocity considered by CKK will be affected by quantum noise. The system of interest is described by the Heisenberg equations of motion (31) and (32). We approximate $`\widehat{\sigma }_z`$ by 1, assuming that pulse durations $`\tau _p`$ and transit times $`L/c`$ are sufficiently small that de-excitation of the initially inverted atoms by radiation (or any other decay process) is negligible. The situation here is different from that describing the onset of SF in that (a) the detuning $`\mathrm{\Delta }`$ is not zero but is instead large (Section 2), and (b) the initial state of the field is not the vacuum but corresponds to a short pulse of radiation from some external source. The equation for $`\widehat{s}(z,t)`$ in the present model is $$\frac{\widehat{s}}{t}=i(\mathrm{\Delta }i\beta )\widehat{s}\frac{id}{\mathrm{}}\widehat{F},$$ (49) or $$\widehat{s}(z,t)=\widehat{s}(z,t_o)e^{i(\mathrm{\Delta }i\beta )(tt_o)}\frac{id}{\mathrm{}}_{t_o}^t𝑑t^{}\widehat{F}(z,t^{})e^{i(\mathrm{\Delta }i\beta )(t^{}t)}.$$ (50) $`t_o`$ is some initial time, before any pulse is injected into the medium. We take $`\widehat{F}(z,t_o)=0`$, although of course what this really means is that there is no nonvanishing field or intensity in the medium at $`t_o`$, so that for practical purposes (normally ordered expectation values) we can in effect ignore the operator $`\widehat{F}(z,t_o)`$ in the equation for $`\widehat{s}(z,t)`$. The pulse is assumed to have a central frequency $`\omega `$ and to have no significant frequency components near the resonance frequency $`\omega _o`$: $`|\mathrm{\Delta }|\tau _p>1`$. We assume that $`|\mathrm{\Delta }|\tau _p`$ is large enough that we can approximate (50) by integrating by parts and retaining only the leading terms: $$\widehat{s}(z,t)\widehat{s}(z,t_o)e^{i(\mathrm{\Delta }i\beta )(tt_o)}\frac{d}{\mathrm{}}\frac{\mathrm{\Delta }+i\beta }{\mathrm{\Delta }^2+\beta ^2}\widehat{F}(z,t)\frac{id}{\mathrm{}\mathrm{\Delta }^2}\frac{\widehat{F}}{t}.$$ (51) As will be clear from the analysis that follows, this approximation implies the undistorted propagation of the incident pulse at the group velocity $`v_g`$, as assumed by CKK. ¿From (31), $$\frac{\widehat{F}}{z}+\frac{1}{c}\frac{\widehat{F}}{t}\left(2\pi iNd\frac{\omega _o}{c}\right)\widehat{s}(z,t_o)e^{i(\mathrm{\Delta }i\beta )(tt_o)}+\frac{g}{2}\widehat{F}+i[n(\omega )1]\frac{\omega }{c}\widehat{F}+\left(\frac{1}{c}\frac{1}{v_g}\right)\frac{\widehat{F}}{t},$$ (52) where $$g\frac{4\pi Nd^2\omega _o}{\mathrm{}c}\frac{\beta }{\mathrm{\Delta }^2+\beta ^2}$$ (53) is the gain coefficient for propagation of a field with frequency $`\omega `$ in the inverted medium. We have used equation (15) for the refractive index $`n(\omega )`$ and (17) for $`v_g/c1`$. Writing $`\widehat{F}(z,t)=\widehat{F}^{}(z,t)e^{i[n(\omega )1]\omega z/c}`$ and $`\widehat{s}(z,t_o)=\widehat{s}^{}(z,t_o)e^{i[n(\omega )1]\omega z/c}`$ yields an equation in terms of the primed variables in which the term $`i[n(\omega )1](\omega /c)z`$ associated with phase velocity is eliminated. Then, ignoring for practical purposes the difference between the primed and unprimed variables, we have $$\frac{\widehat{F}}{z}+\frac{1}{v_g}\frac{\widehat{F}}{t}=\frac{g}{2}\widehat{F}+\left(2\pi iNd\frac{\omega _o}{c}\right)\widehat{s}(z,t_o)e^{i(\mathrm{\Delta }i\beta )(tt_o)},$$ (54) and therefore $`\widehat{F}(z,t)`$ $`=`$ $`\widehat{F}(0,tz/v_g)e^{gz/2}+\left(2\pi iNd{\displaystyle \frac{\omega _o}{c}}\right){\displaystyle _0^z}𝑑z^{}\widehat{s}(z^{},t_o)e^{g(zz^{})/2}`$ (55) $`\times `$ $`e^{i(\mathrm{\Delta }i\beta )[tt_o(zz^{})/v_g]}\theta (tt_o(zz^{})/v_g)`$ $``$ $`\widehat{F}_s(0,tz/v_g)e^{gz/2}+\widehat{F}_n(z,t),`$ where the subscripts s and n denoted signal and noise, respectively. Here $$\widehat{F}_n(z,t)=\left(2\pi iNd\frac{\omega _o}{c}\right)_0^z𝑑z^{}\widehat{s}(z^{},t_o)e^{g(zz^{})/2}e^{i(\mathrm{\Delta }i\beta )[tt_o(zz^{})/v_g]}\theta (tt_o(zz^{})/v_g)$$ (56) is a quantum noise field associated with the quantum fluctuations of the atomic dipoles. To appreciate the significance of $`g`$ as defined by equation (53), consider the gain coefficient $`g_R`$ for a radiatively broadened transition of frequency $`\omega _o`$ and radiative decay rate $`1/\tau _{RAD}=2\beta `$. For light of frequency $`\omega =\omega _o\mathrm{\Delta }`$, $$g_R=\frac{NS}{\tau _{RAD}}\frac{2\beta }{\mathrm{\Delta }^2+\beta ^2}=\frac{4\pi Nd^2\omega _o}{\mathrm{}c}\frac{\beta }{\mathrm{\Delta }^2+\beta ^2}$$ (57) if we assume that all the $`N`$ atoms per unit volume are in the upper state of the amplifying transition. Thus $`g_R=g`$, i.e., $`g`$ is just the gain coefficient for amplification by stimulated emission. We note also that, from equation (17), $$g=2\beta \left(\frac{1}{c}\frac{1}{v_g}\right)$$ (58) in the case under consideration where the amplifying transition is radiatively broadened and the detuning is large compared with the gain bandwidth. The operator $`\widehat{s}(z,t_o)`$ has the expectation-value properties (144) and (145) of Appendix B. These properties imply $`\widehat{F}_n(z,t)=\widehat{F}_n^{}(z,t)=0`$ and $`\widehat{F}_n^{}(z,t)\widehat{F}_n(z,t)`$ $`=`$ $`\left(2\pi Nd{\displaystyle \frac{\omega _o}{c}}\right)^2{\displaystyle \frac{L}{N_T}}e^{2\beta (tt_o)}{\displaystyle _{zv_g(tt_o)}^z}𝑑z^{}e^{g(zz^{})}e^{2\beta (zz^{})/v_g}`$ (59) $`=`$ $`\left(2\pi d{\displaystyle \frac{\omega _o}{c}}\right)^2{\displaystyle \frac{N}{S}}{\displaystyle \frac{c}{2\beta }}\left[e^{gv_gt}e^{2\beta t}\right],`$ where we have used the relations (58) and $`N_T=NSL`$ and, to simplify the notation, we have taken $`t_o=0`$. Since the atom and field are initially uncorrelated, i.e., $$\widehat{F}^{}(0,tz/v_g)\widehat{s}_j(t_o)=\widehat{F}_n^{}(0,tz/v_g)\widehat{s}_j(t_o)=0,$$ (60) we have, at the end of the amplifier, $$\widehat{F}^{}(L,t)\widehat{F}(L,t)=\widehat{F}_s^{}(0,tL/v_g)\widehat{F}_s(0,tL/v_g)e^{gL}+\widehat{F}_n^{}(L,t)\widehat{F}_n(L,t)$$ (61) and the signal–to–noise ratio $`SNR(L,t)`$ $``$ $`{\displaystyle \frac{\widehat{F}_s^{}(0,tL/v_g)\widehat{F}_s(0,tL/v_g)e^{gL}}{\widehat{F}_n^{}(L,t)\widehat{F}_n(L,t)}}`$ (62) $`=`$ $`{\displaystyle \frac{\widehat{F}_s^{}(0,tL/v_g)\widehat{F}_s(0,tL/v_g)e^{gL}}{(2\pi d\omega _o/c)^2(N/S)(c/2\beta )\left[e^{gL}e^{2\beta L/v_g}\right]}}`$ $`=`$ $`{\displaystyle \frac{\widehat{F}_s^{}(0,tL/v_g)\widehat{F}_s(0,tL/v_g)e^{2\beta (1/c1/v_g)L}}{(2\pi d\omega _o/c)^2(N/S)(c/2\beta )\left[e^{2\beta (1/c1/v_g)L}e^{2\beta L/v_g}\right]}}`$ $``$ $`{\displaystyle \frac{\widehat{F}_s^{}(0,tL/v_g)\widehat{F}_s(0,tL/v_g)}{(2\pi d\omega _o/c)^2NL/S}}.`$ In the denominators we have taken $`t=L/v_g`$ for the time over which the atoms radiate, and have used the fact that $`2\beta (1/c1/v_g)L=gL`$, the difference of two numbers that themselves are small according to our assumption that propagation times are small compared with the single–atom radiative decay rate, is $`<<1`$. The numerator in equation (62) can be related to the expectation value $`q`$ of the number of photons in the incident signal pulse as follows. The expectation value of the incident signal intensity is $$I_s(0,t)=\frac{v_g}{2\pi }\widehat{F}_s(0,t)\widehat{F}_s(0,t)=I_oe^{t^2/\tau _p^2}$$ (63) for a Gaussian pulse of duration $`\tau _p`$. Requiring that the energy flux $`_{\mathrm{}}^{\mathrm{}}𝑑tI_s(z,t)`$ be $`q\mathrm{}\omega /Sq\mathrm{}\omega _o/S`$ implies $`I_o=q\mathrm{}\omega _o/(S\tau _p\sqrt{\pi })`$ and therefore $$\widehat{F}_s^{}(0,tL/v_g)\widehat{F}_s(0,tL/v_g)=q\frac{2\pi \mathrm{}\omega _o}{v_gS\tau _p\sqrt{\pi }}e^{(tL/v_g)^2/\tau _p^2}.$$ (64) Thus $`SNR(L,t)`$ $`=`$ $`{\displaystyle \frac{q}{\tau _p\sqrt{\pi }}}{\displaystyle \frac{c}{v_g}}e^{(tL/v_g)^2/\tau _p^2}\left({\displaystyle \frac{2\pi d^2\omega _o}{\mathrm{}cS}}NSL\right)^1`$ (65) $`=`$ $`{\displaystyle \frac{q}{\sqrt{\pi }}}\left({\displaystyle \frac{4c}{\omega _p^2L\tau _p}}\right){\displaystyle \frac{c}{v_g}}e^{(tL/v_g)^2/\tau _p^2}`$ $`=`$ $`{\displaystyle \frac{q}{\sqrt{\pi }}}{\displaystyle \frac{\tau _R}{\tau _p}}{\displaystyle \frac{c}{v_g}}e^{(tL/v_g)^2/\tau _p^2},`$ where we have used equation (24) . Among the criteria given by CKK for the observation of a superluminal pulse is that “The probe-pulse duration \[$`\tau _p`$\] must not exceed $`\tau _R=4c/L\omega _p^2`$.” This criterion implies, from equation (65), that $`SNR(L,t)(q/\sqrt{\pi })c/v_g`$ and therefore that it is possible, even for $`q1`$, to have superluminal propagation with $`SNR(L,t)>1`$ if the pulse duration is short enough: $`\tau _p<\tau _Rc/v_g`$. In order to relate this conclusion to ARS, we use equation (17) to write (65) as $$SNR(L,t)=\frac{q}{\sqrt{\pi }}\frac{\tau _p}{\left(\frac{v_g}{c}1\right)\frac{L}{c}\mathrm{\Delta }^2\tau _p^2}e^{(tL/v_g)^2/\tau _p^2}.$$ (66) We see from this expression that, if we impose the ARS condition (2), i.e., $`(v_g/c1)L/c>>\tau _p`$, then $$SNR(L,t)<<\frac{q}{\sqrt{\pi }}\frac{1}{(\mathrm{\Delta }\tau _p)^2}e^{(tL/v_g)^2/\tau _p^2},$$ (67) so that, given also the condition on $`|\mathrm{\Delta }|\tau _p`$ discussed before equation (51), the signal-to-noise ratio will be very small when the ARS condition for strong distinguishability of superluminal propagation from propagation at the speed $`c`$ is satisfied. In fact if $`(v_g/c1)L/c>>\tau _p`$ and therefore $`SNR(L,t)`$ is very small for $`q1`$, then $$t/\tau _R=\frac{L/c}{\tau _R}=\frac{v_g}{c}\frac{L/v_g}{\tau _R}\stackrel{>}{}\frac{v_g}{c}\frac{\tau _p}{\tau _R},$$ (68) which, from (65), must be large. Then the SF noise must be exponentially large \[equation (41)\]. It follows that $`q`$ must be exponentially large in order to maintain a signal-to-noise ratio greater than unity. This is consistent with the ARS conclusion that “for the signal amplitude to be larger than the amplitude of the fluctuations at the observation time, the signal amplitude should be exponentially large” . Our results are therefore in agreement with ARS in that, if we require the separation of the superluminal pulse and a twin vacuum-propagated pulse to be much larger than the pulse duration, the signal-to-noise ratio will be very small at the one- or few-photon level. On the other hand, the results are not inconsistent with CKK: even at the one-photon level we can achieve a signal-to-noise ratio greater than unity if this separation ($`[v_g/c1]L/c`$) is smaller than the pulse duration $`\tau _p`$ \[Equation (66)\]. ### Physical Origin of the Noise Limiting the Observation of <br>Superluminal Group Velocity Note that, when we set the time $`t`$ in equation (42) for the short-time SF noise intensity equal to the “observation time” $`L/c`$, we obtain exactly the noise intensity appearing in the denominator of equation (62) . Thus the quantum noise that imposes limitations on the observation of superluminal group velocity is attributable to the initiation of SF. ### Operator Ordering and Relation to ARS Approach Less obvious, perhaps, is the relation between the quantum noise we have considered – which stems from the atomic dipole fluctuations characterized by equations (144) and (145) of Appendix B – and the quantum noise of ARS, which is attributed to the quantum fluctuations of the field. To establish the relation to the ARS approach we return to our calculation of the noise intensity, using now anti-normally ordered field operators instead of the normally ordered operators used before. Thus we consider now the expectation value $`\widehat{F}(z,t)\widehat{F}^{}(z,t)`$ instead of $`\widehat{F}^{}(z,t)\widehat{F}(z,t)`$. In this approach the atomic dipole fluctuations play no explicit role, as can be seen from equation (55) and the fact that $$\widehat{s}(z^{},t_o)\widehat{s}^{}(z^{\prime \prime },t_o)=0$$ (69) for excited atoms. In this case, however, the initially unoccupied modes of the field make a nonvanishing contribution as a consequence of non-normal ordering: $`\widehat{F}(0,tL/v_g)\widehat{F}^{}(0,tL/v_g)`$ $`=`$ $`{\displaystyle \underset{k}{}}{\displaystyle \frac{2\pi \mathrm{}\omega _k}{S\mathrm{}}}\widehat{a}_k(0)\widehat{a}_k^{}(0)e^{g(\omega _k)L}`$ (70) $``$ $`{\displaystyle \underset{k}{}}{\displaystyle \frac{2\pi \mathrm{}\omega _k}{S\mathrm{}}}[g(\omega _k)L+1],`$ which follows from (26) and (28) and the approximation $`gL<<1`$ upon which (65) is based. The contribution from the term that does not vanish as $`L0`$ can be ignored, as it corresponds to vacuum quantum noise (energy $`\frac{1}{2}\mathrm{}\omega _k`$ per mode) that is present even in the absence of the amplifier. In other words, the quantum noise of the field in the presence of the amplifier is $`\widehat{F}(0,tz/v_g)\widehat{F}^{}(0,tz/v_g)_n{\displaystyle \underset{k}{}}{\displaystyle \frac{2\pi \mathrm{}\omega _k}{S\mathrm{}}}g(\omega _k)L{\displaystyle \frac{\mathrm{}}{2\pi c}}{\displaystyle 𝑑\omega \frac{2\pi \mathrm{}\omega }{S\mathrm{}}g(\omega )L}`$ $`\pi \left({\displaystyle \frac{2\omega _od}{c}}\right)^2{\displaystyle \frac{NL}{c}}{\displaystyle _0^{\mathrm{}}}𝑑\omega {\displaystyle \frac{\beta }{\mathrm{\Delta }^2+\beta ^2}},`$ (71) where we have gone to the mode continuum limit, approximated $`\omega `$ by $`\omega _o`$ in the numerator of the integrand, and used equation (53) for the gain coefficient. Performing the integration, we obtain exactly the noise term appearing in the denominator in the last line of (62). But now the noise is attributable to the amplification of vacuum field fluctuations . Thus we can attribute the quantum noise that limits the observation of superluminal group velocity to either the quantum fluctuations of the field in the inverted medium, as do ARS, or to the quantum fluctuations of the inverted atoms, as in our derivation of the signal-to-noise ratio. The situation here is similar to that in the theory of the initiation of SF, as discussed by Polder et al. , or, as noted by those authors, to the theory of spontaneous emission by a single atom . ### Limit of Very Small Transition Frequency Since the origin of noise in the optical amplifier is associated ultimately with spontaneous emission, the question arises as to whether the signal-to-noise ratio might be increased by employing a transition having a very small transition frequency $`\omega _o`$ and therefore a very large radiative lifetime. Indeed, since $`\omega _p^2\omega _o`$, the second line of equation (65) suggests at first glance that $`SNR\mathrm{}`$ in the limit $`\omega _o0`$. However, equation (16) shows that $`v_gc`$ in this limit: the superluminal effect itself becomes weaker as the spontaneous emission rate is made smaller. In this connection we invoke once again the form (66) of the signal-to-noise ratio. If we assume $`|\mathrm{\Delta }|\tau _p>1`$ in order that the pulse does not undergo substantial distortion as a consequence of strong absorption, then $$SNR(L,t)<q\frac{c\tau _p}{(v_gc)L/c}.$$ (72) In other words, the signal-to-noise ratio must be smaller than the number of photons in the incident pulse times a factor equal to to length of the vacuum-propagated pulse divided by the separation of the vacuum-propagated pulse and the pulse emerging from the amplifier, independent of the the atomic transition frequency or the radiative lifetime. At the one- or few-photon level the signal-to-noise ratio must therefore be less than unity under the ARS criteria for the observation of superluminal group velocity, regardless of the frequency or strength of the amplifying transition. ## 5 Unitarity and Superluminal Propagation We now turn our attention from the specific example of the optical amplifier to some general features of superluminal propagation that follow generally from the unitary evolution of the state vector, considered here within first quantization. The time evolution of a wave packet can be formulated in terms of a unitary operator $`U(t)`$ or equivalently in terms of a coordinate-space propagator $`G(xx^{},t)=x|U(t)|x^{}`$: $`|\mathrm{\Psi }(t)`$ $`=`$ $`U(t)|\mathrm{\Psi }(0)`$ $`\mathrm{\Psi }(x,t)`$ $`=`$ $`x|\mathrm{\Psi }(t)={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x^{}G(xx^{},t)\mathrm{\Psi }(x^{},0).`$ (73) The assumption that the propagator vanishes identically outside the light cone implies that $`G(xx^{}>ct,t)=0.`$ (74) Given an initial wave packet centered around $`x=X_0<0`$ at $`t=0`$, we assume that at a later time $`t>0`$ it will be centered around $`X_0+v_gt`$, as in the example of pulse propagation in an inverted medium. We divide the wave packet into two parts, which we label as “superluminal” (S) and “luminal” (L), in the following way: $`\mathrm{\Psi }(x,T)\{\begin{array}{cc}\mathrm{\Psi }^S(x,T)\hfill & x>cT,\hfill \\ \mathrm{\Psi }^L(x,T)\hfill & x<cT.\hfill \end{array}`$ (77) $`\mathrm{\Psi }^S`$ vanishes if the group velocity $`v_g<c`$. Suppose that $`v_g>c`$ and that we let the wave packet propagate for a time $`T`$ long enough that a superluminal signal can be clearly identified. That is, we assume that at $`t=T`$, $$\mathrm{\Psi }^L(x,T)0.$$ (78) Now $$\mathrm{\Psi }(T)|\mathrm{\Psi }(T)=\mathrm{\Psi }(0)|\mathrm{\Psi }(0)$$ (79) due to unitarity, and thus $$_{\mathrm{}}^{\mathrm{}}𝑑x^{}|\mathrm{\Psi }(x^{},0)|^2=_{\mathrm{}}^{\mathrm{}}𝑑x^{}|\mathrm{\Psi }(x^{},T)|^2_{cT}^{\mathrm{}}𝑑x^{}|\mathrm{\Psi }^S(x^{},T)|^2.$$ (80) Physically, this means that the superluminal signal, $`\mathrm{\Psi }^S(x,T)`$, is about as large, or contains about “as many photons,” as the initial wave packet. We now combine the two underlying premises of causality and superluminal propagation as they are defined by equations (74) and (78). Using equation (73) for $`x>cT`$, we write $$\mathrm{\Psi }^S(x,T)=_{\mathrm{}}^0𝑑x^{}G(xx^{},T)\mathrm{\Psi }(x^{},0)+_0^{\mathrm{}}𝑑x^{}G(xx^{},T)\mathrm{\Psi }(x^{},0).$$ (81) The first term vanishes because, according to Eq. (74), the integrand differs from zero only if $`x^{}>xcT>0`$. Thus $$\mathrm{\Psi }^S(x,T)=_{\mathrm{}}^{\mathrm{}}𝑑x^{}G(xx^{},T)\left[\mathrm{\Theta }(x^{})\mathrm{\Psi }(x^{},0)\right].$$ (82) This formulates the notion, which is essential to the ARS argument, that for a causal, \[i.e., equation (74)\], superluminal signal \[equation (78)\], the wave packet is reconstructed from its tail \[equation (82)\]. This rather remarkable reconstruction of the signal propagated without distortion and with superluminal group velocity is especially evident in the temporal domain . (See Figure 1.) The construction (82) of the superluminal wave packet from the tail of the initial wave packet motivated ARS to define another, truncated initial wave packet: $$\mathrm{\Phi }(x,0)\mathrm{\Theta }(x)\mathrm{\Psi }(x,0).$$ (83) The two different initial wave functions, $`\mathrm{\Phi }(x,0)`$ and $`\mathrm{\Psi }(x,0)`$, give the same superluminal signal: $$\mathrm{\Psi }(x>cT,T)=\mathrm{\Phi }(x>cT,T)=_{\mathrm{}}^{\mathrm{}}𝑑x^{}G(xx^{},T)\mathrm{\Phi }(x^{},0).$$ (84) Equation (84) implies what ARS call amplification: a “small” signal propagates to become a “large” signal. After all, $`\mathrm{\Phi }(x,0)`$ is “made from a small number of photons,” while we have just seen that $`\mathrm{\Psi }(x>cT,T)`$ has about the same number of photons as the non-truncated initial wave packet. We note that amplification in this sense is a necessary consequence of a superluminal group velocity. One might be tempted to write (84) symbolically as $$|\mathrm{\Phi }(0)|\mathrm{\Psi }^S(T),$$ (85) where $``$ denotes time evolution under $`U(T)`$. This would be incorrect: the truncated initial wave packet $`\mathrm{\Phi }(x,0)`$ is a perfectly well defined initial state, but it does not evolve into $`\mathrm{\Psi }^S(x,T)`$; part of it evolves luminally. It will be prove convenient to introduce “superluminal” and “luminal” parts of the truncated wave packet in a manner similar to the decomposition (77) used for the complete wave packet $`\mathrm{\Psi }(x,T)`$: $`\mathrm{\Phi }(x,T)\{\begin{array}{cc}\mathrm{\Phi }^S(x,T)\hfill & x>cT,\hfill \\ \mathrm{\Phi }^L(x,T)\hfill & x<cT.\hfill \end{array}`$ (88) We note that, while the superluminal part of the time-evolved truncated initial state is the same as the superluminal part of the time-evolved non-truncated initial state, the luminal parts of these signals differ: $`\mathrm{\Phi }^S(x,T)=\mathrm{\Psi }^S(x,T),`$ (89) $`\mathrm{\Phi }^L(x,T)\mathrm{\Psi }^L(x,T)0.`$ (90) That is, while the luminal part of the time-evolved complete wave packet approximately vanishes ($`\mathrm{\Psi }^L(x,T)0`$), the luminal part of the truncated wavepacket, $`\mathrm{\Phi }^L(x,T)`$, does not. We show below that, on the contrary, it grows exponentially with time. ### Momentum space: normal and unstable modes We are comparing the time evolution of two different initial wavepackets, $`\mathrm{\Psi }(x,0)`$ and $`\mathrm{\Phi }(x,0)`$ where $`\mathrm{\Phi }(x,0)=\mathrm{\Theta }(x)\mathrm{\Psi }(x,0)`$. It is useful to define still another initial wave packet, $$R(x,0)=\mathrm{\Theta }(x)\mathrm{\Psi }(x,0).$$ (91) Clearly, $$\mathrm{\Psi }(x,0)=R(x,0)+\mathrm{\Phi }(x,0).$$ (92) After a time $`T`$, $`\mathrm{\Psi }(x,0)`$ evolves into $`\mathrm{\Psi }(x,T)`$, $`\mathrm{\Phi }(x,0)`$ into $`\mathrm{\Phi }(x,T)`$, and $`R(x,0)`$ into $`R(x,T)`$. The time evolution is linear and $$\mathrm{\Psi }(x,T)=R(x,T)+\mathrm{\Phi }(x,T).$$ (93) Fourier transforming into momentum space, we define $`g(k)`$, $`\zeta (k)`$, and $`\xi (k)`$, by: $`\mathrm{\Psi }(x,t){\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑kg(k)\mathrm{exp}[i(kx\omega _kt)],`$ (94) $`\mathrm{\Phi }(x,t){\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑k\zeta (k)\mathrm{exp}[i(kx\omega _kt)],`$ (95) $`R(x,t){\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑k\xi (k)\mathrm{exp}[i(kx\omega _kt)].`$ (96) ¿From these definitions it is straightforward to show that $`\zeta (k)={\displaystyle \frac{i}{2\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑k^{}{\displaystyle \frac{g(k^{})}{kk^{}i\eta }},`$ (97) $`\xi (k)={\displaystyle \frac{+i}{2\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑k^{}{\displaystyle \frac{g(k^{})}{kk^{}+i\eta }},`$ (98) where $`\eta `$ is an infinitesimal positive number. ¿From the identity $$\frac{1}{kk^{}i\eta }+\frac{1}{kk^{}+i\eta }=2\pi i\delta (kk^{})$$ (99) it follows that $$g(k)=\zeta (k)+\xi (k).$$ (100) Equations (94)-(98) can be written as well in the following way: $`\mathrm{\Psi }(x,t){\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑kg(k)\psi _k(x,t),`$ (101) $`\mathrm{\Phi }(x,t){\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑kg(k)\varphi _k(x,t),`$ (102) $`R(x,t){\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑kg(k)\rho _k(x,t),`$ (103) where $`\psi _k(x,t)=\mathrm{exp}[i(kx\omega _kt)],`$ (104) $`\varphi _k(x,t)={\displaystyle \frac{i}{2\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\kappa {\displaystyle \frac{\mathrm{exp}[i(\kappa x\omega _\kappa t)]}{\kappa ki\eta }},`$ (105) $`\rho _k(x,t)={\displaystyle \frac{+i}{2\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\kappa {\displaystyle \frac{\mathrm{exp}[i(\kappa x\omega _\kappa t)]}{\kappa k+i\eta }},`$ (106) $`\psi _k(x,t)=\varphi _k(x,t)+\rho _k(x,t).`$ (107) For $`t=0`$ we obtain, as required by their definitions, $`\varphi _k(x,0)=\mathrm{\Theta }(x)\psi _k(x,0),`$ (108) $`\rho _k(x,0)=\mathrm{\Theta }(x)\psi _k(x,0).`$ (109) We now invoke the premises of causality and superluminal propagation, focusing on the ARS model involving the dispersion relation $$\omega _k=c\sqrt{k^2m^2}.$$ (110) As long as $`|k|>m`$ this dispersion relation describes normal oscillating modes. Unstable modes exist for $`|k|<m`$. One might attempt to avoid the unstable modes altogether by choosing a $`g(k)`$ that vanishes or is negligibly small for $`|k|<m`$. This can be done, for example, by choosing an initial state with a Gaussian $`g(k)`$, centered around $`k_o`$ and having a width $`\mathrm{\Delta }k_o`$ such that $`|k_o\pm \mathrm{\Delta }k_o|>>m`$. This corresponds in the case of the optical amplifier to a pulse detuning large compared with a radiative decay rate. It turns out, however, as might be expected from the example of the optical amplifier, that even for such an initial wave packet $`\mathrm{\Psi }(x,0)`$ the unstable modes play an essential role in the time evolution of both the truncated and the residual wavepackets, $`\mathrm{\Phi }(x,t)`$ and $`R(x,t)`$, respectively. Consider the integrals in (105) and (106) as contour integrals in the complex $`\kappa `$ plane. The integrands, analytically continued into the complex $`\kappa `$ plane, each have a single, simple pole above or below the real $`\kappa `$ axis at $`\kappa =k\pm i\eta `$, and both have two branch points at $`\kappa =\pm m`$, which we connect with a branch cut on the line segment $`(m,m)`$ on the real $`\kappa `$ axis. The contour from $`\mathrm{}`$ to $`\mathrm{}`$ should pass, as usual, slightly above the real $`\kappa `$ axis (at a distance smaller than $`\eta `$). As shown below, this ensures causality according to equation (74). In the limit of infinite $`|\kappa |`$, $$\underset{|\kappa |\mathrm{}}{lim}\omega _\kappa =c\kappa ,$$ (111) and on the circle at infinity, $$\kappa x\omega _\kappa t\kappa (xct).$$ (112) For $`x>ct`$ we can therefore close the contour integral in the upper-half plane, whereas for $`x<ct`$ we close the contour in the lower half. In both cases the contributions to the integral from the arcs at infinity vanish. Using first the residue theorem for $`x>ct`$, we see immediately that the superluminal parts of the time-evolved residual and truncated wave packets satisfy $`\rho _k(x>ct,t)=0,`$ (113) $`\varphi _k(x>ct,t)=\mathrm{exp}[i(kx\omega _kt)].`$ (114) These results are not surprising, as they simply reformulate equations (74) and (84), respectively. For $`x<ct`$, where we close the contour in the lower half-plane, the integral encircles the branch cut on $`(m,m)`$. After deforming the contour and isolating contributions from this branch cut, we use the residue theorem and obtain $`\rho _k(x<ct,t)=\mathrm{exp}[i(kx\omega _kt)]I_k^\rho (x,t),`$ (115) $`\varphi _k(x<ct,t)=I_k^\varphi (x,t),`$ (116) where $`I_k^\rho (x,t)={\displaystyle \frac{i}{2\pi }}{\displaystyle _C}𝑑\kappa {\displaystyle \frac{\mathrm{exp}[i(\kappa xct\sqrt{\kappa ^2m^2})]}{\kappa ki\eta }},`$ (117) $`I_k^\varphi (x,t)={\displaystyle \frac{i}{2\pi }}{\displaystyle _C}𝑑\kappa {\displaystyle \frac{\mathrm{exp}[i(\kappa xct\sqrt{\kappa ^2m^2})]}{\kappa k+i\eta }},`$ (118) and $`_C𝑑\kappa `$ is a closed contour circling counter-clockwise the branch cut on the line segment $`(m,m)`$ while not circling the poles at $`k\pm i\eta `$. Each of the integrals, $`I_k^\rho (x,t)`$ and $`I_k^\varphi (x,t)`$, is dominated by a saddle point on the imaginary $`\kappa `$ axis in the complex $`\kappa `$ plane and exponentially grows with time. Combining terms, we obtain $$R(x,t)=\mathrm{\Theta }(ctx)\mathrm{\Psi }(x,t)\mathrm{\Theta }(ctx)_{\mathrm{}}^{\mathrm{}}𝑑kg(k)I_k^\rho (x,t),$$ (119) $$\mathrm{\Phi }(x,t)=\mathrm{\Theta }(xct)\mathrm{\Psi }(x,t)+\mathrm{\Theta }(ctx)_{\mathrm{}}^{\mathrm{}}𝑑kg(k)I_k^\varphi (x,t).$$ (120) The integrals give exponentially growing contributions to the luminal parts of both the truncated and residual wavepackets. Our choice of $`g(k)`$ enforces $`|k\pm i\eta |>m`$, and as a result, $$_{\mathrm{}}^{\mathrm{}}𝑑kg(k)\left[I_k^\varphi (x,t)I_k^\rho (x,t)\right]=0$$ (121) We see therefore that, when the residual and truncated wavepackets (119) and (120) are combined to form the complete wavepacket $`\mathrm{\Psi }(x,t)`$ \[Eq. (93)\], the exponentially growing luminal parts cancel each other. ### Discussion and implications for quantum noise We are studying the time evolution of three wavepackets: the complete wavepacket $`\mathrm{\Psi }(x,t)`$, the truncated wavepacket $`\mathrm{\Phi }(x,t)`$, and the retarded, residual wavepacket $`R(x,t)`$. These three wavepackets can be decomposed in two different ways. In equations (94)-(98) they were decomposed in the usual way via a Fourier transform at the initial time $`t=0`$ into normal and unstable modes. The Fourier components of the truncated wavepacket $`\zeta (k)`$ and the retarded wavepacket $`\xi (k)`$ are related to the Fourier components of the complete wave packet, $`g(k)`$, by equations (97) and (98), respectively. If we choose to construct the complete wave packet from normal modes $`g(k)`$, where $`|k|m`$, the truncated and retarded wavepackets will have a strong unstable-mode component in them. This was discussed by ARS, who pointed out that because of the unitarity of the time evolution, the unstable modes are accompanied by an enhancement of the quantum noise. In order to identify the noise in a space-time picture we employed in equations (101)-(106) a less common decomposition. The difference between (101)-(106) and (94)-(98) lies in the order of integration. Both decompositions can be derived from $`\mathrm{\Psi }(x,t){\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑kg(k)\mathrm{exp}[i(kx\omega _kt)],`$ $`\mathrm{\Phi }(x,t){\displaystyle \frac{i}{2\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑q{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑p{\displaystyle \frac{g(p)\mathrm{exp}[i(qx\omega _qt)]}{qpi\eta }},`$ (122) $`R(x,t){\displaystyle \frac{i}{2\pi }}{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑q{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑p{\displaystyle \frac{g(p)\mathrm{exp}[i(qx\omega _qt)]}{qp+i\eta }}.`$ (123) Equation (101) describes a wavepacket made of a superposition of oscillating waves $`\psi _k(x,t)\mathrm{exp}[i(kx\omega _kt)]`$, with the momentum distribution $`g(k)`$. In (102) and (103) each of these oscillating waves is replaced by a new wavefunction, $`\varphi _k(x,t)`$ and $`\rho _k(x,t)`$, respectively. The weight function for the superposition forming the respective wavepackets remains $`g(k)`$, but $`k`$ has lost its meaning as a physical momentum. At any time, $`\psi _k=\varphi _k+\rho _k`$. At $`t=0`$, $`\varphi _k`$ and $`\rho _k`$ are obtained from $`\psi _k`$ by truncation. At a later time $`t>0`$, one can distinguish between two regions. In the superluminal region where $`x>ct`$, $`\varphi _k=\psi _k`$ and $`\rho _k=0`$. In the luminal region where $`x<ct`$, $`\varphi _k\psi _k`$: While $`\psi _k`$ is everywhere a periodic wavefunction oscillating in space and time, $`\varphi _k`$ is in this region exponentially growing as a function of both $`t`$ and $`x`$; it is not oscillating in this region. In the same retarded region $`\rho _k`$ has a periodic oscillating component equal to $`\psi _k`$ and an exponentially growing component which exactly cancels the contribution of $`\varphi _k`$ to this region. The three wavepackets we consider are formed by superpositions of these different wavefunctions with the same weight function $`g(k)`$. They evolve in time in the following way. In the superluminal region $`x>ct`$ the oscillating wave functions $`\psi _k=\varphi _k\mathrm{exp}[i(kx\omega _kt)]`$, with $`\omega _k`$ given above, combine to form a wavepacket moving at the group velocity $`v_g>c`$; this is the superluminal signal. In the luminal region $`x<ct`$ the oscillating wave functions combine to cancel each other. This cancellation ensures the unitary time evolution of the complete wavepacket. The residual part of the complete wavepacket is essential for this cancellation to occur. Using the language of truncated wavepackets introduced by ARS, we see that the superluminal signal is constructed completely from the time evolution of the forward tail, i.e., from the time evolution of the truncated wavepacket. This truncated wavepacket evolves with time into a combination of the superluminal signal and an additional, exponentially increasing part in the luminal region $`x<ct`$. As discussed below, this additional part that grows exponentially with time can be expected to be accompanied by substantial quantum noise, as ARS observed using a different decomposition of the same truncated wavepacket. The new decomposition presented here therefore leads us to conclude that the exponentially growing noise is mostly “luminal” and will be delayed compared with the superluminal signal. This conclusion is consistent with the exponentially growing noise due to SF in the case of the optical amplifier . Looking at the complete wavepacket, we observe that contributions from the time-evolved residual wavepacket will cancel in the luminal region $`x<ct`$ the contributions from the time-evolved truncated wavepacket. However, while the signal in this region vanishes by the cancellation of the two exponentially growing contributions, the noise does not cancel – and may be very large . We note that an amplification of the signal in the superluminal region does occur, but our new decomposition indicates that this amplification is mostly a result of a rather efficient constructive interference of oscillating wavefunctions, while the luminal parts of the time-evolved truncated and retarded wavepackets appear to be controlled by the unstable modes. Our analysis in this section, being based on a “first-quantization” approach in which the wave packets are c-numbers, not operators, has not dealt explicity with quantum noise. However, as in the theory of the initiation of superfluorescence , the linearity of the model resulting from the approximation that there is no change in the atomic inversion over the time scales of interest allows a treatment of the operator fields as classical, fluctuating c-number fields . Thus the shaded part of Figure 1a, the “tail” from which the superluminal signal evolves, becomes in such a treatment the truncated signal we have considered plus a fluctuating noise field. In the limit of a very weak incident signal pulse, the superluminal signal will be dominated by the noise part rather than the signal part of the tail shown in Figure 1a, and the signal-to-noise ratio will therefore be small, consistent with the ARS results as well as the results obtained in Section 4 for the model of an optical amplifier. ## 6 Summary We have considered the effects of quantum noise on the propagation of a pulse with superluminal group velocity. In the case considered by CKK , where an off-resonant, short pulse of duration $`\tau _p`$ propagates with superluminal group velocity $`v_g`$ in an optical amplifier, we calculated a signal-to-noise ratio $`SNR`$ and found that, for an incident pulse consisting of a single photon, $`SNR<<1`$ under the condition $`(v_g/c1)L>>\tau _p`$ assumed by ARS for discrimination between the pulse propagating in the amplifier and a twin pulse propagating the same distance in vacuum. This result is fully consistent with the conclusions of ARS based on general considerations and, in particular, the reconstruction of the superluminal pulse from a truncated portion of the initial wave packet. However, if we impose the weaker condition that $`(v_g/c1)L\stackrel{>}{}\tau _p`$, then our conclusion is that $`SNR>1`$ is possible. However, in this case superluminal group velocity is observable in the arrival statistics of many photons, not per shot. We showed that, in the case of the optical amplifier, the quantum noise is attributable to the onset of superfluorescence, and could be associated either with the quantum fluctuations of the field, along the lines of the ARS considerations, or with the quantum fluctuations of the atomic dipoles. We then presented some general considerations based on unitarity and causality and introduced a new wave packet decomposition. In particular, we considered the “residual” wave packet in addition to the complete and truncated wave packets considered by ARS. This led to the conclusion that the noise is mostly luminal, that in the luminal region the truncated and residual signal grow exponentially but cancel each other as required by unitarity, but that the noise is not cancelled. For the case where the propagation time is large enough for the superluminal signal to be clearly distinguished from a twin pulse propagated at the vacuum speed of light, our conclusions were again consistent with ARS. ## Acknowledgements We thank Y. Aharonov, E.L. Bolda, I.H. Deutsch, R.J. Glauber, P.G. Kwiat, B. Reznik, and A.M. Steinberg for helpful discussions or remarks during this work. This work was partially supported by the National Science Foundation through a grant for the Institute for Theoretical Atomic and Molecular Physics (ITAMP) at the Harvard-Smithsonian Center for Astrophysics. ## Appendix A The Heisenberg equations of motion for the Pauli operators follow from the Hamiltonian (25) and the commutation relations $`[\widehat{\sigma }_x,\widehat{\sigma }_y]=2i\widehat{\sigma }_z`$, etc. : $$\dot{\widehat{\sigma }}_{xj}=\omega _o\widehat{\sigma }_{yj},$$ (124) $$\dot{\widehat{\sigma }}_{yj}=\omega _o\widehat{\sigma }_{xj}+\frac{2d}{\mathrm{}}\widehat{\sigma }_{zj}\widehat{E}(z_j,t),$$ (125) $$\dot{\widehat{\sigma }}_{zj}=\frac{2d}{\mathrm{}}\widehat{\sigma }_{yj}\widehat{E}(z_j,t),$$ (126) or, in the rotating–wave approximation, $$\dot{\widehat{\sigma }}_j=i\omega _o\widehat{\sigma }_j\frac{id}{\mathrm{}}\widehat{\sigma }_{zj}\widehat{E}^{(+)}(z_j,t),$$ (127) $$\dot{\widehat{\sigma }}_{zj}=\frac{2id}{\mathrm{}}[\widehat{E}^{()}(z_j,t)\widehat{\sigma }_j\widehat{\sigma }_j^{}E^{(+)}(z_j,t)].$$ (128) ¿From the formal solution of the Heisenberg equation of motion for $`\widehat{a}_k(t)`$ we obtain, using equation (26), $`\widehat{E}^{(+)}(z_j,t)`$ $`=`$ $`\widehat{E}_o^{(+)}(z_j,t)+{\displaystyle \frac{2\pi id}{S\mathrm{}}}{\displaystyle \underset{k}{}}\omega _k{\displaystyle \underset{i=1}{\overset{N_T}{}}}e^{ik(z_jz_i)}{\displaystyle _0^t}𝑑t^{}\widehat{\sigma }_i(t^{})e^{i\omega _k(t^{}t)}`$ (129) $``$ $`\widehat{E}_o^{(+)}(z_j,t)+\widehat{E}_s^{(+)}(z_j,t).`$ Here $$\widehat{E}_o^{(+)}(z,t)=i\underset{k}{}\left(\frac{2\pi \mathrm{}\omega _k}{S\mathrm{}}\right)^{1/2}\widehat{a}_k(0)e^{i\omega _kt}e^{ikz}$$ (130) is the homogeneous (“vacuum”) solution of the Maxwell equation for the quantized field, while $`\widehat{E}_s^{(+)}(z,t)`$ is the “source” part. Now in the mode continuum limit $`_k(\mathrm{}/2\pi )𝑑k=(\mathrm{}/2\pi c)𝑑\omega `$, $`\widehat{E}_s^{(+)}(z_j,t)`$ $`=`$ $`{\displaystyle \frac{2\pi id}{S\mathrm{}}}{\displaystyle \frac{\mathrm{}}{2\pi c}}{\displaystyle \underset{i=1}{\overset{N_T}{}}}{\displaystyle _0^t}𝑑t^{}\widehat{\sigma }_i(t^{}){\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\omega \omega e^{i\omega (t^{}t+[z_jz_i]/c)}`$ (131) $`=`$ $`{\displaystyle \frac{2\pi d}{Sc}}{\displaystyle \underset{i=1}{\overset{N_T}{}}}{\displaystyle _0^t}𝑑t^{}\dot{\widehat{\sigma }}_i(t^{})\delta (t^{}t+[z_jz_i]/c)`$ $``$ $`{\displaystyle \frac{2\pi id\omega _o}{Sc}}{\displaystyle \underset{i=1}{\overset{N_T}{}}}{\displaystyle _0^t}𝑑t^{}\widehat{\sigma }_i(t^{})\delta (t^{}t+[z_jz_i]/c)`$ $`=`$ $`{\displaystyle \frac{i\pi d\omega _o}{Sc}}\widehat{\sigma }_j(t)+{\displaystyle \frac{2\pi id\omega _o}{Sc}}{\displaystyle \underset{ij}{\overset{N_T}{}}}\widehat{\sigma }_i(t[z_jz_i]/c)\theta (z_jz_i)\theta (t[z_jz_i]/c)`$ $`=`$ $`{\displaystyle \frac{i\pi d\omega _o}{Sc}}\widehat{\sigma }_j(t)+\widehat{E}^{(+)}(z_j,t).`$ Here $`\widehat{E}^{(+)}(z_j,t)`$ denotes the field, at the position $`z_j`$ of atom $`j`$, that is produced by all the other atoms of the medium. We now use this result, and the operator identity $`\widehat{\sigma }_{zj}\widehat{\sigma }_j(t)=\widehat{\sigma }_j(t)`$, in equation (127). The result is $$\dot{\widehat{\sigma }}_j(t)=i\omega _o\widehat{\sigma }_j(t)\beta \widehat{\sigma }_j(t)\frac{id}{\mathrm{}}\widehat{\sigma }_{zj}(t)\widehat{E}^{(+)}(z_j,t),$$ (132) where $$\beta =\frac{\pi d^2\omega _o}{S\mathrm{}c}$$ (133) and $`\widehat{E}^{(+)}(z_j,t)=\widehat{E}_o^{(+)}(z_j,t)+\widehat{E}^{(+)}(z_j,t)`$. Similarly, using the operator identity $`\widehat{\sigma }_j^{}(t)\widehat{\sigma }_j(t)=\frac{1}{2}[1+\widehat{\sigma }_{zj}(t)]`$, we obtain from (126) and (131) $$\dot{\widehat{\sigma }}_{zj}(t)=2\beta [1+\widehat{\sigma }_{zj}(t)]\frac{2id}{\mathrm{}}[\widehat{E}^{()}(z_j,t)\widehat{\sigma }_j(t)\widehat{\sigma }_j^{}(t)\widehat{E}^{(+)}(z_j,t)].$$ (134) Since the expectation value $`\widehat{\sigma }_z`$ of the TLA inversion operator is $`p_2p_1=2p_21`$, where $`p_1`$ and $`p_2`$ are the lower– and upper–state probabilities, respectively, it follows that $`2\beta `$ is the radiative (spontaneous emission) decay rate: $$\frac{1}{\tau _{RAD}}=2\beta =\frac{2\pi d^2\omega _o}{S\mathrm{}c}.$$ (135) This is not the more familiar Einstein $`A`$ coefficient for spontaneous emission, $`A=4|𝐝|^2\omega _o^3/3\mathrm{}c^3`$, because it gives the spontaneous emission rate into modes propagating unidirectionally with a single polarization, whereas $`A`$ is the spontaneous emission rate into all possible field modes in free space. In fact $`1/\tau _{RAD}`$ is the spontaneous emission rate implicit in much of laser theory: the coefficient $`\lambda ^2A/8\pi `$ appearing in the standard expression for the gain coefficient $`g(\nu )`$, where $`\lambda `$ is the wavelength, is just $`1/\tau _{RAD}`$ times the cross–sectional area $`S`$: $$g(\nu )=\frac{\lambda ^2A}{8\pi }(N_2N_1)(\nu )=\frac{1}{\tau _{RAD}}(N_2N_1)S(\nu )$$ (136) for (nondegenerate) upper- and lower-level population densities $`N_2`$ and $`N_1`$, respectively, an atomic lineshape function $`(\nu )`$, and $`|𝐝|^2/3=d^2`$. Finally we use the definitions (28) and (30) of $`\widehat{F}`$ and $`\widehat{s}`$ to obtain $$\widehat{\dot{s}}_j(t)=i(\mathrm{\Delta }i\beta )\widehat{s}_j(t)\frac{id}{\mathrm{}}\widehat{\sigma }_{zj}(t)\widehat{F}(z_j,t),$$ (137) $$\dot{\widehat{\sigma }}_{zj}(t)=2\beta [1+\widehat{\sigma }_{zj}(t)]\frac{2id}{\mathrm{}}[\widehat{F}^{}(z_j,t)\widehat{s}_j(t)\widehat{s}_j^{}(t)\widehat{F}(z_j,t)]$$ (138) in the rotating–wave approximation. The detuning between the TLA resonance frequency $`\omega _o`$ and the central field frequency $`\omega `$ is defined as $`\mathrm{\Delta }=\omega _o\omega .`$ Replacing $`\widehat{s}_j(t)`$ and $`\widehat{\sigma }_{zj}(t)`$ by $`\widehat{s}(z_j,t)`$ and $`\widehat{\sigma }_z(z_j,t)`$, respectively, or $`\widehat{s}(z,t)`$ and $`\widehat{\sigma }_z(z,t)`$ in the continuum limit, we obtain equations (32) and (33). ## Appendix B For the initial state in which all the TLAs are in the upper state, $`\widehat{s}_i(t_o)=0`$ and $`\widehat{s}_i^{}(t_o)\widehat{s}_j(t_o)=\delta _{ij}`$. Then the operator $$\widehat{𝒮}=\underset{i=1}{\overset{N_T}{}}\widehat{s}_i(t_o)$$ (139) satisfies $$\widehat{𝒮}=0,$$ (140) $$\widehat{𝒮}^{}\widehat{𝒮}=\underset{i=1}{\overset{N_T}{}}\underset{j=1}{\overset{N_T}{}}\widehat{s}_i^{}(t_o)\widehat{s}_j(t_o)=N_T.$$ (141) In the continuum limit $$\widehat{𝒮}=\frac{N_T}{L}_0^L𝑑z\widehat{s}(z,0),$$ (142) $$\widehat{𝒮}^{}\widehat{𝒮}=\frac{N_T^2}{L^2}_0^L𝑑z^{}_0^L𝑑z^{\prime \prime }\widehat{s}^{}(z^{},t_o)\widehat{s}(z^{\prime \prime },t_o),$$ (143) and we can satisfy (140) and (141) by taking $$\widehat{s}(z,t_o)=0,$$ (144) $$\widehat{s}^{}(z^{},t_o)\widehat{s}(z^{\prime \prime },t_o)=\frac{L}{N_T}\delta (z^{}z^{\prime \prime }).$$ (145) Figure
warning/0004/math0004075.html
ar5iv
text
# Convexity of domains of Riemannian manifolds ## 1 Introduction In this paper we shall discuss the problem of the geodesic connectedness of subsets of Riemannian manifolds. In particular, we shall prove the geodesic connectedness of open domains (i.e. connected open subsets) $`𝒟`$ of a smooth Riemannian manifold $`(,,)`$, under reasonable assumptions; moreover, in the more relevant cases $`𝒟`$ will be shown to be convex, i.e. any pair of its points can be joined by a (non necessarily unique) minimizing geodesic<sup>1</sup><sup>1</sup>1The word “convex” is used in different non-equivalent ways in the literature. Sometimes, it is reserved for open domains such that each two points can be joined by an unique minimizing geodesic; for those following this convention, a better name for our domains would be weakly convex.. As pointed out by Gordon , this problem is important not only by its own but also because of its relation, via the Jacobi metric, to the problem of connecting two points by means of a trajectory of fixed energy for a Lagrangian system. Until now this topic has been faced by using different techniques and under different assumptions which allow us to control the non–completeness of $`𝒟`$. In particular, geodesic connectedness can be proved by using variational methods. In this case, the right assumption to get existence, and multiplicity in same cases, of geodesics connecting two fixed points, is a convexity assumption on the boundary of $`𝒟`$ (see e.g. ); here we shall work under weaker assumptions. Our study makes necessary to discuss the different notions of convexity for the (possibly singular) boundary points of the open domain $`𝒟`$. In what follows, differentiability will mean $`𝒞^4`$; indeed, we shall need just $`𝒞^3`$ for our main results (cfr. Theorems 1.6, 1.8, and $`𝒞^2`$ for Theorem 1.5), but in our references the highest assumption of differentiability is $`𝒞^4`$; so, we prefer stating the results assuming it. At first, recall that, by the well–known Hopf–Rinow theorem, if a Riemannian manifold $``$ is complete then it is geodesically connected. From a variational point of view, the Hopf–Rinow theorem can be easily proved by using the functional $$f(x)=\frac{1}{2}_0^1\dot{x}(s),\dot{x}(s)𝑑s$$ (1.1) defined on a suitable Hilbert manifold, see Section 3. It is well–known that the critical points of $`f`$ are geodesics and it is not difficult to prove that $`f`$ admits a minimum point. Then, in the complete case, this result guarantees not only that the manifold is geodesically connected, but also that it is convex. We examine now the case when the boundary $`𝒟`$ of $`𝒟`$ in $``$ is differentiable, that is, $`\overline{𝒟}=𝒟𝒟`$ is a Riemannian manifold with (differentiable) boundary. Recall the following two natural notions of convexity around a point of the boundary. ###### Definition 1.1 (Infinitesimal convexity) We say that $`𝒟`$ is infinitesimally convex at $`p𝒟`$ if the second fundamental form $`\sigma _p`$, with respect to the interior normal, is positive semidefinite. ###### Definition 1.2 (Local convexity) We say that $`𝒟`$ is locally convex at $`p𝒟`$ if there exists a neighborhood $`\overline{U}\overline{𝒟}`$ of $`p`$ such that $$\mathrm{exp}_p\left(T_p𝒟\right)\left(\overline{U}𝒟\right)=\mathrm{}.$$ (1.2) It is not difficult to show that the local convexity implies the infinitesimal one, but the converse is not true. Nevertheless, if the infinitesimal convexity is assumed on a neighbourhood of a point of the boundary, then the notions are equivalent, as proved by Bishop . In order to apply variational methods to the study of geodesic connectedness, a characterization of the infinitesimal convexity is useful. Firstly, note that, by the differentiability of the boundary, for each $`p𝒟`$ there exist a neighborhood $`U`$ of $`p`$ and a differentiable function $`\varphi :U\overline{𝒟}𝐑`$ such that $$\{\begin{array}{ccc}\varphi ^1(0)=U𝒟\hfill & & \\ \varphi >0\hfill & \text{on }U𝒟\text{ }\hfill & \\ \varphi (q)0\hfill & \text{for any }qU𝒟.\hfill & \end{array}$$ (1.3) Then, it is easy to check that $`𝒟`$ is infinitesimally convex at $`p𝒟`$ if and only if for one (and then for all) function $`\varphi `$ satisfying (1.3) we have $$H_\varphi (p)[v,v]0vT_p𝒟.$$ (1.4) Now, we shall go from considerations around a point of the boundary, through considerations on all the boundary, and, so, the completeness of $`\overline{𝒟}`$ becomes essential. Note that if $``$ is complete then so is $`\overline{𝒟}`$ and, even though the converse is not true, there is no loss of generality assuming it, because the Riemannian metric can be modified out of $`\overline{𝒟}`$ to obtain completeness (see for example ). By standard arguments the function $`\varphi `$ in (1.3) can be found on all $`\overline{𝒟}`$, and, thus, we have the following equivalent definitions for the convexity of all the boundary. ###### Definition 1.3 (Global convexity, variational point of view) Assume that $``$ is complete. $`𝒟`$ is convex if and only if for one, and then for all, nonnegative function $`\varphi `$ on $`\overline{𝒟}`$ such that $$\{\begin{array}{ccc}\varphi ^1(0)=𝒟\hfill & & \\ \varphi >0\hfill & \text{on }𝒟\hfill & \\ \varphi (q)0,\hfill & \text{for any }q𝒟\hfill & \end{array}$$ (1.5) we have $$H_\varphi (q)[v,v]0q𝒟,vT_q𝒟.$$ (1.6) It is worth pointing out that condition (1.6) is equivalent to a geometric notion of convexity. Indeed the following definition is equivalent too, see . ###### Definition 1.4 (Global convexity, geometrical point of view) Assume that $``$ is complete. $`𝒟`$ is convex if for any $`p,q𝒟`$ the range of any geodesic $`\gamma :[0,1]\overline{𝒟}`$ such that $`\gamma (0)=p,\gamma (1)=q`$ satisfies $$\gamma \left([0,1]\right)𝒟.$$ (1.7) Condition (1.7) is a generalization to Riemannian manifolds of the usual notion of convexity given in Euclidean spaces. Moreover, all the above conditions provide different ways to prove that, when $``$ is complete: $`𝒟`$ is convex if and only if $`𝒟`$ is convex. In fact, a variational technique based on the use of the functional (1.1) and a penalization argument make possible to prove that if $`𝒟`$ is convex, then $`𝒟`$ is convex. On the other hand, by using Definition 1.4 it is easy to show that if $`𝒟`$ is not convex then neither is $`𝒟`$. Now we are ready to examine the general case where $`𝒟`$ is not differentiable or $`\overline{𝒟}`$ is not complete. By using the results above, it is clear that if there exists a sequence $$\left(\overline{𝒟}_m\right)_{m𝐍}$$ of complete submanifolds with convex (differentiable) boundary such that $$\overline{𝒟}_m\overline{𝒟}_{m+1}\text{and}𝒟=\underset{m𝐍}{}\overline{𝒟}_m,$$ (1.8) then $`𝒟`$ is geodesically connected. As a first question we can wonder if $`𝒟`$ must be convex. In Section 2 we answer this question, by showing that if $`\overline{𝒟}`$ is complete then $`𝒟`$ is convex. More precisely, let $`\overline{𝒟}^c`$ be the canonical completation of $`𝒟`$ by using Cauchy sequences, and $`_c𝒟`$ the corresponding boundary points, $`\overline{𝒟}^c=𝒟_c𝒟`$ ($`\overline{𝒟}^c`$ is always complete as a metric space, but the boundary points in $`_c𝒟`$ are not necessarily differentiable and, if they are, the metric may be non–extendible or degenerate). Note that any point of $`𝒟`$ naturally determines one or more points in $`_c𝒟`$, and $`\overline{𝒟}`$ is complete if and only if all the points in $`_c𝒟`$ are of this type (in this case, we can assume that $``$ is complete). Then we will prove: ###### Theorem 1.5 Assume that $`\overline{\varphi }:\overline{𝒟}^c[0,\mathrm{}]`$ is a continuous function such that: * $`\overline{\varphi }^1(0)=_c𝒟`$; * there exists an infinitesimal sequence $`(a_m)_{m𝐍}`$ such that $`\overline{\varphi }^1(]a_m,\mathrm{}])`$ is a Riemannian submanifold with convex (differentiable) boundary $`\overline{\varphi }^1(a_m)`$ for any $`m𝐍`$. Then $`𝒟`$ is geodesically connected. Moreover, if $``$ is complete then $`𝒟`$ is convex. We point out that in order to obtain convexity, the assumption of completeness on $``$ cannot be removed, as we shall shown in Section 2 by a counterexample. This result is proved by using geometrical methods, and it makes possible to generalize the results by Gordon in by showing that his hypotheses imply those in Theorem 1.5 (Section 2). Our main objective in this paper is achieved in Section 3, where we use variational methods (under the natural assumption of completeness for $``$) to show that $`𝒟`$ is convex when there exists a sequence $`\left(𝒟_m\right)_{m𝐍}`$ of open domains invading $`𝒟`$ as in (1.8), whose boundaries are differentiable but not necessarily convex (and, thus, each $`𝒟_m`$ may be non–geodesically connected), if a suitable estimate of the loss of convexity of $`𝒟_m`$ and boundness of the sequence is assured. More precisely, the following result will be proved. ###### Theorem 1.6 Let $``$ be a complete Riemannian manifold and $`𝒟`$ an open domain of $``$. Assume that there exists a positive differentiable function $`\varphi `$ on $`𝒟`$ such that * $`lim_{x𝒟}\varphi (x)=0`$; * each $`y𝒟`$ admits a neighbourhood $`U`$ and constants $`a,b>0`$ such that $$a\varphi (x)bx𝒟U;$$ * the first and second derivatives of the normalized flow of $`\varphi `$ are locally bounded close to $`𝒟`$, that is: each $`y𝒟`$ admits a neighbourhood $`U`$ such that the induced local flow on $`𝒟U`$ has first and second derivatives with bounded norms; * there exist a decreasing and infinitesimal sequence $`(a_m)_{m𝐍}`$ such that each $`y𝒟`$ admits a neighbourhood $`U`$ and a constant $`M𝐑`$ satisfying: $$H_\varphi (x)[v,v]Mv,v\varphi (x)x\varphi ^1(a_m)U,vT_x\varphi ^1(a_m),m𝐍.$$ (1.9) Then $`𝒟`$ is convex. Moreover if $`𝒟`$ is not contractible in itself, then for any $`p,q𝒟`$ there exists a sequence $`(x_m)_{m𝐍}`$ of geodesics in $`𝒟`$ joining them such that $$\underset{m\mathrm{}}{lim}f(x_m)=\mathrm{}.$$ ###### Remarks 1.7 When the boundary $`𝒟`$ is smooth and convex in the sense of Definition 1.3, the function $`\varphi `$ in (1.5) always satisfies all conditions (i)–(iv) above. Let us examine the role of each one of these hypotheses. (1) Hypotheses (i), (iv): they imply first that, taking $`𝒟_m=\varphi ^1(]a_m,\mathrm{}[)`$, condition (1.8) is satisfied, and, second, that even when the boundaries $`𝒟_m`$ may be non–convex, their loss of convexity is (locally) bounded by (1.9). (2) Hypothesis (ii): from Theorem 1.5, if (iv) is satisfied with $`M0`$ then (ii) can be replaced just by: $$\varphi (x)0,x\varphi ^1(a_m),m𝐍.$$ But when $`M>0`$ the hypothesis (ii) must be imposed to make (iv) meaningful. The reason is that the left hand side in (1.9) describes the shape of $`\varphi ^1(a_m)`$, but the value of $`\varphi `$ in the right hand side can be almost arbitrarily changed, if no bound on the gradient is imposed. More precisely, consider any smooth function $`\phi :]0,\mathrm{}[]0,\mathrm{}[`$ such that $`lim_{s0}\phi (s)=0`$ and its derivative satisfies $`\dot{\phi }>0`$. Then $`\varphi ^{}=\phi \varphi `$ also satisfies (i), and $$H_\varphi ^{}(x)[v,v]=\dot{\phi }(\varphi (x))H_\varphi (x)[v,v]x\varphi ^1(a_m^{}),vT_x\varphi ^1(a_m^{}),$$ where $`a_m^{}=\phi (a_m),m𝐍`$. When $`\varphi `$ satisfies (ii) then $`\varphi ^{}`$ satisfies (ii) if and only if $`a^{}\dot{\phi }b^{}`$, close to 0, for some $`a^{},b^{}>0`$. In this case, $`\varphi `$ satisfies (iv) if and only if so does $`\varphi ^{}`$. But if (ii) were not imposed, it would be possible that one of the functions satisfies (iv) and the other does not<sup>2</sup><sup>2</sup>2In fact, assume that $`\varphi `$ satisfies (iv), but there exists an infinitesimal sequence $`(a_m)_{m𝐍}`$ such that $`H_\varphi (x_m)[v_m,v_m]>0`$ for some $`x_m`$ at each $`\varphi ^1(a_m)`$ and unitary vector $`v_mT_{x_m}\varphi ^1(a_m)`$. Then take $`(k_m)_{m𝐍}`$ such that $`k_m>0,`$ $`\left(k_mH_\varphi (x_m)[v_m,v_m]\right)_{m𝐍}`$ is not infinitesimal. Clearly any $`\phi `$ as above such that $`\dot{\phi }(a_m)=k_m`$ yields the required example.. This shows that (iv) is not reasonable by itself as a measure of the loss of convexity of the hypersurfaces $`\varphi ^1(a_m)=\varphi ^1(a_m^{})`$, being hypothesis (ii) natural. (3) Hypothesis (iii): the bounds on the normalized flow (i.e. the flow of $`\varphi /\varphi ^2`$) are technical, and they express the unique control we impose on the intermediate hypersurfaces between two consecutive $`\varphi ^1(a_m)`$. Note that if (iii) were not imposed then hypersurfaces arbitrarily close to $`𝒟`$ “very distorted” by the flow could exist. Technical condition (iii) and even the completeness of the ambient manifold $``$ can be weakened if (iv) is imposed on all points and directions enough close to the boundary. So, a straightforward consequence of the technique in the proof of Theorem 1.6 is the following result (compare with ): ###### Theorem 1.8 Let $``$ be a Riemannian manifold, $`𝒟`$ an open domain, and $`\overline{𝒟}^c=𝒟_c𝒟`$ its canonical Cauchy completation. Assume that there exists a positive differentiable function $`\varphi `$ on $`𝒟`$ such that * $`lim_{x_c𝒟}\varphi (x)=0`$; * each $`y_c𝒟`$ admits a neighbourhood $`U\overline{𝒟}^c`$ and constants $`a,b>0`$ such that $$a\varphi (x)bxU𝒟;$$ * each $`y_c𝒟`$ admits a neighbourhood $`U\overline{𝒟}^c`$ and a constant $`M𝐑`$ such that inequality (1.9) holds for all $`x𝒟U`$ and for all $`vT_x`$. Then $`𝒟`$ is convex. Moreover if $`𝒟`$ is not contractible in itself, for any $`p,q𝒟`$ there exists a sequence $`(x_m)_{m𝐍}`$ of geodesics in $`𝒟`$ joining them such that $$\underset{m\mathrm{}}{lim}f(x_m)=\mathrm{}.$$ ###### Remarks 1.9 Note that when $`_c𝒟`$ is convex in the sense of Definition 1.3 the function $`\varphi `$ in (1.5) does not necessarily satisfy (iii) because this condition is now imposed on all tangent vectors $`v`$. Nevertheless, when Definition 1.3 is applicable, it is independent of the chosen $`\varphi `$; thus, varying $`\varphi `$, all tangent $`v`$ can be considered as tangent to a level hypersurface. So, we can conclude that the hypotheses in Theorems 1.6 and 1.8 imply, for each point of the boundary, local conditions which extend those for differentiable boundaries. In Section 3 Theorem 1.6 will be proved. To this aim, we shall penalize the functional $`f`$ of (1.1) with a term depending on a positive parameter $`ϵ`$ and we shall study the Euler–Lagrange equation associated to the penalized functionals $`f_ϵ`$. The crucial point is to prove that a critical point of $`f_ϵ`$ in a sublevel of $`f_ϵ`$ is uniformly far (with respect to $`ϵ`$) from $`𝒟`$. In the proof we shall “project” the critical points of the penalized functionals (using the normalized flow of $`\varphi `$) on the hypersurface $`\varphi ^1(a_m)`$ for $`m`$ large enough. This makes possible to get critical points of $`f`$ (i.e. geodesics) not touching $`𝒟`$ by means of a limit process. Finally, in Section 4 the discussion of Theorems 1.6, 1.8 is completed by giving: (a) some examples which show the applicability and independence of the hypotheses of Theorems 1.5, 1.6, 1.8, and (b) an application to the existence of trajectories of fixed energy for dynamical systems. ## 2 Proof of Theorem 1.5 and Gordon’s theorem Proof of Theorem 1.5. Given $`p,q𝒟`$ choose $`a_m<\mathrm{min}\{\overline{\varphi }(p),\overline{\varphi }(q)\}`$. By (i) and the continuity of $`\overline{\varphi }`$ at $`_c𝒟`$, the manifold $`\overline{𝒟}_m=\overline{\varphi }^1([a_m,\mathrm{}])`$ is complete. Then $`p`$ and $`q`$ belong to the interior of $`\overline{\varphi }^1([a_m,\mathrm{}])`$, which is intrinsically convex. Now let us show that if $``$ is complete then $`𝒟`$ is convex. Let $`m_0𝐍`$ be the first positive integer such that $`a_{m_0}<\mathrm{min}\{\overline{\varphi }(p),\overline{\varphi }(q)\}`$. As before we get for any $`mm_0`$ the existence of a minimizing geodesic $`\gamma _m`$ in $`𝒟_m`$ joining $`p`$ and $`q`$. Let us consider $`\gamma _m`$ parametrized by arc length, so $`\gamma _m:[0,l_m]𝒟_m`$, $`\dot{\gamma }_m(s),\dot{\gamma }_m(s)=1`$, for any $`s`$. Note that, up to a subsequence, $`\dot{\gamma }_m(0)v`$, with $`v`$ unitary vector, and that $`(l_m)_{m𝐍}`$ is a decreasing sequence converging to the distance $`l`$ between $`p`$ and $`q`$. By standard arguments, the geodesic $`\gamma :[0,l]`$ with $`\dot{\gamma }(0)=v`$ has range in $`𝒟𝒟`$, joins $`p`$ and $`q`$, and satisfies $$\gamma _m\gamma \text{uniformly in }[0,l]\text{.}$$ If, for infinitely many $`m𝐍`$, $`\gamma \gamma _m`$ the proof is complete. So let us assume $`\dot{\gamma }_m(0)v`$ for infinitely many $`m`$ and $`(l_m)_{m𝐍}`$ strictly decreasing: note that, necessarily, $`p`$ and $`q`$ are conjugate along $`\gamma `$. Let $`U`$ be a star–shaped neighbourhood of $`q`$, with $`U𝒟_{m_0}`$. We shall prove that, for small $`\delta `$, $`q_\delta =\gamma (l\delta )U`$ is always conjugate to $`p`$ along $`\gamma `$ getting a contradiction. Since $`p,q_\delta 𝒟_m`$ for any $`mm_0`$, reasoning as before, we can find a sequence of minimizing geodesics parametrized by arc length $`\overline{\gamma }_m:[0,\overline{l}_m]𝒟_m`$ joining $`p`$ and $`q_\delta `$, a subsequence $`\dot{\overline{\gamma }}_m(0)\overline{v}`$, with $`\overline{v}`$ unitary vector (which we can assume distinct from $`v`$, otherwise $`q_\delta `$ is also conjugate and the proof is complete) and $`\overline{\gamma }:[0,\overline{l}]𝒟𝒟`$ with $`\dot{\overline{\gamma }}(0)=\overline{v}`$. Again $`(\overline{l}_m)_{m𝐍}`$ is decreasing and converging to the distance $`\overline{l}`$ between $`p`$ and $`q_\delta `$. We claim $`\overline{l}l\delta `$; otherwise, modifying slightly the curves $`\gamma _m`$, we could find a curve joining $`p`$ and $`q_\delta `$ contained in $`𝒟_m`$ for $`m`$ large, with length less than $`\overline{l}`$, which is a contradiction. Now we can define the union curve of $`\overline{\gamma }`$ on $`[0,\overline{l}]`$ and $`\gamma `$ restricted to $`[l\delta ,l]`$ which joins $`p`$ and $`q`$ and with length less or equal than $`l`$. As this union curve is not differentiable at $`\overline{l}`$, we can slightly modify it to obtain a curve $`\widehat{\gamma }`$ with length less than $`l`$ and which agrees with $`\overline{\gamma }`$ out of $`U`$. Finally, we could find for large $`m`$ a curve joining $`p`$ and $`q`$, with length less than $`l`$ and which agrees with some $`\overline{\gamma }_m`$ out of $`U`$, getting an absurd with the minimality of $`\gamma _m`$ since $`U𝒟_m`$. $`\mathrm{}`$ At the end of this section we shall give a counterexample for the case $``$ non complete. The possibility of extending Gordon’s results by using variational methods has already been pointed out in \[5, Chapter 4\], . Nevertheless our point of view is quite different, and the extension we obtain is, at any case, elementary and stronger. In fact, in Gordon’s result a convexity assumption is done on the whole manifold; in the quoted references it is claimed that this global assumption must imply a convexity property close to the boundary, which should be enough from a variational point of view; finally, we will check now that the global assumption imply a convexity property for a sequence of hypersurfaces close to the boundary, which is enough from any of the points of view sketched in Section 1. We recall that a map $`h`$ between manifolds is said to be proper if $`h^1(K)`$ is compact whenever $`K`$ is compact. In particular, if $`h:𝒟𝐑`$ is proper necessarily $`|h(p)|\mathrm{}`$ as $`p𝒟`$. Recall also that a (real–valued $`𝒞^2`$) function is called convex when its Hessian is positive semidefinite. Gordon’s result \[4, Theorem 1\] asserts: ###### Theorem 2.1 If the open domain $`𝒟`$ of the Riemannian manifold $``$ supports a proper positive convex function $`h`$, then it is geodesically connected. We will reprove this result, by showing that its hypotheses imply the ones in Theorem 1.5. There is a second theorem in Gordon’s paper, which can be reproved in the same way. Recall first that Theorem 1.5 can also be stated assuming $`\overline{\varphi }^1(\mathrm{})=_c𝒟`$ and $`(a_m)_{m𝐍}`$ diverging to $`\mathrm{}`$. Proof of Theorem 2.1. By Sard’s theorem, almost all the values of $`h`$ are regular. Thus, as $`h`$ is proper (and positive), there exists a diverging sequence of regular values $`(a_m)_{m𝐍}`$ contained in the range of $`h`$. Moreover, $`h^1([0,a_m])`$ is a compact Riemannian manifold with boundary $`h^1(a_m)`$. From Definition 1.3, this boundary is convex (put $`\varphi =h(a_m)h`$). So, extending continuously $`h`$ to a function $`\overline{h}:\overline{𝒟}^c[0,\mathrm{}]`$, Theorem 1.5 (in the version above) can be claimed. $`\mathrm{}`$ A counterexample. The following counterexample shows that the result in Theorem 1.5 on geodesic connectedness cannot be strengthened to obtain convexity, when $``$ is not complete. Consider two open hemispheres $`H_0,H_1`$ in $`𝐑^3`$ and let $`x_0,x_1`$ be their north poles (see Fig. 1). Put a sequence of inmersed tubes $`(T_m)_{m𝐍}`$ connecting $`H_0`$ and $`H_1`$ of decreasing length and such that any curve joining $`x_0`$ and $`x_1`$ through $`T_m`$ is longer than a minimizing curve joining them through $`T_{m+1}`$. We also assume that the width of these tubes goes to zero, and their mouths in each hemisphere go to a point $`e_i,i=0,1`$ in the equator, being all their centers in the same meridian (the shape of the resulting hemispheres is shown in Fig. 2). Let $``$ be this manifold inmersed in $`𝐑^3`$, and $`𝒟=`$. Recall that $`_c𝒟`$ is canonically identifiable to the equators, and let $`\varphi :𝒟𝐑`$ be the height function with $`lim_{p_c𝒟}\varphi (p)=0`$. Clearly, a sequence $`(a_m)0`$ can be chosen such that $`\overline{𝒟}_m=\varphi ^1([a_m,\mathrm{}[)`$ is a complete Riemannian manifold with convex boundary $`𝒟_m=\varphi ^1(a_m)`$, containing the tube $`T_m`$ but not $`T_{m+1}`$, and satisfying (1.8). So, in each convex manifold $`𝒟_m=\varphi ^1(]a_m,\mathrm{}[)`$ there exists a minimizing geodesic $`\gamma _m`$ connecting $`x_0,x_1`$, and, by the condition on the lengths of the tubes, $$\text{length}(\gamma _{m+1})<\text{length}(\gamma _m).$$ So, if there was a minimizing geodesic $`\gamma `$ between $`x_0,x_1`$ in $`𝒟`$, necessarily it should be included in some $`𝒟_m`$ and $`\text{length}(\gamma _{m+1})<\text{length}(\gamma )`$, a contradiction. ## 3 Proof of Theorems 1.6 and 1.8 Before introducing the functional framework, we recall that, by the well–known Nash embedding Theorem (see ), any smooth Riemannian manifold $``$ is isometric to a submanifold of $`𝐑^N`$, with $`N`$ sufficiently large, equipped with the metric induced by the Euclidean metric in $`𝐑^N`$. So, henceforth, we shall assume that $``$ is a submanifold of $`𝐑^N`$ and $`,`$ is the Euclidean metric. It is well–known that the geodesics in $`𝒟`$ joining two fixed points $`p`$ and $`q`$ of $`𝒟`$ are the critical points of the action integral (1.1) defined on $`\mathrm{\Omega }^1\left(𝒟\right)`$ where $$\mathrm{\Omega }^1\left(𝒟\right)=\{xH^{1,2}([0,1],𝒟)x(0)=p,x(1)=q\}$$ and $$H^{1,2}([0,1],𝒟)=\{xH^{1,2}([0,1],𝐑^N)x([0,1])𝒟\}.$$ It can be proved that $`\mathrm{\Omega }^1\left(𝒟\right)`$ is a Hilbert submanifold of $`H^{1,2}([0,1],𝒟)`$ whose tangent space at $`x\mathrm{\Omega }^1\left(𝒟\right)`$ is given by $$T_x\mathrm{\Omega }^1\left(𝒟\right)=\{vH^{1,2}([0,1],T𝒟)v(s)T_{x(s)}𝒟v(0)=0=v(1)\}.$$ We recall the following definition. ###### Definition 3.1 Let $`(X,g)`$ be a Riemannian manifold modelled on a Hilbert space and let $`F𝒞^1(X,𝐑)`$. We say that $`F`$ satisfies the Palais–Smale condition if every sequence $`(x_m)_{m𝐍}`$ such that $$(F(x_m))_{m𝐍}\text{ is bounded},$$ (3.1) $$F(x_m)0,$$ (3.2) contains a converging subsequence, where $`F(x)`$ denotes the gradient of $`F`$ at the point $`x`$ with respect to the metric $`g`$ and $``$ is the norm on the tangent bundle induced by $`g`$. A sequence satisfying (3.1)–(3.2) is said a Palais–Smale sequence. In our case there are Palais–Smale sequences that could converge to a curve which “touches” the boundary $`𝒟`$, so we penalize the functional $`f`$ in a suitable way, following . For any $`ϵ]0,1]`$, we consider on $`\mathrm{\Omega }^1\left(𝒟\right)`$ the functional $$f_ϵ(x)=f(x)+_0^1\frac{ϵ}{\varphi ^2(x)}𝑑s$$ (3.3) where $`\varphi `$ has been introduced in Theorem 1.6. For any $`ϵ]0,1]`$ $`f_ϵ`$ is a $`𝒞^2`$ functional and if $`x\mathrm{\Omega }^1\left(𝒟\right)`$ is a critical point of $`f_ϵ`$, by using a boot–strap argument it can be proved that it is $`𝒞^2`$ and satisfies $$D_s\dot{x}=\frac{2ϵ}{\varphi ^3(x)}\varphi (x).$$ (3.4) For any $`ϵ]0,1]`$, $`s[0,1]`$, we set $$\lambda _ϵ(s)=\frac{2ϵ}{\varphi ^3(x(s))},$$ (3.5) which represents the multiplier in (3.4). Multiplying (3.4) by $`\dot{x}`$, it is easy to get the existence of a constant $`E_ϵ(x)𝐑`$ such that $$\frac{1}{2}\dot{x}(s),\dot{x}(s)\frac{ϵ}{\varphi ^2(x(s))}=E_ϵ(x)s[0,1].$$ (3.6) To prove that the penalized functionals satisfy the Palais–Smale condition, we recall the following lemma which holds with slight variants of the proof in also under our local assumptions (i)–(ii) of Theorem 1.6. ###### Lemma 3.2 Let $`(x_m)_{m𝐍}`$ be a sequence in $`\mathrm{\Omega }^1\left(𝒟\right)`$ such that $$\underset{m𝐍}{sup}_0^1\dot{x}_m,\dot{x}_m𝑑s<\mathrm{}$$ (3.7) and assume the existence of a sequence $`(s_m)_{m𝐍}`$ in $`[0,1]`$ such that $$\underset{m\mathrm{}}{lim}\varphi (x_m(s_m))=0.$$ (3.8) Then $$\underset{m\mathrm{}}{lim}_0^1\frac{1}{\varphi ^2(x_m(s))}𝑑s=\mathrm{}.$$ (3.9) ###### Proposition 3.3 Let $`f_ϵ`$ be as in (3.3). Then * for any $`ϵ]0,1]`$ and for any $`c𝐑`$ the sublevels $$f_ϵ^c=\{x\mathrm{\Omega }^1\left(𝒟\right)|f_ϵ(x)c\}$$ are complete metric subspaces of $`\mathrm{\Omega }^1\left(𝒟\right)`$; * for any $`ϵ]0,1]`$, $`f_ϵ`$ satisfies the Palais–Smale condition. Proof: For any $`ϵ]0,1],c𝐑`$, let $`(x_m)_{m𝐍}`$ be a Cauchy sequence in $`f_ϵ^c`$, then it is a Cauchy sequence also in $`H^{1,2}([0,1],𝐑^N)`$, so it converges strongly to a curve $`x`$ in $`H^{1,2}([0,1],𝐑^N)`$. Since this convergence is also uniform, by Lemma 3.2 it results that $`x\mathrm{\Omega }^1\left(𝒟\right)`$ and by the continuity of $`f_ϵ`$, we obtain the first part of the proposition. Now let $`(x_m)_{m𝐍}`$ be a Palais–Smale sequence; in particular it results that $$_0^1\dot{x}_m,\dot{x}_m𝑑s\text{is bounded}.$$ (3.10) Then, up to a subsequence, we get the existence of a $`xH^{1,2}([0,1],𝐑^N)`$ such that $$x_mx\text{ weakly in }H^{1,2}([0,1],𝐑^N).$$ (3.11) Arguing as in the first part of the proof, we get that $`x\mathrm{\Omega }^1\left(𝒟\right)`$. Using standard arguments, it can be proved that $$x_mx\text{ strongly in }H^{1,2}([0,1],𝐑^N).\mathrm{}$$ ###### Remark 3.4 By Proposition 3.3, for any $`ϵ]0,1]`$, $`f_ϵ`$ has a minimum point $`x_ϵ\mathrm{\Omega }^1\left(𝒟\right)`$; it is easy to see that there exists $`k>0`$ such that $$f_ϵ(x_ϵ)k,$$ for any $`ϵ]0,1]`$. Moreover, by (3.6) we get for any $`ϵ]0,1]`$ $$E_ϵ(x_ϵ)=f_ϵ(x_ϵ)2_0^1\frac{ϵ}{\varphi ^2(x_ϵ(s))}𝑑sk,$$ hence $$\frac{1}{2}\dot{x}_ϵ(s),\dot{x}_ϵ(s)k+\frac{ϵ}{\varphi ^2(x_ϵ(s))},$$ (3.12) for any $`ϵ]0,1]`$, $`s[0,1]`$. ###### Remark 3.5 In the sequel, we shall need to relate the Hessian of a $`\widehat{\mathrm{\Psi }}𝒞^2(𝐑^N,𝐑)`$ to the one of its restriction $`\mathrm{\Psi }`$ on $``$. For any $`y`$ let $$P(y):𝐑^NT_y,$$ (3.13) $$Q(y):𝐑^NT_y^{},$$ (3.14) be respectively the projections on $`T_y`$ and $`T_y^{}`$. Since $``$ is a $`𝒞^3`$ submanifold of $`𝐑^N`$, there exist $`A_j^i𝒞^2(,𝐑)`$, $`i`$,$`j\{1,\mathrm{},N\}`$ such that for any $`y,v=(v^1,\mathrm{},v^N)𝐑^N`$ $$Q(y)[v]=\underset{i,j=1}{\overset{N}{}}A_j^i(y)v^je_i,$$ where $`e_1,\mathrm{},e_N`$ is the canonical basis of $`𝐑^N`$. We locally extend the functions $`A_j^i`$ to $`𝒞^2`$ functions (still denoted by $`A_j^i`$) on $`𝐑^N`$. For any $`y𝐑^N`$, we define the differential map $`\mathrm{d}Q(y):𝐑^N\times 𝐑^N𝐑^N`$ as $$\mathrm{d}Q(y)[v,w]=\underset{i,j,k=1}{\overset{N}{}}\frac{A_j^i(y)}{x_k}v^kw^je_i,v,w𝐑^N.$$ Even if $`\mathrm{d}Q`$ could depend on the extensions of the functions $`A_j^i`$, for any $`y`$, the restriction of $`\mathrm{d}Q(y)`$ to $`T_y\times T_y`$ is well–defined. It can be proved (see e.g. \[3, Lemma 8\]) that for any $`y`$, $`vT_y`$: $$H_\mathrm{\Psi }(y)[v,v]=\mathrm{d}^2\widehat{\mathrm{\Psi }}(y)[v,v]\mathrm{d}\widehat{\mathrm{\Psi }}(y)[\mathrm{d}Q(y)[v,v]],$$ where $`d`$ and $`d^2`$ are the differential map and the second differential map on $`𝐑^N`$. ###### Lemma 3.6 Let $`\left(x_ϵ\right)_{ϵ>0}`$ be a family in $`\mathrm{\Omega }^1(𝒟)`$ of critical points of $`f_ϵ`$ such that $$f_ϵ(x_ϵ)kϵ]0,1],$$ (3.15) for a suitable positive constant $`k`$. Then $`\left(\lambda _ϵ(s)={\displaystyle \frac{2ϵ}{\varphi ^3(x_ϵ(s))}}\right)_{ϵ>0}`$ is bounded in $`L^{\mathrm{}}([0,1],𝐑)`$. Proof: Let $`\left(ϵ_m\right)_{m𝐍}`$ be a decreasing and infinitesimal sequence in $`]0,1]`$ and let $`\left(x_{ϵ_m}\right)_{m𝐍}`$ be a sequence of critical points of $`f_{ϵ_m}`$ satisfying (3.15). For the sake of simplicity, in the following we set $`x_{ϵ_m}x_m`$, $`\lambda _{ϵ_m}\lambda _m`$. Now let $`u_m(s)=\varphi (x_m(s))`$, for $`s[0,1],m𝐍`$ and $`u_m(s_m)=\mathrm{min}_{s[0,1]}u_m(s)`$, for any $`m𝐍`$. It suffices to prove the lemma when, up to a subsequence, $$\underset{m\mathrm{}}{lim}u_m(s_m)=0.$$ (3.16) By (3.15) and the Poincaré inequality there exists $`xH^{1,2}([0,1],𝐑^N)`$ such that $$x_mx\text{uniformly},$$ (3.17) and since, up to a subsequence, $$\underset{m\mathrm{}}{lim}s_m=s_0,$$ (3.18) by (3.16) and (3.17) it easily follows $`s_0]0,1[`$ since $`\varphi (p),\varphi (q)>0`$. It is not difficult to see that $`\left(x_m(s_m)\right)_{m𝐍}`$ converges to $`x(s_0)=y𝒟`$. Let $`U`$ be a neighborhood of $`y`$ such that (ii)(iv) of Theorem 1.6 hold, then there exists $`\mu >0`$ such that $$x_m(s)U𝒟$$ (3.19) for any $`sJ=[s_0\mu ,s_0+\mu ]`$ and $`m`$ sufficiently large. By (3.4), (3.5), (3.16) and (ii) of Theorem 1.6, we get for m large enough $`\ddot{u}_m(s_m)=H_\varphi (x_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]\lambda _m(s_m)\varphi (x_m(s_m))^2`$ $`H_\varphi (x_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]a^2\lambda _m(s_m).`$ (3.20) By the assumptions of Theorem 1.6 and (3.16), for any $`m𝐍`$ there exists $`k_m𝐍`$ such that $$a_{k_m}\varphi (x_m(s_m))\varphi (x_m(s))$$ for any $`s[0,1]`$. Let us consider the Cauchy problem $$\{\begin{array}{cc}\dot{\eta }=\frac{\varphi (\eta )}{\varphi (\eta )^2}\hfill & \\ \eta (0)=xU𝒟\hfill & \end{array}$$ (3.21) and call $`\eta (s,x):𝒰𝐑\times \overline{𝒟}\overline{𝒟}`$ the flow associated to the Cauchy problem (3.21), where $`𝒰`$ is the maximal domain where the flow can be defined. Set for any $`sJ`$ $`\tau _m(s)=\varphi (x_m(s))a_{k_m}`$ (3.22) $`y_m(s)=\eta (\tau _m(s),x_m(s)).`$ (3.23) Observe that if $`m`$ is sufficiently large and $`\mu `$ is opportunely chosen $$(\tau _m(s),x_m(s))𝒰y_m(s)UsJ.$$ Then, for $`m`$ large enough, we can define the projection $`\mathrm{\Pi }_m:U𝒟\varphi ^1(a_{k_m})`$ $$\mathrm{\Pi }_m(x_m(s))=y_m(s)sJ.$$ (3.24) Note that, by the definition of $`\mathrm{\Pi }_m`$, $`\dot{y}_m(s),\varphi (y_m(s))=0`$, for any $`sJ`$, so since $$\dot{y}_m(s)=\eta _x(\tau _m(s),x_m(s))[\dot{x}_m(s)]\frac{\varphi (y_m(s))}{\varphi (y_m(s))^2}\dot{u}_m(s),$$ by assumption (iii) of Theorem 1.6 we get for any $`sJ`$ $`|\dot{y}_m(s)|^2=|\eta _x(\tau _m(s),x_m(s))[\dot{x}_m(s)]|^2{\displaystyle \frac{\dot{u}_m^2(s)}{\varphi (y_m(s))^2}}`$ $`C|\dot{x}_m(s)|^2.`$ (3.25) Hence by (3) and (iv) of Theorem 1.6 $`a^2\lambda _m(s_m)H_\varphi (x_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]H_\varphi (y_m(s_m))[\dot{y}_m(s_m)),\dot{y}_m(s_m)]+`$ $`M\dot{y}_m(s_m),\dot{y}_m(s_m)a_{k_m}.`$ (3.26) Now, following Remark 3.5, set for any $`x𝒟`$, $`v,w𝐑^N`$ $$L(x)[v,w]=\mathrm{d}^2\varphi (x)[v,w]\mathrm{d}\varphi (x)[\mathrm{d}Q(x)[v,w]],$$ (3.27) $`H_\varphi (x_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]H_\varphi (y_m(s_m))[\dot{y}_m(s_m),\dot{y}_m(s_m)]=`$ $`L(x_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]L(y_m(s_m))[\dot{y}_m(s_m),\dot{y}_m(s_m)]=`$ $`L(x_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]L(y_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]+`$ $`L(y_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]L(y_m(s_m))[\dot{y}_m(s_m),\dot{y}_m(s_m)]=`$ $`L(x_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]L(y_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]+`$ $`L(y_m(s_m))[\dot{x}_m(s_m)+\dot{y}_m(s_m),\dot{x}_m(s_m)\dot{y}_m(s_m)].`$ (3.28) In the following we shall denote by $`M_1,\mathrm{},M_7`$ suitable positive constants. Since $`\varphi 𝒞^3`$, $`\eta 𝒞^2`$, using the mean value theorem, the boundness of $`\left(x_m_{\mathrm{}}\right)_{m𝐍}`$ and $`\left(y_m_{\mathrm{}}\right)_{m𝐍}`$, (iii) of Theorem 1.6 and (3.16): $`L(x_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]L(y_m(s_m))[\dot{x}_m(s_m),\dot{x}_m(s_m)]`$ $`M_1|\dot{x}_m(s_m)|^2|x_m(s_m)y_m(s_m)|`$ $`M_2|\dot{x}_m(s_m)|^2\left(u_m(s_m)a_{k_m}\right)`$ $`M_3|\dot{x}_m(s_m)|^2.`$ (3.29) By the boundness of $`\left(y_m_{\mathrm{}}\right)_{m𝐍}`$ and (3) we get $`L(y_m(s_m))[\dot{x}_m(s_m)+\dot{y}_m(s_m),\dot{x}_m(s_m)\dot{y}_m(s_m)]`$ $`M_4|\dot{x}_m(s_m)|^2.`$ (3.30) Then by (3), (3), (3), (3), (3), we get $$a^2\lambda _m(s_m)M_5|\dot{x}_m(s_m)|^2$$ so that (3.12) implies $$\frac{ϵ_m}{\varphi ^3(x_m(s_m))}M_6+M_7\frac{ϵ_m}{\varphi ^2(x_m(s_m))}$$ from which the boundedness of the multiplier follows. $`\mathrm{}`$ ###### Proposition 3.7 Let $`\left(x_ϵ\right)_{ϵ>0}`$ be a sequence of critical points of $`f_ϵ`$ and $`k`$ a positive constant such that (3.15) holds. Then there exists a positive constant $`\beta `$ such that $$\varphi (x_ϵ(s))\beta s[0,1],ϵ]0,1].$$ (3.31) Proof: Assume by contradiction that there exist a decreasing and infinitesimal sequence $`\left(ϵ_m\right)_{m𝐍}]0,1]`$ and a sequence $`\left(x_{ϵ_m}\right)_{m𝐍}`$ of critical points of $`f_{ϵ_m}`$ satisfying (3.15) and such that $$\underset{s[0,1]}{\mathrm{min}}\varphi (x_{ϵ_m}(s))0\text{ as }m\mathrm{}.$$ (3.32) Defining $`x_m,u_m,s_m,s_0,U,J`$ and $`a_{k_m}`$ as in Lemma 3.6, since for $`sJ`$, $`m𝐍`$ sufficiently large $`y_m(s)U`$, we get, by (iv) of Theorem 1.6 $`\ddot{u}_m(s)H_\varphi (x_m(s))[\dot{x}_m(s),\dot{x}_m(s)]H_\varphi (y_m(s))[\dot{y}_m(s),\dot{y}_m(s)]+`$ $`M\varphi (y_m(s))\dot{y}_m(s),\dot{y}_m(s){\displaystyle \frac{2ϵ_m}{\varphi ^3(x_m(s))}}\varphi (x_m(s))^2`$ $`H_\varphi (x_m(s))[\dot{x}_m(s),\dot{x}_m(s)]H_\varphi (y_m(s))[\dot{y}_m(s),\dot{y}_m(s)]+`$ $`M\varphi (y_m(s))\dot{y}_m(s),\dot{y}_m(s)`$ (3.33) where $`y_m`$ is defined by (3.22) and (3.23). In the following we shall always assume $`sJ`$ and $`m𝐍`$ large enough and we shall denote by $`C_1,\mathrm{},C_{11}`$ suitable positive constants. Now by (3.27) $`H_\varphi (x_m(s))[\dot{x}_m(s),\dot{x}_m(s)]H_\varphi (y_m(s))[\dot{y}_m(s),\dot{y}_m(s)]=`$ $`L(x_m(s))[\dot{x}_m(s),\dot{x}_m(s)]L(y_m(s))[\dot{y}_m(s),\dot{y}_m(s)]=`$ $`L(x_m(s))[\dot{x}_m(s),\dot{x}_m(s)]L(y_m(s))[\dot{x}_m(s),\dot{x}_m(s)]+`$ $`L(y_m(s))[\dot{x}_m(s)+\dot{y}_m(s),\dot{x}_m(s)\dot{y}_m(s)].`$ (3.34) Since $`\varphi 𝒞^3`$, $`\eta 𝒞^2`$, using the mean value theorem, the boundness of $`\left(x_m_{\mathrm{}}\right)_{m𝐍}`$ and $`\left(y_m_{\mathrm{}}\right)_{m𝐍}`$ and (iii) of Theorem 1.6, there results $`L(x_m(s))[\dot{x}_m(s),\dot{x}_m(s)]L(y_m(s))[\dot{x}_m(s),\dot{x}_m(s)]`$ $`C_1|\dot{x}_m(s)|^2u_m(s).`$ (3.35) Since $`\eta _x(0,x)[v]=v`$, $`\dot{x}_m(s)\dot{y}_m(s)=\eta _x(0,x_m(s))[\dot{x}_m(s)]`$ (3.36) $`\eta _x(\tau _m(s),x_m(s))[\dot{x}_m(s)]\eta _s(\tau _m(s),x_m(s))\dot{u}_m(s),`$ so by (3.36) the mean value theorem and (iii) of Theorem 1.6 $`L(y_m(s))[\dot{x}_m(s)+\dot{y}_m(s),\dot{x}_m(s)\dot{y}_m(s)]=`$ $`L(y_m(s))[\dot{x}_m(s)+\dot{y}_m(s),\eta _x(0,x_m(s))[\dot{x}_m(s)]\eta _x(\tau _m(s),x_m(s))[\dot{x}_m(s)]]`$ $`L(y_m(s))[\dot{x}_m(s)+\dot{y}_m(s),\eta _s(\tau _m(s),x_m(s))\dot{u}_m(s)]`$ $`C_2|\dot{x}_m(s)|^2u_m(s)+B_m(s)\dot{u}_m(s)`$ (3.37) where $$B_m(s)=L(y_m(s))[\dot{x}_m(s)+\dot{y}_m(s),\eta _s(\tau _m(s),x_m(s))]$$ (3.38) is a $`𝒞^1`$ function. By (3.12), (3.19), Lemma 3.6 and (ii) of Theorem 1.6 we get $$\dot{x}_m(s),\dot{x}_m(s)C_3$$ (3.39) for a suitable constant $`C_3`$, hence by (3), (3), (3) and (3) we get $$\ddot{u}_m(s)C_4u_m(s)+B_m(s)\dot{u}_m(s).$$ (3.40) Note that by (3.39) and (iii) of Theorem 1.6 $$|B_m(s)|C_5$$ (3.41) and by (3.4), (ii) of Theorem 1.6 $$|D_s\dot{x}_m(s)|C_6.$$ Since $`\ddot{x}_m(s)=P(x_m(s))\ddot{x}_m(s)+Q(x_m(s))\ddot{x}_m(s)`$ $`=D_s\dot{x}_m(s)dQ(x_m(s))[\dot{x}_m(s),\dot{x}_m(s)]`$ where $`P,Q`$ are as in (3.13), (3.14), we also get $$|\ddot{x}_m|C_7.$$ Then by using (iii) of Theorem 1.6, standard arguments show that $$|\dot{B}_m(s)|C_8.$$ (3.42) Therefore by (3.40), (3.41) and (3.42) integrating by parts, for $`s>s_m`$ there results $`\dot{u}_m(s)={\displaystyle _{s_m}^s}\ddot{u}_m(\tau )𝑑\tau C_4{\displaystyle _{s_m}^s}u_m(\tau )𝑑\tau +{\displaystyle _{s_m}^s}B_m(\tau )\dot{u}_m(\tau )𝑑\tau =`$ $`C_4{\displaystyle _{s_m}^s}u_m(\tau )𝑑\tau +B_m(s)u_m(s)B_m(s_m)u_m(s_m){\displaystyle _{s_m}^s}\dot{B}_m(\tau )u_m(\tau )𝑑\tau `$ $`C_9{\displaystyle _{s_m}^s}u_m(\tau )𝑑\tau +C_5u_m(s)+C_5u_m(s_m).`$ (3.43) Hence $`u_m(s)u_m(s_m)+C_9{\displaystyle _{s_m}^s}\left({\displaystyle _{s_m}^\tau }u_m(r)𝑑r\right)𝑑\tau +C_5{\displaystyle _{s_m}^s}u_m(\tau )𝑑\tau +C_5u_m(s_m)`$ $`C_{10}u_m(s_m)+C_{11}{\displaystyle _{s_m}^s}u_m(\tau )𝑑\tau .`$ (3.44) By the Gronwall lemma we get $$u_m(s)C_{10}u_m(s_m)\mathrm{exp}(C_{11}(ss_m))$$ so by (3.32) we have $$u_m0\text{ uniformly}$$ getting a contradiction. $`\mathrm{}`$ Proof of Theorem 1.6. By Remark 3.4 and Proposition 3.7, we can find a family $`\left(x_ϵ\right)_{ϵ>0}`$ of critical points of $`f_ϵ`$ such that (3.31) holds. By (3.15) and the Poincaré inequality, $`\left(x_ϵ\right)_{ϵ>0}`$ is bounded in $`H^{1,2}([0,1],𝐑^N)`$, so there is a subsequence $`\left(x_{ϵ_m}\right)_{m𝐍}`$ such that $$x_{ϵ_m}x\text{ weakly in }H^{1,2}([0,1],𝐑^N),$$ (3.45) where $`x\mathrm{\Omega }^1\left(𝒟\right)`$ since the convergence is also uniform and (3.31) holds. Now it is easy to prove that $`f`$ attains a minimum value at $`x`$. Indeed by (3.45) and (3.3) $$f(x)\underset{m\mathrm{}}{lim\; inf}f_{ϵ_m}(x_{ϵ_m})f(y)$$ for any $`y\mathrm{\Omega }^1\left(𝒟\right)`$, so $`𝒟`$ is convex. Finally, if $`𝒟`$ is not contractible in itself, the proof can be carried out exactly as the one of Theorem 0.2 of . $`\mathrm{}`$ Proof of Theorem 1.8. The only difference with the proof of Theorem 1.6 concerns the a priori estimates of Proposition 3.7 that here are simpler because they do not require a projection. Indeed, defining $`x_m`$, $`u_m`$, $`s_m`$, $`J`$ as in Proposition 3.7, we get by (ii) and (iii) of Theorem 1.8, for $`m`$ sufficiently large $$\ddot{u}_m(s)M\varphi (x_m(s))\dot{x}_m(s),\dot{x}_m(s)\frac{2ϵ_ma^2}{\varphi ^3(x_m(s))}.$$ As (3.12) holds, we get $$\ddot{u}_m(s)M_1u_m(s)+M_2\frac{ϵ_m}{u_m(s)}M_3\frac{ϵ_m}{u_m^3(s)}$$ for $`M_1,M_2,M_3>0`$, so that, if $`\delta `$ is sufficiently small $$\ddot{u}_m(s)M_1u_m(s).$$ Then, by the Gronwall Lemma, we immediately get a contradiction. $`\mathrm{}`$ ## 4 Applications (a) Some examples. (1) First, we will check that Theorems 1.6, and 1.8 can be applied in cases where neither the elementary considerations for differentiable boundary nor Theorem 1.5 are appliable. Let $`(,,)`$ be a cylinder $`C`$ in $`𝐑^3`$ and let $`H`$ be an helix in $`C`$. Set $`𝒟=CH`$ and take $`\varphi `$ equal to the distance to $`H`$ on a strip $`S`$ around $`H`$. Recall that $`𝒟`$ cannot be considered as a manifold with boundary $`𝒟=H.`$ Clearly, Theorems 1.6 and 1.8 are applicable and, moreover, we can perturb the metric to make the constant $`M`$ in (1.9) positive (for example, by conformally changing the metric symmetrically around $`H`$ on $`S`$, see formula (4.3)); thus, Theorem 1.5 may be not applicable now. (2) Nevertheless, the following example shows that Theorem 1.5 may be applicable when Theorems 1.6, 1.8 are not, even if the condition for the Hessian in Theorem 1.8 is satisfied. Consider $`𝒟=]0,\mathrm{}[\times ]0,\mathrm{}[`$, and $`=𝐑^2`$ equipped with the Euclidean metric (this metric can be perturbed as suggested in last example to make this one non–trivial; note that conditions (ii), (iii) in Theorem 1.6 are independent of the metric on $``$). If $`\varphi (x,y)=\sqrt{xy}`$, then $$\varphi (x,y)^2=\frac{x^2+y^2}{4xy}>\frac{1}{4},$$ so the first inequality in (ii) of Theorems 1.6, 1.8 is satisfied, but not the second one. On the other hand, if we choose $`\widehat{\varphi }(x,y)=xy`$ then $$\widehat{\varphi }(x,y)^2=x^2+y^2,$$ which satisfies the upper local bound around each point, but not the lower one (compare with Remark 1.7(2)). Nevertheless, it is straightforward to check that Theorem 1.5 is applicable for $`\varphi `$ as well as for $`\widehat{\varphi }`$. (3) It is clear that condition (iii) of Theorem 1.8 on all tangent vectors $`v`$ may not hold under the corresponding hypothesis (iv) of Theorem 1.6. The following example shows that Theorem 1.8 may be appliable when Theorem 1.6 is not, because of the hypothesis (iii) of this theorem and the requirement that $`𝒟`$ must be a subset of a complete manifold. Consider $`𝒟==𝐑\{0\}\times 𝐑`$ endowed with the Riemannian metric given in polar coordinates $`r,\theta `$ by $$dx^2=dr^2+\frac{1}{r^2}d\theta ^2.$$ Choosen $`\varphi (r,\theta )=r`$, it follows $`\varphi =_r`$, so $`\varphi (r,\theta )=1`$. Standard calculations show that the (normalized) flow of $`\varphi `$ is $`\eta (s,(r,\theta ))=(rs,\theta )`$ and the norm of the partial derivative $`\eta _{(r,\theta )}`$ is not bounded, so we cannot apply Theorem 1.6. Moreover, note that $``$ is not complete and its curvature along incomplete radial geodesics diverges, so $`𝒟`$ is not isometric to a domain of a complete Riemannian manifold ($`_c𝒟`$ is topologically a circumference). However, it is easy to check that Theorem 1.8 is applicable. (b) Trajectories of Lagrangian systems. The interest in the study of the geodesic connectedness of a complete Riemannian manifold is related to the existence of trajectories of a Lagrangian system joining two fixed points. More precisely, consider a potential $`VC^2(,𝐑)`$ bounded from above in a domain $`𝒟`$ and the system $$\{\begin{array}{cc}D_s\dot{x}=V(x)\hfill & \\ x(0)=p,x(1)=q.\hfill & \end{array}$$ (4.1) Each solution $`x:[0,1]𝒟`$ of (4.1) has constant energy, that is, there exists $`E𝐑`$ such that, for any $`s[0,1]`$ $$\frac{1}{2}\dot{x},\dot{x}+V(x)=E.$$ We fix $`E>sup_𝒟V`$ and consider the Jacobi metric $$,_E=(EV(x)),\text{on }\overline{𝒟}.$$ (4.2) As $`\overline{𝒟}`$ is complete for $`,`$ then it is also complete for $`,_E`$ and, thus, we can extend $`,_E`$ to a complete Riemannian metric on all $``$. It is well–known that the geodesics on $`𝒟`$ with respect to the Riemannian metric $`,_E`$ are, up to reparametrizations, solutions of (4.1) with energy $`E`$ and vice–versa. Let $`\varphi `$ be a function satisfying conditions (i), (ii), (iii) in Theorem 1.6, which are independent of the metric. If assumption (iv) is satisfied with respect to $`,_E`$ then the existence of at least one solution of (4.1) with energy $`E`$ is proved. Notice that the Hessian of $`\varphi `$ with respect the two metrics is linked by the following relation $$H_\varphi ^E(x)[v,v]=H_\varphi (x)[v,v]+\varphi (x),u(x)v,v2u(x),v\varphi (x),v$$ (4.3) for any $`x`$, $`vT_x`$ where $$u(x)=\frac{1}{2}\mathrm{log}(EV(x)).$$ Thus, if (iv) of Theorem 1.6 is verified with respect to $`,`$, then it is satisfied by $`,_E`$ when for each $`y𝒟`$ there exists a neighborhood $`U`$ and a (positive) constant $`M^{}𝐑`$ such that $$\varphi (x),V(x)M^{}\varphi (x)$$ (4.4) $`x\varphi ^1(a_m)U`$. Summing up the following result holds. ###### Corollary 4.1 Let $`(,,)`$ be a complete Riemannian manifold and assume that there exists a positive and differentiable function $`\varphi `$ on $`𝒟`$ satisfying (i)–(iv) of Theorem 1.6. Then, if $`VC^2(,𝐑)`$ is bounded from above on $`𝒟`$ and (4.4) holds, for any $`E>sup_𝒟V`$ there exists a solution with energy $`E`$ of (4.1). Moreover, if $`𝒟`$ is non contractible in itself, then for any $`E>supV`$ there exist infinitely many solutions of (4.1) with energy $`E`$. Acknowledgment. We wish to thank Domingo Rodríguez for his kind help for the figures of this paper.
warning/0004/nlin0004017.html
ar5iv
text
# On some nondecaying potentials and related Jost solutions for the heat conduction equation Work supported in part by PRIN 97 “Sintesi” ## 1 Introduction In this article we investigate into the direct and inverse scattering transform for the heat conduction operator $$L=_{x_2}+_{x_1}^2u(x_1,x_2),$$ (1.1) in the case in which $`u`$ is a real function with “ray” type behavior. More exactly, $`u`$ is supposed to be rapidly decaying in all directions on the $`x`$-plane with the exception of some finite number of directions, where it has finite and nontrivial limits, i.e. $$u_{n,\pm }(x_1)=\underset{x_2\pm \mathrm{}}{lim}u(x_12\mu _nx_2,x_2),n=1,2,\mathrm{},N,$$ (1.2) for $`N`$ real constants $`\mu _n`$. The spectral theory of operator $`L`$ with potential in this class is interesting per se and because it is associated to the Kadomtsev–Petviashvili equation in its version called KPII $$(u_t6uu_{x_1}+u_{x_1x_1x_1})_{x_1}=3u_{x_2x_2}.$$ (1.3) In investigating KPII equation it was found that there are significant differences with respect to the case of KPI (equation (1.3) with opposite sign in the rhs). Specifically, the inverse problem cannot be formulated as a Riemann–Hilbert boundary value problem . This results from the fact that for the associated spectral problem there exist eigenfunctions which, though bounded, are nowhere analytic in the spectral parameter. The role of the Riemann–Hilbert problem is now played by what is usually called a $`\overline{}`$-problem. The basic idea is to compute the $`\overline{}`$-derivative of these eigenfunctions (modified Jost solutions) and then to exploit the Cauchy–Green formula to get a linear Fredholm integral equation for such solutions (under the assumption that the homogeneous integral equation has no nontrivial solutions). One can prove the smoothness of the potential $`u(x,t)`$ and also some other properties . Anyway, it is always assumed that the potential is rapidly decaying for $`x_1^2+x_2^2\mathrm{}`$ (which excludes in particular the presence of line solitons) and a rigorous investigation of the IST and of the properties of solutions of KPII equation with nontrivial (for instance, one-dimensional) asymptotic behavior is still missing. Such extension of the spectral theory for the heat operator would provide the possibility to extend correspondingly the class of solutions of the KPII equation. The same problem has been faced for the nonstationary Schrödinger equation in the framework of the so-called resolvent approach . This extension in the general case resulted to be particularly involved, in particular as far as the definition of spectral data and their characterization equations are concerned. Some general results were obtained in with some (partially implicit) assumptions. Even though it was not possible to define precisely the subclass of potentials satisfing these assumptions, it was shown that potentials belonging to this special subclass can be obtained via Bäcklund transformations (BT’s). For getting some experience with a case less involved from a mathematical point of view but sufficiently general, it was considered the case of potentials obtained by applying recursively binary BT’s to an arbitrary decaying potential and describing $`N`$ solitons superimposed to a generic background . At least in the case $`N=1`$ it was possible to obtain explicitly the resolvent, the Jost solutions and the spectral data and to show that they satisfy all the required assumptions . Moreover, there is evidence that the structural properties of the spectral data can be even more complicated in the general case. Spectral theory of the operator (1.1) with ray potential is essentially more involved than the standard case of rapidly decaying potential. One of the problems is that the standard integral equations defining the Jost solution and its dual become senseless, since the Green’s function is slowly decaying at space infinity and cannot ensure convergency. In Sec. 2, after reviewing some properties of Green’s function and Jost solutions for the heat conduction equation in the case of rapidly decaying potentials, we propose an extension of these integral equations which allows also to include potentials with ray type behavior and which in the one-dimensional limit give the standard integral equations for the Jost solutions of KdV equation. We call the solutions of these modified integral equations Jost-like solutions. In Sec. 3 we consider the potentials obtained by superimposing, via binary Bäcklund transformations, one soliton to a generic decaying background. A significant difference with respect to the case of the nonstationary Schrödinger operator is due to the fact that in order to get a real potential we are obliged to use as functions generating the BT’s the Jost solution and dual Jost solution computed for purely imaginary values of the spectral parameter and thus an accurate study of their properties at these values is required. We get the explicit expression of the potential and we are able to formulate the conditions to be imposed on the parameters of the Bäcklund transformation in order to obtain a real and regular solution. Moreover, we obtain also the explicit expression of the Jost-like solutions and study their spectral properties. Finally, in Sec. 4 we iterate this procedure to get two solitons on a generic background. Also in this case we determine the regularity conditions for the solutions obtained by these means and we explicitly compute Jost solutions and study their spectral properties. An interesting feature of these solutions describing two solitons on a background is that, unlike KPI case, the directions of the solitons are not fixed by the parameters of the Bäcklund transformations but depend also on the “constants” of integration of the Darboux procedure. If the matrix which appears in the expression of the potential is diagonal, the two BT’s simply “superimpose” the two solitons with only a phase shift. But in the general case, that is when both off-diagonal entries of such matrix are different from zero, the solitons are rotated with respect to the directions determined by the single BT’s. When one of the off-diagonal entries is null but the other one is not, only one soliton is rotated while the other one is unchanged and we could say that the “interaction” does not preserve the soliton direction. If a row or a column in matrix $`C_2`$ is made up of zeros the resulting potential has three “rays”. Note that all these features are absent in the case of KPI equation. In fact, for KPI the soliton directions do not depend on the choice of such matrix . Another difference with respect to the case of KPI equation is that the KPI wave soliton emerging from the background at large distances divides the $`x`$-plane into two regions with different asymptotic behavior. In the first region (to the left of the soliton) the original potential $`u(x)`$ is modified at large distances by an exponentially decreasing term, while in the second region (to the right) by a term decreasing as $`1/x_2^3`$. In the case of KPII the corrections are always decaying at infinity faster that any power of $`x_2`$. ## 2 On Jost solutions In the standard case one introduces the complex spectral parameter $`k=k_{\mathrm{}}+ik_{\mathrm{}}`$ and defines the Jost solution $`\mathrm{\Phi }(x,k)`$ and dual Jost solution $`\mathrm{\Psi }(x,k)`$ as the solutions of the heat conduction equation and its dual, respectively, such that $`\chi (x,k)`$ $`=`$ $`e^{ikx_1+k^2x_2}\mathrm{\Phi }(x,k)`$ (2.4) $`\xi (x,k)`$ $`=`$ $`e^{ikx_1k^2x_2}\mathrm{\Psi }(x,k)`$ (2.5) satisfy the integral equations $`\chi (x,k)`$ $`=`$ $`1+{\displaystyle 𝑑x^{}G_0(xx^{},k)u(x^{})\chi (x^{},k)}`$ (2.6) $`\xi (x,k)`$ $`=`$ $`1+{\displaystyle 𝑑x^{}G_0(x^{}x,k)u(x^{})\xi (x^{},k)}`$ (2.7) where the Green’s function $`G_0(x,k)`$ is given by $$G_0(x,k)=\frac{signx_2}{2\pi }𝑑\alpha \theta (\alpha (\alpha +2k_{\mathrm{}})x_2)e^{i\alpha x_1\alpha (\alpha +2k)x_2}.$$ (2.8) It is evident from expression (2.8) that $`G_0`$ and, consequently, $`\chi `$ and $`\xi `$ are nowhere analytic in the complex $`k`$-plane. Functions $`\chi `$ and $`\xi `$ are bounded for all $`x_1,x_2`$ (see for instance ) and $$\underset{k\mathrm{}}{lim}\chi (x,k)=\underset{k\mathrm{}}{lim}\xi (x,k)=1.$$ The Jost solutions defined by (2.4) and (2.5) obey the following normalization and completness conditions $`{\displaystyle 𝑑x_1\mathrm{\Psi }(x_1,x_2,k+p)\mathrm{\Phi }(x_1,x_2,k)}=2\pi \delta (p)`$ (2.9) $`{\displaystyle 𝑑k_{\mathrm{}}\mathrm{\Psi }(x_1^{},x_2,k)\mathrm{\Phi }(x_1,x_2,k)}=2\pi \delta (x_1x_1^{})`$ (2.10) One can check that $$G_0(x,k)=\frac{signk_{\mathrm{}}}{2i\pi \left|x_1\right|}\left[e^{2ik_{\mathrm{}}(x_1+2k_{\mathrm{}}x_2)}1\right]+o(\left|x_1\right|^1)$$ (2.11) and $$G_0(x,k)=\frac{signk_{\mathrm{}}}{4\pi kx_2}+o(\left|x_2\right|^1).$$ (2.12) Despite these asymptotics are discountinuous at $`k_{\mathrm{}}=0`$, the Green’s function is not and one can easily check from (2.8) that $$G_0(x,ik_{\mathrm{}})=\frac{\theta (x_2)}{2\sqrt{\pi x_2}}e^{\frac{(x_1+2k_{\mathrm{}}x_2)^2}{4x_2}}$$ which is real and, up to the exponential factor $`e^{k_{\mathrm{}}x_1k_{\mathrm{}}^2x_2},`$ is the well-known Gauss-Weierstrass kernel (or heat kernel). Then $`\chi `$ and $`\xi `$ can be computed at purely imaginary values of the spectral parameter $`k`$ and they are real as well. Solvability of the integral equations (2.6)–(2.7) under some small norm assumptions was proved in and thanks to (2.11) it is easy to show that $`\chi (x,k)`$ and $`\xi (x,k)`$ have the asymptotic behavior $`\underset{\left|x_1\right|\mathrm{}}{lim}\chi (x,k)`$ $`=`$ $`1`$ (2.13) $`\underset{\left|x_1\right|\mathrm{}}{lim}\xi (x,k)`$ $`=`$ $`1`$ (2.14) Moreover, one can prove the following propositions. ###### Proposition 1 Let $`u`$ be a real, regular and rapidly decaying potential. The modified Jost solutions $`\chi `$ and $`\xi `$ obey for any $`n`$ the following asymptotics $`\underset{x_2\pm \mathrm{}}{lim}x_2^n\left(\chi (x,ik_{\mathrm{}})1\right)`$ $`=`$ $`0,`$ (2.15) $`\underset{x_2\pm \mathrm{}}{lim}x_2^n\left(\xi (x,ik_{\mathrm{}})1\right)`$ $`=`$ $`0.`$ (2.16) Proof. Let us consider the integral equation (2.6) for $`\chi `$. We have for $`k_{\mathrm{}}=\kappa `$ $$\chi (x,i\kappa )1=\frac{1}{2\sqrt{\pi }}𝑑x^{}\frac{\theta (x_2x_2^{})}{\sqrt{x_2x_2^{}}}e^{\frac{\left(x_1x_1^{}+2\kappa (x_2x_2^{})\right)^2}{4(x_2x_2^{})}}u(x^{})\chi (x^{},i\kappa )$$ and we can make shift and proper change of variables to get $$\chi (x,i\kappa )1=\frac{1}{\sqrt{\pi }}𝑑x^{}\theta (x_2^{})e^{(x_1^{})^2}(u\chi )(x_12x_1^{}\left|x_2^{}\right|^{1/2}+2\kappa \left|x_2^{}\right|,x_2\left|x_2^{}\right|,i\kappa ).$$ Note that $`\left|{\displaystyle 𝑑x_1^{}_0^+\mathrm{}𝑑x_2^{}e^{(x_1^{})^2}(u\chi )(x_12x_1^{}\left|x_2^{}\right|^{1/2}+2\kappa \left|x_2^{}\right|,x_2\left|x_2^{}\right|,i\kappa )}\right|`$ $``$ $`{\displaystyle _{\mathrm{}}^+\mathrm{}}𝑑x^{}e^{(x_1^{})^2}\left|(u\chi )(x_12x_1^{}\left|x_2^{}\right|^{1/2}+2\kappa \left|x_2^{}\right|,x_2\left|x_2^{}\right|,i\kappa )\right|.`$ Since $`\chi `$ is bounded and $`u`$ is in the Schwartz class the limit $`\left|x_2\right|\mathrm{}`$ can be exchanged with the integral and the integrand decays as fast as $`u`$, that is faster than any power of $`x_2`$. The same result holds for $`\xi `$. ###### Proposition 2 Let $`u`$ be a real potential in the Schwartz class. The modified Jost solutions $`\chi `$ and $`\xi `$ for the heat conduction equation with potential $`u`$ are such that $`\chi (x,ik_{\mathrm{}})0,\xi (x,ik_{\mathrm{}})0`$ $`x^2,`$ or equivalently, taking into account asymptotics (2.15)–(2.16), $`\chi (x,ik_{\mathrm{}})`$ $`>`$ $`0`$ (2.17) $`\xi (x,ik_{\mathrm{}})`$ $`>`$ $`0,x^2.`$ (2.18) Proof. In order to simplify the notation in the following we will put $`k_{\mathrm{}}=\kappa `$. Let us write the solution of the integral equation (2.6) as the (formal) Neumann series $$\chi (x,i\kappa )=\underset{n=0}{\overset{\mathrm{}}{}}\chi _n(x,i\kappa )$$ where $`\chi _0(x,i\kappa )`$ $`=`$ $`1`$ $`\chi _n(x,i\kappa )`$ $`=`$ $`{\displaystyle \frac{1}{2\sqrt{\pi }}}{\displaystyle 𝑑x^{}\frac{\theta (x_2x_2^{})}{\sqrt{x_2x_2^{}}}e^{\frac{\left[\left(x_1x_1^{}\right)+2\kappa (x_2x_2^{})\right]^2}{4(x_2x_2^{})}}u\left(x^{}\right)\chi _{n1}(x^{},i\kappa )}`$ Obviously we have to consider the class of potentials for which the existence of solution of the integral equation, that is the convergency of the Neumann series, is estabilished (see and the following remark). Now let us assume that for all $`x^2`$ $$\left|u(x)\right|<U(x_2)$$ with $`U`$ such that $$_0^+\mathrm{}𝑑x_2^{}\left|U(x_2x_2^{})\right|<1.$$ If we introduce $$a(x)=𝑑x^{}\frac{\theta (x_2x_2^{})}{\sqrt{x_2x_2^{}}}e^{\frac{\left[\left(x_1x_1^{}\right)+2\kappa (x_2x_2^{})\right]^2}{4(x_2x_2^{})}}\left|u(x^{})\right|,$$ (2.19) due to (2)–(2.19) we have $$0<\frac{a(x)}{2\sqrt{\pi }}<\frac{1}{2}.$$ (2.20) Then by induction on $`n`$ one can easily prove that $$\left|\chi _{n+1}(x,i\kappa )\right|<\left(\frac{1}{2}\right)^{n+1},$$ so that $$\left|\chi (x,i\kappa )1\right|<1.$$ The main problem when dealing with a potential $`u(x)`$ not vanishing in all directions at large distances is that the integral equations (2.6) and (2.7) cannot be applied as the Green’s function is slowly decaying at space infinity. We suggest the following modification of these integral equations: $`\chi (x,k)`$ $`=`$ $`1+{\displaystyle \underset{k_{\mathrm{}}\mathrm{}}{\overset{x_1}{}}}𝑑y_1{\displaystyle 𝑑x^{}_{y_1}G_0(y_1x_1^{},x_2x_2^{},k)u(x^{})\chi (x^{},k)},`$ (2.21) $`\xi (x,k)`$ $`=`$ $`1+{\displaystyle \underset{k_{\mathrm{}}\mathrm{}}{\overset{x_1}{}}}𝑑y_1{\displaystyle 𝑑x^{}_{y_1}G_0(x_1^{}y_1,x_2^{}x_2,k)u(x^{})\xi (x^{},k)},`$ (2.22) where the order of operations is explicitly prescribed. Here and below we use notations of the type $`k_{\mathrm{}}\mathrm{}`$ in the limits of integrals to indicate the sign of infinity. If the solution of this equation exists and is bounded, then like in the standard case $`\underset{x_1k_{\mathrm{}}\mathrm{}}{lim}\chi (x,k)`$ $`=`$ $`1,`$ (2.23) $`\underset{x_1k_{\mathrm{}}\mathrm{}}{lim}\xi (x,k)`$ $`=`$ $`1,k_{\mathrm{}}0,`$ (2.24) while in contrast to (2.13) and (2.14) they can be different from 1 in the opposite direction. These modified integral equations are applicable to the simplest case of a potential of type (1.2), i.e. to the case $`u(x)=u(x_1)`$ and one recovers the standard one-dimensional integral equation for Jost solution of KdV equation. Since the general case is extremely involved, we are studying here the special but rather wide subclass of potentials of type (1.2) that is obtained by applying recursively the so-called binary Bäcklund transformations with complex spectral parameter to a decaying potential. As we are interested in the spectral properties of potentials $`u`$ having nontrivial limits, following an approach close to the one in , we study also the corresponding Darboux transformations furnishing the Jost solutions of the transformed potentials and their analytical properties, as well as transformations of the spectral data. ## 3 Binary Bäcklund transformations We have at our disposal in and a rather simple and transparent method for performing binary Bäcklund transformations of the potential $`u`$ and corresponding Darboux transformations of solutions of (1.1) and its dual. Let $`u_0`$ be a smooth, real and rapidly decaying potential and $`L_0`$ the corresponding heat operator $$L_0=_{x_2}+_{x_1}^2u_0(x).$$ (3.25) A new potential $`u_0^{}(x)`$ can be generated through an elementary BT $$L_0^{}B_0=B_0L_0$$ (3.26) using a gauge operator $$B_0=_{x_1}\left(_{x_1}\mathrm{log}\phi _0(x)\right).$$ (3.27) If we impose (3.26) we find that $$u_0^{}(x)=u_0(x)2_{x_1}^2\mathrm{log}\phi _0(x)$$ (3.28) and that $`\phi _0`$ has to be a solution of the original spectral problem, that is $$\left(_{x_2}+_{x_1}^2u_0(x)\right)\phi _0(x)=0.$$ (3.29) Let us consider also the spectral operator dual to (3.25), that is $$L_0^d=_{x_2}+_{x_1^2}u_0(x)$$ (3.30) and the gauge operator $$B_0^d=_{x_1}+\left(_{x_1}\mathrm{log}\phi _0(x)\right).$$ (3.31) Eq. (3.26) is equivalent to the corresponding dual relation $$B_0^dL_0^d=L_0^dB_0^d.$$ (3.32) Moreover, given any two solutions of the spectral problem (3.25) and of its dual (3.30), say $`\varphi _0`$ and $`\psi _0`$, their Darboux transforms $`\varphi _0^{}`$ and $`\psi _0^{}`$, defined by $`\varphi _0^{}=B_0\varphi _0`$ (3.33) $`B_0^d\psi _0^{}=\psi _0`$ (3.34) solve, respectively, the spectral problem for the transformed potential and its dual up to annihilators of $`B_0^d`$, that is $`L_0^{}\varphi _0^{}=0`$ and $`B_0^dL_0^{}\psi _0^{}=0`$. One can easily check that if $`f`$ and $`g`$ solve (3.29) and its dual, respectively, then $$_{x_2}\left(f(x)g(x)\right)=_{x_1}W(f(x),g(x))$$ (3.35) where $`W`$ is standard Wronskian of $`f`$ and $`g`$, that is $$W(f(x),g(x))=f(x)_{x_1}g(x)g(x)_{x_1}f(x).$$ (3.36) Now let us consider an inverse elementary BT $$L_0^{}B_0^{}=B_0^{}L_1$$ (3.37) through the gauge $$B_0^{}=_{x_1}+\left(_{x_1}\mathrm{log}\psi _0^{}(x)\right).$$ (3.38) In this case we have $$u_1(x)=u_0^{}(x)2_{x_1}^2\mathrm{log}\psi _0^{}$$ (3.39) where $`\psi _0^{}`$ is an arbitrary solution of $$\left(_{x_2}+_{x_1}^2u_0^{}(x)\right)\psi _0^{}(x)=0.$$ (3.40) Now we perform direct and inverse elementary Bäcklund transformations (3.26) and (3.32) and study the properties of operator $`L_1`$, omitting all intermediate constructions associated to operator $`L_0^{}`$. First of all, from (3.28) and (3.39) we see that $$u_1(x)=u_0(x)2_{x_1}^2\mathrm{log}\mathrm{\Delta }_1(x)$$ (3.41) where we introduced $$\mathrm{\Delta }_1(x)=\phi _0(x)\psi _0^{}(x).$$ (3.42) Moreover, using (3.28), from (3.29) and (3.40) we find that $`\phi _0`$ and $`\psi _0^{}`$ have to be chosen in such a way that they satisfy the following system of equations $$\{\begin{array}{c}(_{x_2}+_{x_1}^2u_0(x))\phi _0(x)=0\hfill \\ \left(_{x_2}+_{x_1}^22\left(_{x_1}\mathrm{log}\phi _0(x)\right)_{x_1}\right)\mathrm{\Delta }_1(x)=0\hfill \end{array}.$$ (3.43) It is clear that a sufficient condition for $`u_1`$ in (3.41) to be real is to choose $`\mathrm{\Delta }_1`$ real, but this means that $`\phi _0`$ cannot be an arbitrary solution of the original spectral problem. In particular, one can choose a real solution such as the Jost solution computed for purely imaginary value of spectral parameter, that is $$\phi _0(x)=\mathrm{\Phi }_0(x,i\kappa _1).$$ (3.44) In order to obtain $`\psi _0^{}(x)`$, and consequently $`\mathrm{\Delta }_1`$, we consider the Darboux transform (3.34) of a solution $`\psi _0(x)`$ of the spectral problem $`L_0^d\psi _0=0`$ and we choose $$\psi _0(x)=\mathrm{\Psi }_0(x,i\alpha _1)$$ (3.45) where $`\mathrm{\Psi }_0`$ is the Jost solution of the spectral problem (3.30) computed for $`k=i\alpha _1,`$ $`\alpha _1`$. With this choice we get $$\psi _0^{}(x)=\frac{1}{\phi _0(x)}\left[C(x_2)+_{(\kappa _1\alpha _1)\mathrm{}}^{x_1}𝑑x_1^{}\mathrm{\Psi }_0(x_1^{},x_2,i\alpha _1)\mathrm{\Phi }_0(x_1^{},x_2,i\kappa _1)\right]$$ (3.46) and the integral is convergent due to the asymptotic behaviors (2.13) and (2.14). “Constant” of integration $`C`$ depends, in general, on $`x_2`$ as well, but one can check that $`\psi _0^{}`$ is indeed a solution of (3.40) iff $`C`$ does not depend on $`x_2`$ . So finally we get $$\mathrm{\Delta }_1(x)=c_1+_{(\kappa _1\alpha _1)\mathrm{}}^{x_1}𝑑x_1^{}\mathrm{\Phi }_0(x_1^{},x_2,i\kappa _1)\mathrm{\Psi }_0(x_1^{},x_2,i\alpha _1),$$ (3.47) where $`c_1`$ and $`\alpha _1,\kappa _1`$ are real constants with $`\alpha _1\kappa _1`$. Taking into account (3.47), it is always possible to formulate a small norm condition which ensures regularity of the dressed potential (3.41). Indeed, Prop. 2 proves that $`\mathrm{\Phi }_0(x,ik_{\mathrm{}})`$ and $`\mathrm{\Psi }_0(x,ik_{\mathrm{}})`$ are real and strictly positive for all $`x^2`$; then the integral in (3.47) is a monotonic function of $`x_1`$ and if we choose $`c_1`$ such that $$(\kappa _1\alpha _1)c_10,$$ (3.48) eq. (3.47) gives a function $`\mathrm{\Delta }_1`$ which is regular, has no zeros in the $`x`$-plane and has the same sign as $`\kappa _1\alpha _1`$. Consequently, condition (3.48) ensures regularity of potential (3.41). Finally, one can easily show that $`\mathrm{\Delta }_1`$ indeed superimposes one soliton on the generic (smooth and rapidly decaying) background $`u_0`$ along the direction $`x_1+(\kappa _1+\alpha _1)x_2=\mathrm{c}onst`$. ### 3.1 Darboux procedure We need now to express the Jost solutions $`\mathrm{\Phi }_1`$ and $`\mathrm{\Psi }_1`$ of the spectral equations $$L_1\mathrm{\Phi }_1=0,L_1^d\mathrm{\Psi }_1=0$$ (3.49) in terms of the Jost solutions $`\mathrm{\Phi }_0`$ and $`\mathrm{\Psi }_0`$ of the original spectral problem, that is we have to construct the Darboux version of the binary BT. From (3.37) we see that if $`\varphi _0^{}`$ satisfies $`L_0^{}\varphi _0^{}=0`$ and we define $`\varphi _1`$ through $$B_0^{}\varphi _1=\varphi _0^{},$$ (3.50) then $`\varphi _1`$ is such that $`B_0^{}L_1\varphi _1=0`$. So, taking into account definition of $`B_0^{}`$ in (3.38), from (3.50) we get $$_{x_1}\left(\varphi _1(x,k)\psi _0^{}(x)\right)=\varphi _0^{}(x,k)\psi _0^{}(x).$$ (3.51) In order to have transformations parametrized by constants and not by functions of $`x_2`$ obeying some differential equation, it would be convenient to integrate this equation from infinity to $`x_1.`$ From the other side, it is natural to use (3.33) to get $`\varphi _0^{}`$, that is $$\varphi _0^{}(x,k)=_{x_1}\varphi _0(x,k)\frac{_{x_1}\mathrm{\Phi }_0(x,i\kappa _1)}{\mathrm{\Phi }_0(x,i\kappa _1)}\varphi _0(x,k),$$ (3.52) and in particular to choose $`\varphi _0(x,k)\mathrm{\Phi }_0(x,k)`$. However, in this case $`\varphi _0^{}`$ has at large $`x_1`$ the same behavior as $`\mathrm{\Phi }_0`$, and then from (3.46) it follows that for $`(k_{\mathrm{}}\alpha _1)(k_{\mathrm{}}\kappa _1)<0`$ it is not possible to integrate (3.51) on the infinite interval. Then we integrate (3.51) from some $`x_1^0`$ to $`x_1`$ getting $$\varphi _1(x,k)=\frac{1}{\psi _0^{}(x)}\left[C_1^{}(x_2,k)+_{x_1^0}^{x_1}𝑑x_1^{}\varphi _0^{}(x_1^{},x_2,k)\psi _0^{}(x_1^{},x_2)\right].$$ (3.53) $`C_1^{}(x_2,k)`$ is a “constant” of integration and $`\varphi _1(x,k)`$ is a solution of $`L_1\varphi _1=0`$ iff $$_{x_2}\left(C_1^{}(x_2,k)\right)=W(\varphi _0^{}(x,k),\psi _0^{}(x))|_{x_1=x_1^0}.$$ Substituting $`\varphi _0^{}`$ given by (3.52) into (3.53) and integrating by parts we get $$\varphi _1(x,k)=\mathrm{\Phi }_0(x,k)+\frac{1}{\psi _0^{}(x)}\left[C_1(x_2,k)_{(k_{\mathrm{}}\alpha _1)\mathrm{}}^{x_1}𝑑x_1^{}\mathrm{\Phi }_0(x_1^{},x_2,k)\mathrm{\Psi }_0(x_1^{},x_2,i\alpha _1)\right],$$ (3.54) where $$C_1(x_2,k)=C_1^{}(x_2,k)\mathrm{\Phi }_0(x_1^0,x_2,k)\psi _0^{}(x_1^0,x_2)+_{(k_{\mathrm{}}\alpha _1)\mathrm{}}^{x_1^0}𝑑x_1^{}\mathrm{\Phi }_0(x_1^{},x_2,k)\mathrm{\Psi }_0(x_1^{},x_2,i\alpha _1).$$ (3.55) One can check that due to (3.1) and (3.55) $`C_1`$ does not depend on $`x_2`$. Finally, we obtain from (3.53) a solution of the heat conduction equation with potential $`u_1`$ parametrized by an arbitrary function of $`k`$ $$\varphi _1(x,k)=\mathrm{\Phi }_0(x,k)\frac{\mathrm{\Phi }_0(x,i\kappa _1)}{\mathrm{\Delta }_1(x)}\left[C_1(k)+_{(k_{\mathrm{}}\alpha _1)\mathrm{}}^{x_1}𝑑x_1^{}\mathrm{\Phi }_0(x_1^{},x_2,k)\mathrm{\Psi }_0(x_1^{},x_2,i\alpha _1)\right]$$ (3.56) and we can introduce also the solution $`F_1`$ defined by $$F_1(x,k)=\mathrm{\Phi }_0(x,k)\frac{\mathrm{\Phi }_0(x,i\kappa _1)}{\mathrm{\Delta }_1(x)}_{(k_{\mathrm{}}\alpha _1)\mathrm{}}^{x_1}𝑑x_1^{}\mathrm{\Phi }_0(x_1^{},x_2,k)\mathrm{\Psi }_0(x_1^{},x_2,i\alpha _1).$$ (3.57) Eq. (3.57) shows that $`F_1`$ has the same analytical properties of $`\mathrm{\Phi }_0`$ and an additional pole for $`k=i\alpha _1`$. In fact, since $`\mathrm{\Phi }_0`$ is continuous for $`k=i\alpha _1`$, it follows from (2.9) that $$F_1(x,k_{\mathrm{}}+i(\alpha _1+0))F_1(x,k_{\mathrm{}}+i(\alpha _10))=2\pi \frac{\mathrm{\Phi }_0(x,i\kappa _1)}{\mathrm{\Delta }_1(x)}\delta (k_{\mathrm{}})$$ (3.58) and then $$F_1(x,k)=\frac{1}{ik+\alpha _1}\frac{\mathrm{\Phi }_0(x,i\kappa _1)}{\mathrm{\Delta }_1(x)}+\mathrm{reg}$$ where reg denotes terms which are regular in the limit $`ki\alpha _1`$. Moreover one can check that there exist limits $$\underset{x_1\pm \mathrm{}}{lim}e^{ikx_1+k^2x_2}F_1(x,k)=A_1(\pm ,k)$$ (3.59) where $$A_1(\pm ,k)=1+\frac{(\kappa _1\alpha _1)}{(ik+\alpha _1)}\theta \left(\pm (\kappa _1\alpha _1)\right)$$ (3.60) and so the solution of the modified integral equation (2.21) with potential $`u_1`$ is given by $$\chi _1(x,k)=e^{^{ikx_1+k^2x_2}}\mathrm{\Phi }_1(x,k)$$ (3.61) with $$\mathrm{\Phi }_1(x,k)=\frac{F_1(x,k)}{A_1(signk_{\mathrm{}},k)}.$$ (3.62) Indeed, since $`\mathrm{\Phi }_1`$ satisfies the heat conduction equation with potential $`u_1`$ we have $`{\displaystyle }dx^{}\left(_{x_1}G_0(xx^{},k)\right)u_1(x^{})\chi _1(x^{},k)={\displaystyle }dx^{}\left(_{x_1}G_0(xx^{},k)\right)\times `$ $`\times \left(_{x_2^{}}+_{x_1^{}}^22ik_{x_1^{}}\right)\chi _1(x^{},k)`$ and $`_{x_1}`$ cancels all slowly decaying terms of the Green’s function (see (2.11) and (2.12)) so we can integrate by parts in the right hand side getting $$𝑑x^{}_{x_1}G_0(xx^{},k)u_1(x^{})\chi _1(x^{},k)=_{x_1}\chi _1(x,k).$$ It follows that $`1+{\displaystyle _k_{\mathrm{}}\mathrm{}^{x_1}}𝑑y_1{\displaystyle 𝑑x^{}_{y_1}G_0(y_1x_1^{},x_2x_2^{},k)u_1(x^{})\chi _1(x^{},k)}=`$ $`=`$ $`1+{\displaystyle _k_{\mathrm{}}\mathrm{}^{x_1}}𝑑y_1_{y_1}\chi _1(y_1,x_2,k)=1+\chi _1(x,k)\underset{x_1k_{\mathrm{}}\mathrm{}}{lim}\chi _1(x,k)`$ and this completes the proof since, due to definition (3.61), $`lim_{x_1k_{\mathrm{}}\mathrm{}}\chi _1(x,k)=1.`$ Note that from (3.60) and (3.62) and taking into account (3.1), we see that if $`\kappa _1\alpha _1<0`$ $`\mathrm{\Phi }_1`$ has no poles in the complex $`k`$-plane; if $`\kappa _1\alpha _1>0`$ and $`0<\alpha _1<\kappa _1`$ or $`\kappa _1<\alpha _1<0`$ it has a pole at $`k=i\alpha _1`$ while if $`\alpha _1<\kappa _1<0`$ or $`0<\kappa _1<\alpha _1`$ it has a pole for $`k=i\kappa _1`$ (in any case the pole corresponds to the one which is smaller by modulo). So, strictly speaking, $`\mathrm{\Phi }_1`$ cannot be considered a Jost solution since it may have a pole. We will call it Jost-like solution. As far as the dual solutions are concerned, in the same way as before we get $$\psi _1(x,k)=\mathrm{\Psi }_0(x,k)\frac{\mathrm{\Psi }_0(x,i\alpha _1)}{\mathrm{\Delta }_1(x)}\left[D_1(k)+_{(k_{\mathrm{}}\kappa _1)\mathrm{}}^{x_1}𝑑x_1^{}\mathrm{\Psi }_0(x_1^{},x_2,k)\mathrm{\Phi }_0(x_1^{},x_2,i\kappa _1)\right]$$ (3.63) $$\mathrm{{\rm Y}}_1(x,k)=\mathrm{\Psi }_0(x,k)\frac{\mathrm{\Psi }_0(x,i\alpha _1)}{\mathrm{\Delta }_1(x)}_{(k_{\mathrm{}}\kappa _1)\mathrm{}}^{x_1}𝑑x_1^{}\mathrm{\Psi }_0(x_1^{},x_2,k)\mathrm{\Phi }_0(x_1^{},x_2,i\kappa _1).$$ (3.64) Note that $`\mathrm{{\rm Y}}_1`$, as a function of the spectral parameter $`k`$, has the same analytical properties of $`\mathrm{\Psi }_0`$ and an additional pole for $`k=i\kappa _1`$. Indeed, using (2.9) one can easily check from (3.64) that $$\mathrm{{\rm Y}}_1(x,k)=\frac{1}{ik+\kappa _1}\frac{\mathrm{\Psi }_0(x,i\alpha _1)}{\mathrm{\Delta }_1(x)}+\mathrm{reg}.$$ (3.65) If we compute its asymptotic behavior, we see that $$\underset{x_1\pm \mathrm{}}{lim}e^{ikx_1k^2x_2}\mathrm{{\rm Y}}_1(x,k)=B_1(\pm ,k)$$ (3.66) with $$B_1(\pm ,k)=1+\frac{(\alpha _1\kappa _1)}{(ik+\kappa _1)}\theta (\pm (\kappa _1\alpha _1))$$ (3.67) and then, to obtain the dual Jost solution, satisfying the modified integral equation (2.22), we have to consider $$\mathrm{\Psi }_1(x,k)=\frac{\mathrm{{\rm Y}}_1(x,k)}{B_1(signk_{\mathrm{}},k)}.$$ (3.68) Using (3.67) and (3.65) we can prove that $`\mathrm{\Psi }_1`$ has no poles iff $`\alpha _1\kappa _1<0.`$ For $`0<\alpha _1<\kappa _1`$ or $`\kappa _1<\alpha _1<0`$ it has a pole at $`k=i\alpha _1`$ while if $`\alpha _1<\kappa _1<0`$ or $`0<\kappa _1<\alpha _1`$ it has a pole behavior at $`k=i\kappa _1.`$ ## 4 Two solitons on a generic background Now we are going to iterate once the procedure illustrated above. We consider $$\varphi _1(x,k)=\varphi _0(x,k)a_1(x)\left[C_1(k)+_{(k_{\mathrm{}}\alpha _1)\mathrm{}}^{x_1}𝑑x_1^{}\varphi _0(x_1^{},x_2,k)\psi _0(x_1^{},x_2,i\alpha _1)\right]$$ (4.69) and $$\psi _1(x,k)=\psi _0(x,k)b_1(x)\left[D_1(k)+_{(k_{\mathrm{}}\kappa _1)\mathrm{}}^{x_1}𝑑x_1^{}\psi _0(x_1^{},x_2,k)\varphi _0(x_1^{},x_2,i\kappa _1)\right]$$ (4.70) where $`\varphi _0`$ and $`\psi _0`$ are, respectively, the Jost solution of the original spectral problem and of its dual, that is $$\varphi _0(x,k)=\mathrm{\Phi }_0(x,k),\psi _0(x,k)=\mathrm{\Psi }_0(x,k),$$ and we introduced $`a_1(x)`$ $`=`$ $`{\displaystyle \frac{\varphi _0(x,i\kappa _1)}{\mathrm{\Delta }_1(x)}}b_1(x)={\displaystyle \frac{\psi _0(x,i\alpha _1)}{\mathrm{\Delta }_1(x)}}`$ (4.71) $`\mathrm{\Delta }_1(x)`$ $`=`$ $`c_1+{\displaystyle _{(\kappa _1\alpha _1)\mathrm{}}}𝑑x_1^{}\varphi _0(x_1^{},x_2,i\kappa _1)\psi _0(x_1^{},x_2,i\alpha _1)`$ (4.72) and $`C_1(k)`$ and $`D_1(k)`$ are arbitrary functions of $`k`$ such that $`C_1(ik_{\mathrm{}}),D_1(ik_{\mathrm{}})`$ are real, so that $`\varphi _1(x,ik_{\mathrm{}})`$ and $`\psi _1(x,ik_{\mathrm{}})`$ are real and can be used to generate the binary BT. We proved that $`a_1(x)`$ and $`b_1(x)`$ are solutions of the spectral problem (3.25) with potential $`u_1`$ and of its dual, respectively. Now we consider, starting from potential $`u_1`$, a binary BT to generate $$u_2(x)=u_1(x)2_{x_1}^2\mathrm{log}\mathrm{\Delta }_2(x)$$ (4.73) or $$u_2(x)=u_0(x)2_{x_1}^2\mathrm{log}\mathrm{\Delta }_1(x)\mathrm{\Delta }_2(x)$$ (4.74) with $$\mathrm{\Delta }_2(x)=c_2+_{(\kappa _2\alpha _2)\mathrm{}}^{x_1}𝑑x_1^{}\varphi _1(x_1^{},x_2,i\kappa _2)\psi _1(x_1^{},x_2,i\alpha _2),$$ (4.75) provided it is possible to choose $`\kappa _2`$ and $`\alpha _2`$ such that the integral is convergent. In the general case, when both $`C_1(i\kappa _2)`$ and $`D_1(i\alpha _2)`$ are different from zero, it can be shown from (4.69) and (4.70) that the integral in (4.75) is convergent if and only if $`\kappa _2`$ and $`\alpha _2`$ are such that the intervals $`(\alpha _1,\kappa _1)`$ and $`(\alpha _2,\kappa _2)`$ have non void intersection. When either one or both are equal to zero less stringent conditions are required. Let us observe that, due to (3.56)–(3.57) and (3.63)–(3.64), we have $`\underset{x_1(\kappa _2\alpha _2)\mathrm{}}{lim}\mathrm{\Delta }_2(x)`$ $`=`$ $`c_2`$ $`\underset{x_1(\kappa _2\alpha _2)\mathrm{}}{lim}\mathrm{\Delta }_2(x)e^{(\kappa _2\alpha _2)x_1(\kappa _2^2\alpha _2^2)x_2}`$ $`=`$ $`{\displaystyle \frac{A_1(sign(\kappa _2\alpha _2),i\kappa _2)B_1(sign(\kappa _2\alpha _2),i\alpha _2)}{\kappa _2\alpha _2}}`$ where the coefficient is given by (3.60)–(3.67), that is $`A_1(sign(\kappa _2\alpha _2),i\kappa _2)B_1(sign(\kappa _2\alpha _2),i\alpha _2)=`$ $`=`$ $`1+{\displaystyle \frac{(\kappa _1\alpha _1)(\kappa _2\alpha _2)}{(\alpha _1\kappa _2)(\kappa _2\alpha _1)}}\theta ((\kappa _2\alpha _2)(\kappa _1\alpha _1)).`$ When $`\kappa _2`$ and $`\alpha _2`$ are both inside the interval of $`\kappa _1`$ and $`\alpha _1`$ or both outside, the asymptotics (4) have the same sign at both infinities iff $$(\kappa _2\alpha _2)c_20,$$ (4.77) while for one inside and the other one outside the interval $`(\alpha _1,\kappa _1)`$ a necessary condition for regularity becomes $$(\alpha _1\kappa _1)c_20.$$ (4.78) ### 4.1 Properties of potential The easiest way to study properties of potential (4.74) is to solve the recursion relation and express $`\mathrm{\Delta }_1\mathrm{\Delta }_2`$ in terms of the Jost solutions of the original spectral problem. Inserting (4.69) and (4.70) into (4.75) and integrating by parts using the identity $$\frac{\varphi _0(x,i\kappa _1)\psi _0(x,i\alpha _1)}{\mathrm{\Delta }_1(x)}=_{x_1}\left(\frac{1}{\mathrm{\Delta }_1(x)}\right),$$ (4.79) and the following condition $$(\kappa _2\alpha _2)(\kappa _1\alpha _1)<0,$$ (4.80) we can write $`\mathrm{\Delta }_1\mathrm{\Delta }_2`$ as $$\mathrm{\Delta }_1(x)\mathrm{\Delta }_2(x)=detA_2(x)$$ (4.81) where we introduced the $`2\times 2`$ matrices $`A_2,`$ $`B_2,`$ and$`C_2`$ given by $$A_2(x)=B_2(x)+C_2$$ (4.82) and $`B_2`$ and $`C_2`$ with entries $$c_{11}=c_1,c_{22}=c_2,c_{12}=C_1(i\kappa _2),c_{21}=D_1(i\alpha _2)$$ (4.83) $$\left(B_2\right)_{jl}(x)=_{(\kappa _l\alpha _j)\mathrm{}}^{x_1}𝑑x_1^{}\varphi _0(x_1^{},x_2,i\kappa _l)\psi _0(x_1^{},x_2,i\alpha _j).$$ (4.84) Consequently, from (4.74) and (4.81) we obtain for potential $`u_2`$ the following expression $$u_2=u_02_{x_1}^2\mathrm{log}detA_2(x).$$ (4.85) Due to the result of Prop. 2, it follows from (4.81) that when $`\kappa _2`$ and $`\alpha _2`$ obey (4.80) and are such that the intervals $`\kappa _2,\alpha _2`$ and $`\kappa _1,\alpha _1`$ are contained one into another, the following conditions $$c_{ij}(\kappa _j\alpha _i)0,i,j=1,2,$$ (4.86) ensure that $`\mathrm{\Delta }_1\mathrm{\Delta }_2`$ has no zeros in the $`x`$-plane and, consequently, that the dressed potential (4.85) is regular. When the two intervals, though having non void intersection, are not contained one into another, these conditions are not sufficient to conclude as before that $`\mathrm{\Delta }_1\mathrm{\Delta }_2`$ has no zeros. Then we’ll choose the parameters ordered either according to $`\kappa _1<\alpha _2<\kappa _2<\alpha _1`$ $`\alpha _1<\kappa _2<\alpha _2<\kappa _1`$ (4.87) or to $`\kappa _2<\alpha _1<\kappa _1<\alpha _2`$ $`\alpha _2<\kappa _1<\alpha _1<\kappa _2.`$ (4.88) Now let us compute asymptotic behavior of potential $`u_2`$ along a generic direction $`d=x_1+hx_2`$. From (4.81) and using asymptotics of Jost solutions we get that for $`x_2\pm \mathrm{}`$ at fixed $`d`$ $`\left(\mathrm{\Delta }_1\mathrm{\Delta }_2\right)(dhx_2,x_2)detC_2+`$ $`+{\displaystyle \frac{c_{22}}{\kappa _1\alpha _1}}e^{(\kappa _1\alpha _1)d+(\kappa _1\alpha _1)(\kappa _1+\alpha _1h)x_2}+{\displaystyle \frac{c_{11}}{\kappa _2\alpha _2}}e^{(\kappa _2\alpha _2)d+(\kappa _2\alpha _2)(\kappa _2+\alpha _2h)x_2}`$ $`+{\displaystyle \frac{c_{12}}{\alpha _2\kappa _1}}e^{(\kappa _1\alpha _2)d+(\kappa _1\alpha _2)(\kappa _1+\alpha _2h)x_2}+{\displaystyle \frac{c_{21}}{\alpha _1\kappa _2}}e^{(\kappa _2\alpha _1)d+(\kappa _2\alpha _1)(\kappa _2+\alpha _1h)x_2}`$ $`+Ke^{(\kappa _1\alpha _1+\kappa _2\alpha _2)d+(\kappa _1\alpha _1)(\kappa _1+\alpha _1h)x_2+(\kappa _2\alpha _2)(\kappa _2+\alpha _2h)x_2}`$ (4.89) where we introduced $$K=(\kappa _2\kappa _1)(\alpha _2\alpha _1)\underset{l,j=1,2}{}\frac{1}{\kappa _j\alpha _l}.$$ It is tedious but not difficult to prove that along any direction $`d=x_1+hx_2`$ with $`h\kappa _i+\alpha _j,\kappa _i+\kappa _{i+1},\alpha _i+\alpha _{i+1}`$ for $`i,j=1,2`$ and where $`i+1`$ is intended modulo 2, the contribution of $`\mathrm{\Delta }_1\mathrm{\Delta }_2`$ to the potential $`u_2`$ is exponentially decaying. However, the asymptotic behavior (4.89) along these “special” directions depends widely on matrix $`C_2`$, in the sense that not only the phase shift but also the number and directions of the solitons emerging from the background are determined by the entries of this matrix. We proved that if $`c_{ij}0`$ but $`c_{i+1,j}=c_{i,j+1}=c_{i+1,j+1}=0`$ the potential is decaying along all directions but the one corresponding to $`h=\alpha _i+\kappa _j`$ and along this direction it exhibits asymptotically the expected soliton-like behavior. For $`c_{ii}0`$ and $`c_{i,i+1}=0`$ or $`c_{ii}=0`$ and $`c_{i,i+1}0`$ $`i=1,2`$ we have two “shifted” solitons along the directions $`h=\alpha _i+\kappa _i`$ $`i=1,2`$ or, respectively, $`h=\alpha _i+\kappa _{i+1}`$ $`i=1,2`$. When $`c_{ii}0`$, $`c_{i,i+1}0`$ but $`c_{i+1,i+1}=c_{i+1,i}=0`$ or $`c_{ii}0`$, $`c_{i+1,i}0`$ but $`c_{i+1,i+1}=c_{i,i+1}=0`$ matrix $`C_2`$ is singular and in these cases we have three “tails” along the directions $`h=\alpha _i+\kappa _i,\alpha _i+\kappa _{i+1},\kappa _i+\kappa _{i+1}`$ or, respectively, $`h=\alpha _i+\kappa _i,\alpha _{i+1}+\kappa _i,\alpha _i+\alpha _{i+1}`$. When all entries but one, say $`c_{ij}`$, are different from zero, the soliton directions are given by $`h=\kappa _i+\kappa _{i+1},\alpha _i+\alpha _{i+1}`$ and $`h=\alpha _i+\kappa _i,\alpha _{i+1}+\kappa _{i+1}`$ if $`ij`$ and $`h=\alpha _{i+1}+\kappa _i,\alpha _i+\kappa _{i+1}`$ if $`i=j`$. Finally, when $`c_{ij}0`$ for all $`i,j=1,2`$ the potential exhibits asymptotically two “shifted” solitons along the directions $`h=\alpha _i+\alpha _{i+1}`$ and $`h=\kappa _i+\kappa _{i+1}`$. ### 4.2 Darboux transform As far as the Darboux transforms of $`\varphi _1`$ and $`\psi _1`$ are concerned, we can formally define $$\varphi _2(x,k)=\varphi _1(x,k)a_2(x)\left[C_2(k)+_{(k_{\mathrm{}}\alpha _2)\mathrm{}}^{x_1}𝑑x_1^{}\varphi _1(x_1^{},x_2,k)\psi _1(x_1^{},x_2,i\alpha _2)\right]$$ (4.90) and $$\psi _2(x,k)=\psi _1(x,k)b_2(x)\left[D_2(k)+_{(k_{\mathrm{}}\kappa _2)\mathrm{}}^{x_1}𝑑x_1^{}\psi _1(x_1^{},x_2,k)\varphi _1(x_1^{},x_2,i\kappa _2)\right]$$ (4.91) with $$a_2(x)=\frac{\varphi _1(x,i\kappa _2)}{\mathrm{\Delta }_2(x)},b_2(x)=\frac{\psi _1(x,i\alpha _2)}{\mathrm{\Delta }_2(x)}.$$ (4.92) However, due to (3.60) and (3.67), the integrals in (4.90)–(4.91) are convergent for arbitrary $`k`$ only if $$(\alpha _2\kappa _1)(\alpha _2\alpha _1)<0(\alpha _1\kappa _2)(\kappa _1\kappa _2)<0,$$ (4.93) and these conditions require the parameters to be ordered according to (4.87). In other words, like in KPI case , the recursion procedure for getting the solutions is well defined only for a proper choice of order for parameters, even though we proved that the potential is well defined and regular in more general situations. Now we have to solve the recursion procedure. If we substitute into (4.90) the expressions (4.69) and (4.70) for $`\varphi _1`$ and $`\psi _1`$ and integrate by parts using again (4.79), we find $$\varphi _2(x,k)=\frac{1}{detA_2(x)}\left|\begin{array}{ccc}A_{11}(x)\hfill & A_{12}(x)\hfill & \beta _1(x,k)+C_1(k)\hfill \\ A_{21}(x)\hfill & A_{22}(x)\hfill & \beta _2(x,k)+C_2(k)\hfill \\ \varphi _0(x,i\kappa _1)\hfill & \varphi _0(x,i\kappa _2)\hfill & \varphi _0(x,k)\hfill \end{array}\right|$$ (4.94) where matrix $`A_2`$ is given by (4.82) and $$\beta _j(x,k)=_{(k_{\mathrm{}}\alpha _j)\mathrm{}}^{x_1}𝑑x_1^{}\varphi _0(x_1^{},x_2,k)\psi _0(x_1^{},x_2,i\alpha _j).$$ (4.95) Clearly, since $`\varphi _2`$ given by (4.94) is a solution of the spectral problem for $`u_2`$ for any $`C_1(k)`$ and $`C_2(k)`$, properties of determinants ensure that also $$F_2(x,k)=\frac{1}{detA_2(x)}\left|\begin{array}{ccc}A_{11}(x)\hfill & A_{12}(x)\hfill & \beta _1(x,k)\hfill \\ A_{21}(x)\hfill & A_{22}(x)\hfill & \beta _2(x,k)\hfill \\ \varphi _0(x,i\kappa _1)\hfill & \varphi _0(x,i\kappa _2)\hfill & \varphi _0(x,k)\hfill \end{array}\right|$$ (4.96) and $$f_2(x,k)=\frac{1}{detA_2(x)}\left|\begin{array}{ccc}A_{11}(x)\hfill & A_{12}(x)\hfill & C_1(k)\hfill \\ A_{21}(x)\hfill & A_{22}(x)\hfill & C_2(k)\hfill \\ \varphi _0(x,i\kappa _1)\hfill & \varphi _0(x,i\kappa _2)\hfill & 0\hfill \end{array}\right|$$ (4.97) are solutions of the same spectral problem and $$\varphi _2(x,k)=F_2(x,k)+f_2(x,k).$$ (4.98) One can easily obtain the analogous expression for dual solutions, that is $$\psi _2(x,k)=\frac{1}{detA_2(x)}\left|\begin{array}{ccc}A_{11}(x)\hfill & A_{12}(x)\hfill & \beta _1^d(x,k)+D_1(k)\hfill \\ A_{21}(x)\hfill & A_{22}(x)\hfill & \beta _2^d(x,k)+D_2(k)\hfill \\ \psi _0(x,i\alpha _1)\hfill & \psi _0(x,i\alpha _2)\hfill & \psi _0(x,k)\hfill \end{array}\right|$$ (4.99) and $$\mathrm{{\rm Y}}_2(x,k)=\frac{1}{detA_2(x)}\left|\begin{array}{ccc}A_{11}(x)\hfill & A_{12}(x)\hfill & \beta _1^d(x,k)\hfill \\ A_{21}(x)\hfill & A_{22}(x)\hfill & \beta _2^d(x,k)\hfill \\ \psi _0(x,i\alpha _1)\hfill & \psi _0(x,i\alpha _2)\hfill & \psi _0(x,k)\hfill \end{array}\right|$$ (4.100) where $$\beta _j^d(x,k)=_{(k_{\mathrm{}}\kappa _j)\mathrm{}}^{x_1}𝑑x_1^{}\psi _0(x_1^{},x_2,k)\varphi _0(x_1^{},x_2,i\kappa _j).$$ (4.101) For $`F_2`$ and $`\mathrm{{\rm Y}}_2`$ one has the expressions $$F_2(x,k)=F_1(x,k)a_2(x)_{(k_{\mathrm{}}\alpha _2)\mathrm{}}^{x_1}𝑑x_1^{}F_1(x_1^{},x_2,k)\psi _1(x_1^{},x_2,i\alpha _2),$$ (4.102) $$\mathrm{{\rm Y}}_2(x,k)=\mathrm{{\rm Y}}_1(x,k)b_2(x)_{(k_{\mathrm{}}\kappa _2)\mathrm{}}^{x_1}𝑑x_1^{}\mathrm{{\rm Y}}_1(x_1^{},x_2,k)\varphi _1(x_1^{},x_2,i\kappa _2).$$ (4.103) Then, if we compute $$A_2(\pm ,k)=\underset{x_1\pm \mathrm{}}{lim}e^{ikx_1+k^2x_2}F_2(x,k)$$ (4.104) we find $$A_2(\pm ,k)=A_1(\pm ,k)+\frac{\kappa _2\alpha _2}{ik+\alpha _2}A_1(\pm ,k)B_1(\pm ,i\alpha _2)\theta (\pm (\kappa _2\alpha _2))$$ where we used that, due to (4.87), $$\underset{x_1\pm \mathrm{}}{lim}\frac{e^{(\kappa _2\alpha _1)x_1}}{\left(1+e^{(\kappa _2\alpha _2)x_1}\right)\left(1+e^{(\kappa _1\alpha _1)x_1}\right)}=0.$$ Taking into account (4.80) and (3.67), we can write the following recursion relation $$A_2(\pm ,k)=A_1(\pm ,k)\left[1+\frac{\kappa _2\alpha _2}{ik+\alpha _2}\theta (\pm (\kappa _2\alpha _2))\right].$$ (4.105) Analogously, from (4.103) one can check that $`B_2(\pm ,k)`$ defined as $$B_2(\pm ,k)=\underset{x_1\pm \mathrm{}}{lim}e^{ikx_1k^2x_2}\mathrm{{\rm Y}}_2(x,k)$$ (4.106) obeys the following recursive relation $$B_2(\pm ,k)=B_1(\pm ,k)\left[1+\frac{\alpha _2\kappa _2}{ik+\kappa _2}\theta (\pm (\kappa _2\alpha _2))\right].$$ (4.107) Then we obtain the Jost solution and the dual Jost solution for potential $`u_2`$ as $`\mathrm{\Phi }_2(x,k)`$ $`=`$ $`{\displaystyle \frac{F_2(x,k)}{A_2(k_{_{\mathrm{}}},k)}}`$ $`\mathrm{\Psi }_2(x,k)`$ $`=`$ $`{\displaystyle \frac{\mathrm{{\rm Y}}_2(x,k)}{B_2(k_{_{\mathrm{}}},k)}},`$ since it is easy to check that they satisfy integral equations (2.21) and (2.22) with $`u`$ substituted by $`u_2`$. ## Acknowledgments I am very grateful to M. Boiti, F. Pempinelli and A. K. Pogrebkov for help, suggestions and fruitful discussions.
warning/0004/hep-ph0004100.html
ar5iv
text
# 1 Introduction ## 1 Introduction The Super-Kamiokande results on atmospheric neutrino flux measurement show a deficit of the $`\nu _\mu `$ flux . Two generation analyses of the SK data show that the $`\nu _\mu \nu _\tau `$ oscillation hypothesis provides a very good fit to the SK data <sup>4</sup><sup>4</sup>4The $`\nu _\mu \nu _s`$ solution is now ruled out at 99% C.L. by the SK collaboration .. The high statistics of SK also makes it possible to study the zenith-angle dependence of the neutrino flux from which one can conclude that the $`\nu _\mu `$’s show signs of oscillation but the $`\nu _e`$ events are consistent with the no-oscillation hypothesis. Independently the results from the reactor experiment CHOOZ disfavours the $`\nu _\mu \nu _e`$ oscillation hypothesis in a two-generation analysis. It is important however to see the implications of these results in a three-generation picture. The most popular three-generation picture in the context of the SK data is the scenario shown in fig. 1a, where one of the mass squared differences is in the solar neutrino range and the other is suitable for atmospheric neutrino oscillations . In such a scheme one mass scale dominance applies for atmospheric neutrinos and the relevant probabilities are functions of two of the mixing angles and one mass squared difference. This picture however cannot explain the LSND results . In this paper we perform a three flavor $`\chi ^2`$-analysis of the SK atmospheric neutrino data assuming a mass pattern with $`\mathrm{\Delta }_{12}\mathrm{\Delta }_{13}`$ fixed in the eV<sup>2</sup> range and allowing the other mass scale to vary arbitrarily. This mass pattern is shown in fig. 1b. Apart from being suitable to explain the SK atmospheric neutrino data this spectrum is also interesting for the laboratory based neutrino oscillation experiments as the higher mass scale is explorable in the short base line experiments, whereas the lower mass scale can be probed in the long base line experiments. In this scheme to a good approximation, neutrino oscillation in the short-base line accelerators and reactors will be governed by one (the higher) mass scale – and only two of the mixing angles appear in the expressions for the oscillation probabilities. For the atmospheric and the long baseline experiments the characteristic energy and length scales are such that in general both mass differences are of relevance and the probabilities involve all the three mixing angles. However the higher mass scale gives rise to $`\mathrm{\Delta }m^2`$ independent average oscillations and it does not enter the $`\chi ^2`$ fit directly. We determine the best-fit values of $`\mathrm{\Delta }_{23}`$ and the three mixing angles by performing a $`\chi ^2`$ analysis of * the SK atmospheric neutrino data * SK atmospheric and CHOOZ data Finally we compare the allowed values of the mixing angles as obtained from the above analysis with those allowed by the other accelerator and reactor neutrino oscillation data including LSND and KARMEN2. The mass scheme of this paper was first considered in after the declaration of the LSND result. These papers performed a combined three generation analysis of accelerator and reactor results as well as the Kamiokande atmospheric neutrino data. Three-generation picture with the higher mass difference in the eV<sup>2</sup> range and the lower mass difference in the atmospheric range has also been considered in (pre-SK) and (post-SK). These papers attempted to explain both solar and atmospheric neutrino anomalies mainly by maximal $`\nu _\mu \nu _e`$ oscillations driven by $`\mathrm{\Delta }_{ATM}10^3`$ eV<sup>2</sup>. Although it was claimed in that this scenario can provide a good fit to all the available data on neutrino oscillations, it was shown in and also later in that this scenario cannot reproduce the zenith angle dependence of the SK atmospheric neutrino data. In this paper our aim is to determine the allowed oscillation parameter ranges consistent with SK atmospheric, CHOOZ, LSND and other accelerator and reactor experiments. The solar neutrino problem can be explained by invoking a sterile neutrino. We discuss in the conclusions how the solar neutrino flux suppression can be explained in our scenario. The plan of the paper is as follows. In section 2 we discuss very briefly the atmospheric neutrino code employed for the analysis of the SK data. In section 3 we present the formalism for three-generation oscillation analysis and calculate the required probabilities including the earth matter effects. In section 4 we present the $`\chi ^2`$ analysis of only SK atmospheric neutrino data. In section 5 we present the combined $`\chi ^2`$ analysis of SK and CHOOZ data. In section 6 we compare the allowed values of mixing angles from the above analyses with those allowed by the other accelerator and reactor data including the latest results from LSND and KARMEN2. In section 7 we discuss the implications of our results for the future long baseline experiments and end in section 8 with some discussions and conclusions. ## 2 The Atmospheric Neutrino Code We define the quantities $`N_{\mu }^{}{}_{osc}{}^{}`$ and $`N_{e}^{}{}_{osc}{}^{}`$ as $`N_{\mu }^{}{}_{osc}{}^{}=N_{\mu \mu }+N_{e\mu }`$ $`N_{e}^{}{}_{osc}{}^{}=N_{ee}+N_{\mu e}`$ $`N_{e,\mu }^{}{}_{osc}{}^{}`$ are the numbers of $`e`$-like and $`\mu `$-like events in the detector and $`N_{ll^{}}`$ is defined as $`N_{ll^{}}`$ $`=`$ $`n_T{\displaystyle _0^{\mathrm{}}}𝑑E{\displaystyle _{(E_l^{})_{min}}^{(E_l^{})_{\mathrm{max}}}}𝑑E_l^{}{\displaystyle _1^{+1}}d\mathrm{cos}\psi {\displaystyle _1^{+1}}d\mathrm{cos}\xi {\displaystyle \frac{1}{2\pi }}{\displaystyle _0^{2\pi }}𝑑\varphi `$ (1) $`\times `$ $`{\displaystyle \frac{d^2F_l(E,\xi )}{dEd\mathrm{cos}\xi }}{\displaystyle \frac{d^2\sigma _l^{}(E,E_l^{},\mathrm{cos}\psi )}{dE_l^{}d\mathrm{cos}\psi }}ϵ(E_l^{})P_{\nu _l\nu _l^{}}(E,\xi ).`$ $`n_T`$ denotes the number of target nucleons, $`E`$ is the neutrino energy, $`E_l^{}`$ is the energy of the final charged lepton, $`\psi `$ is the angle between the incoming neutrino $`\nu _l`$ and the scattered lepton $`l^{}`$, $`\xi `$ is the zenith angle of the neutrino and $`\varphi `$ is the azimuthal angle corresponding to the incident neutrino direction. The zenith angle of the charged lepton is given by $$\mathrm{cos}\mathrm{\Theta }=\mathrm{cos}\xi \mathrm{cos}\psi +\mathrm{sin}\xi \mathrm{cos}\varphi \mathrm{sin}\psi $$ (2) $`d^2\sigma _l^{}/dE_ld\mathrm{cos}\psi `$ is the differential cross section for $`\nu _l^{}Nl^{}X`$ scattering, $`ϵ(E_l^{})`$ is the detection efficiency for the 1 ring events in the detector and $`P_{\nu _l\nu _l^{}}`$ is the probability of a neutrino flavour $`l`$ to convert to a neutrino of flavour $`l^{}`$. We use the atmospheric neutrino fluxes $`\frac{d^2F_l(E,\xi )}{dEd\mathrm{cos}\xi }`$ from . For further details regarding the calculation of number of events we refer to . ## 3 Three-Flavor Analysis ### 3.1 The vacuum oscillation probabilities The general expression for the probability that an initial $`\nu _\alpha `$ of energy $`E`$ gets converted to a $`\nu _\beta `$ after traveling a distance $`L`$ in vacuum is given by, $$P(\nu _\alpha ,0;\nu _\beta ,t)=\delta _{\alpha \beta }4\underset{j>i}{}U_{\alpha i}U_{\beta i}U_{\alpha j}U_{\beta j}\mathrm{sin}^2\left(\frac{\pi L}{\lambda _{ij}}\right)$$ (3) where $`\lambda _{ij}`$ is defined to be the neutrino vacuum oscillation wavelength given by, $$\lambda _{ij}=(2.47\mathrm{m})\left(\frac{E}{MeV}\right)\left(\frac{\mathrm{eV}^2}{\mathrm{\Delta }_{ij}}\right)$$ (4) which denotes the scale over which neutrino oscillation effects can be significant and $`\mathrm{\Delta }_{ij}=m_j^2m_i^2`$. The actual forms of the various survival and transition probabilities depend on the spectrum of $`\mathrm{\Delta }m^2`$ assumed and the choice of the mixing matrix $`U`$ relating the flavor eigenstates to the mass eigenstates. We choose the flavor states $`\alpha =`$ 1,2, and 3 to correspond to e, $`\mu `$ and $`\tau `$ respectively. The most suitable parameterization of $`U`$ for the mass spectrum chosen by us is $`U=R_{13}R_{12}R_{23}`$ where $`R_{ij}`$ denotes the rotation matrix in the $`ij`$-plane. This yields: $$U=\left(\begin{array}{ccc}c_{12}c_{13}& s_{12}c_{13}c_{23}s_{13}s_{23}& c_{13}s_{12}s_{23}+s_{13}c_{23}\\ s_{12}& c_{12}c_{23}& c_{12}s_{23}\\ s_{13}c_{12}& s_{13}s_{12}c_{23}c_{13}s_{23}& s_{12}s_{13}s_{23}+c_{13}c_{23}\end{array}\right)$$ (5) where $`c_{ij}=\mathrm{cos}\theta _{ij}`$ and $`s_{ij}=\mathrm{sin}\theta _{ij}`$ here and everywhere else in the paper. We have assumed CP-invariance so that $`U`$ is real. The above choice of $`U`$ has the advantage that $`\theta _{23}`$ does not appear in the expressions for the probabilities for the laboratory experiments . The probabilities relevant for atmospheric neutrinos are $`P_{\nu _e\nu _e}`$ $`=`$ $`12c_{13}^2c_{12}^2+2c_{13}^4c_{12}^44(c_{13}s_{12}c_{23}s_{13}s_{23})^2(c_{13}s_{12}s_{23}+s_{13}c_{23})^2\mathrm{S}_{23}`$ (6a) $`P_{\nu _\mu \nu _e}`$ $`=`$ $`2c_{13}^2c_{12}^2s_{12}^24c_{12}^2c_{23}s_{23}(c_{13}s_{12}c_{23}s_{13}s_{23})(c_{13}s_{12}s_{23}+s_{13}c_{23})\mathrm{S}_{23}`$ (6b) $`P_{\nu _\mu \nu _\mu }`$ $`=`$ $`12c_{12}^2s_{12}^24c_{12}^4c_{23}^2s_{23}^2\mathrm{S}_{23}`$ (6c) where $`\mathrm{S}_{23}=\mathrm{sin}^2(\pi \mathrm{L}/\lambda _{23})`$. Apart from the most general three generation regime, the following limits are of interest, as we will see later in the context of the SK data: 1. The two-generation limits Because of the presence of more parameters as compared to the one mass scale dominance picture there are twelve possible two-generation limits with the oscillations driven by either $`\mathrm{\Delta }_{LSND}`$ or $`\mathrm{\Delta }_{ATM}`$. Below we list these limits specifying the mass scales that drive the oscillations: * $`s_{12}0,s_{13}0(\nu _\mu \nu _\tau ,\mathrm{\Delta }_{ATM}`$) $`s_{12}1,s_{13}0(\nu _e\nu _\tau ,\mathrm{\Delta }_{ATM}`$) $`s_{12}0,s_{13}1(\nu _\mu \nu _e,\mathrm{\Delta }_{ATM}`$) $`s_{12}1,s_{13}1(\nu _e\nu _\tau ,\mathrm{\Delta }_{ATM}`$) * $`s_{13}0,s_{23}0(\nu _\mu \nu _e,\mathrm{\Delta }_{LSND}`$) $`s_{13}0,s_{23}1(\nu _\mu \nu _e,\mathrm{\Delta }_{LSND}`$) $`s_{13}1,s_{23}0(\nu _\mu \nu _\tau ,\mathrm{\Delta }_{LSND}`$) $`s_{13}1,s_{23}1(\nu _\mu \nu _\tau ,\mathrm{\Delta }_{LSND}`$) * $`s_{12}0,s_{23}0(\nu _e\nu _\tau ,\mathrm{\Delta }_{LSND}`$) $`s_{12}0,s_{23}1(\nu _e\nu _\tau ,\mathrm{\Delta }_{LSND}`$) $`s_{12}1,s_{23}0(\nu _e\nu _\tau ,\mathrm{\Delta }_{ATM}`$) $`s_{12}1,s_{23}1(\nu _e\nu _\tau ,\mathrm{\Delta }_{ATM}`$) 2. $`s_{12}^2`$ = 0.0 In this limit the relevant probabilities become $`P_{\nu _e\nu _e}`$ $`=`$ $`12c_{13}^2s_{13}^2+4s_{13}^2c_{23}^2s_{23}^2S_{23}`$ (7a) $`P_{\nu _e\nu _\mu }`$ $`=`$ $`4s_{13}^2s_{23}^2c_{23}^2S_{23}`$ (7b) $`P_{\nu _\mu \nu _\mu }`$ $`=`$ $`14c_{23}^2s_{23}^2S_{23}`$ (7c) Thus $`P_{\nu _\mu \nu _\mu }`$ is the same as the two generation limit, $`P_{\nu _\mu \nu _e}`$ is governed by two of the mixing angles and one mass scale and $`P_{\nu _e\nu _e}`$ is governed by two mixing angles and both mass scales. 3. $`s_{13}^2`$ = 0.0 For this case the probabilities take the form $`P_{\nu _e\nu _e}`$ $`=`$ $`12c_{12}^2s_{12}^24s_{12}^4c_{23}^2s_{23}^2S_{23}`$ (8a) $`P_{\nu _e\nu _\mu }`$ $`=`$ $`2c_{12}^2s_{12}^24c_{12}^2s_{12}^2c_{23}^2s_{23}^2S_{23}`$ (8b) $`P_{\nu _\mu \nu _\mu }`$ $`=`$ $`12c_{12}^2s_{12}^24c_{12}^4c_{23}^2s_{23}^2S_{23}`$ (8c) In this case the probabilities are governed by two mass scales and two mixing angles. We note that for cases (2) and (3) the probabilities are symmetric under the transformation $`\theta _{23}\pi /2\theta _{23}`$. The probabilities for these cases are functions of at most two mixing angles as in the OMSD case but they are governed by both mass scales making these limits different from the OMSD limit. ### 3.2 Earth matter effects Since on their way to the detector the upward going neutrinos pass through the earth, it is important in general to include the matter effect in the atmospheric neutrino analysis. The matter contribution to the effective squared mass of the electron neutrinos: $`A=2\sqrt{2}G_FEn_e`$ (9) where $`E`$ is the neutrino energy and $`n_e`$ is the ambient electron density. Assuming a typical density of 5 gm/cc and $`E`$ = 10 GeV, the matter potential $`A3.65\times 10^3`$ eV<sup>2</sup> and since this is of the same order as $`\mathrm{\Delta }_{23}`$ in our case, matter effects should be studied carefully. The mass matrix in the flavor basis in presence of matter is given by $`M_F^2=UM^2U^{}+M_A`$ (10) where $`M^2`$ is the mass matrix in the mass eigenbasis, $`U`$ is the mixing matrix and $`M_A=\left(\begin{array}{ccc}A& 0& 0\\ 0& 0& 0\\ 0& 0& 0\end{array}\right)`$ (11) Since $`\mathrm{\Delta }_{12}\mathrm{\Delta }_{13}\mathrm{\Delta }_{23}A`$, one can solve the eigenvalue problem using the degenerate perturbation theory, where the $`\mathrm{\Delta }_{23}`$ and $`A`$ terms are treated as a perturbation to the dominant $`\mathrm{\Delta }_{12}`$ and $`\mathrm{\Delta }_{13}`$ dependent terms. The mixing angle in matter is then given by $`\mathrm{tan}2\theta _{23}^M={\displaystyle \frac{\mathrm{\Delta }_{23}\mathrm{sin}2\theta _{23}As_{12}\mathrm{sin}2\theta _{13}}{\mathrm{\Delta }_{23}\mathrm{cos}2\theta _{23}A(s_{13}^2c_{13}^2s_{12}^2)}}`$ (12) while the mass squared difference in matter turns out to be $`\mathrm{\Delta }_{23}^M=\left[(\mathrm{\Delta }_{23}\mathrm{cos}2\theta _{23}A(s_{13}^2c_{13}^2s_{12}^2))^2+(\mathrm{\Delta }_{23}\mathrm{sin}2\theta _{23}As_{12}\mathrm{sin}2\theta _{13})^2\right]^{1/2}`$ (13) The mixing angles $`\theta _{12}`$ and $`\theta _{13}`$ as well as the larger mass squared difference $`\mathrm{\Delta }_{12}`$ remain unaltered in matter. From eq. (12) and (13) we note the following * In the limit of both $`s_{12}0`$ and $`s_{13}`$ $`0`$, the matter effect vanishes and we recover the two-generation $`\nu _\mu \nu _\tau `$ limit. * The resonance condition now becomes $`\mathrm{\Delta }_{23}\mathrm{cos}2\theta _{23}=A(s_{13}^2c_{13}^2s_{12}^2)`$. So that for $`\mathrm{\Delta }_{23}>0`$, one can have resonance for both neutrinos – if $`s_{13}^2>c_{13}^2s_{12}^2`$ – as well as for antineutrinos – if $`s_{13}^2<c_{13}^2s_{12}^2`$. This is different from the OMSD picture where for $`\mathrm{\Delta }m^2>0`$ only neutrinos can resonate . * In the limit of $`s_{12}0`$ $$\mathrm{tan}2\theta _{23}^M=\frac{\mathrm{\Delta }_{23}\mathrm{sin}2\theta _{23}}{\mathrm{\Delta }_{23}\mathrm{cos}2\theta _{23}As_{13}^2}$$ (14) Here one gets resonance for neutrinos only (if $`\mathrm{\Delta }_{23}>0`$) and this is similar to the OMSD case. * In the limit $`s_{13}0`$ $$\mathrm{tan}2\theta _{23}^M=\frac{\mathrm{\Delta }_{23}\mathrm{sin}2\theta _{23}}{\mathrm{\Delta }_{23}\mathrm{cos}2\theta _{23}+As_{12}^2}$$ (15) For this case for $`\mathrm{\Delta }_{23}>0`$, there is no resonance for neutrinos but antineutrinos can resonate. * In the limit where $`\mathrm{\Delta }_{23}0`$ $`\mathrm{tan}2\theta _{23}^M={\displaystyle \frac{s_{12}\mathrm{sin}2\theta _{13}}{s_{13}^2c_{13}^2s_{12}^2}},\mathrm{\Delta }_{23}^M=A(s_{13}^2+c_{13}^2s_{12}^2)`$ (16) Thus even for small values of $`\mathrm{\Delta }_{23}<10^4`$ the mass squared difference in matter is $`A`$ and one may still hope to see oscillations for the upward neutrinos due to matter effects. The other point to note is that the mixing angle in matter $`\theta _{23}^M`$ depends only on $`\theta _{12}`$ and $`\theta _{13}`$ and is independent of the vacuum mixing angle $`\theta _{23}`$ and $`\mathrm{\Delta }_{23}`$. Contrast this with the OMSD case (where the expressions for $`\mathrm{tan}2\theta _{23}^M`$ is given by an expression similar to eq. (14) ) and the two-generation $`\nu _\mu \nu _e`$ oscillations. For both the two-generation $`\nu _\mu \nu _e`$ as well as the three-generation OMSD case, for $`\mathrm{\Delta }_{23}0`$, the mixing angle $`\mathrm{tan}2\theta _{23}^M0`$, but for the mass spectrum considered in this paper the $`\mathrm{tan}2\theta _{23}^M`$ maybe large depending on the values of $`s_{12}^2`$ and $`s_{13}^2`$. Hence we see that the demixing effect which gives the lower bound on allowed values of $`\mathrm{\Delta }m^2`$ in the two generation $`\nu _\mu \nu _e`$ or the three-generation OMSD case, does not arise here and we hope to get allowed regions even for very low values of $`\mathrm{\Delta }_{23}`$. On the other hand even small values of $`\theta _{23}`$ in vacuum can get enhanced in matter. This special case where $`\mathrm{\Delta }_{23}0`$ was considered in an earlier paper . * In the limit of $`s_{23}^20`$ $`\mathrm{tan}2\theta _{23}^M={\displaystyle \frac{As_{12}\mathrm{sin}2\theta _{13}}{\mathrm{\Delta }_{23}A(s_{13}^2c_{13}^2s_{12}^2)}}`$ (17) * While for $`s_{23}^21`$ $`\mathrm{tan}2\theta _{23}^M={\displaystyle \frac{As_{12}\mathrm{sin}2\theta _{13}}{\mathrm{\Delta }_{23}A(s_{13}^2c_{13}^2s_{12}^2)}}`$ (18) For the last two cases, corresponding to $`\mathrm{sin}^22\theta _{23}0`$, again the mixing angle $`\theta _{23}`$ in matter is independent of its corresponding value in vacuum and hence for appropriate choices of the other three parameters, $`\mathrm{\Delta }_{23}`$, $`s_{12}^2`$ and $`s_{13}^2`$, one can get large values for $`\mathrm{sin}^22\theta _{23}^M`$ even though the vacuum mixing angle is zero. The amplitude that an initial $`\nu _\alpha `$ of energy $`E`$ is detected as $`\nu _\beta `$ after traveling through the earth is $`A(\nu _\alpha ,t_0,\nu _\beta ,t)={\displaystyle \underset{\sigma ,\lambda ,\rho }{}}{\displaystyle \underset{i,j,k,l}{}}`$ $`[(U_{\beta l}^{M_m}e^{iE_l^{M_m}(tt_3)}U_{\sigma l}^{M_m})(U_{\sigma k}^{M_m}e^{iE_k^{M_c}(t_3t_2)}U_{\lambda k}^{M_c})\times `$ (19) $`(U_{\lambda j}^{M_m}e^{iE_j^{M_m}(t_2t_1)}U_{\rho j}^{M_m})(U_{\rho i}e^{iE_i(t_1t_0)}U_{\alpha i})]`$ where we have considered the earth to be made of two slabs, a mantle and a core with constant densities of 4.5 gm/cc and 11.5 gm/cc respectively and include the non-adiabatic effects at the boundaries. The mixing matrix in the mantle and the core are given by $`U^{M_m}`$ and $`U^{M_c}`$ respectively. $`E_i^Xm_{iX}^2/2E`$, $`X`$ = core(mantle) and $`m_{iX}`$ is the mass of the $`i^{th}`$ neutrino state in the core(mantle). The neutrino is produced at time $`t_0`$, hits the earth mantle at $`t_1`$, hits the core at $`t_2`$, leaves the core at $`t_3`$ and finally hits the detector at time $`t`$. The Greek indices ($`\sigma ,\lambda ,\rho `$) denote the flavor eigenstates while the Latin indices ($`i,j,k,l`$) give the mass eigenstates. The corresponding expression for the probability is given by $`P(\nu _\alpha ,t_0,\nu _\beta ,t)=|A(\nu _\alpha ,t_0,\nu _\beta ,t)|^2`$ (20) For our calculations of the number of events we have used the full expression given by eq.(19) and (20). ## 4 $`\chi ^2`$-analysis of the SK data We minimize the $`\chi ^2`$ function defined as $$\chi ^2=\underset{i,j=1,40}{}\left(N_i^{th}N_i^{exp}\right)(\sigma _{ij}^2)\left(N_j^{th}N_j^{exp}\right)$$ (21) where the sum is over the sub-GeV and multi-GeV electron and muon bins. The experimentally observed number of events are denoted by the superscript “exp” and the theoretical predictions for the quantities are labeled by “th”. The element of the error matrix $`\sigma _{ij}`$ is calculated as in , including the correlations between the different bins. For contained events there are forty experimental data points. The probabilities for the atmospheric neutrinos are explicit functions of one mass-squared difference and three mixing angles making the number of degrees of freedom (d.o.f) 36. The other mass squared difference gives rise to $`\mathrm{\Delta }m^2`$ independent average oscillations and hence does not enter the fit as an independent parameter. For two-flavour $`\nu _\mu \nu _\tau `$ oscillation the 1144 days of data gives the following best-fits and $`\chi _{min}^2`$: * $`\chi _{min}^2/d.o.f.=36.23/38`$, $`\mathrm{\Delta }m^2`$ = 0.0027 eV<sup>2</sup>, $`\mathrm{sin}^22\theta `$ = 1.0 This corresponds to a goodness of fit of 55.14%. For the general three-generation scheme the $`\chi _{min}^2`$ and the best-fit values of parameters that we get are * $`\chi _{min}^2/d.o.f.=34.65/36`$, $`\mathrm{\Delta }_{23}=0.0027`$ eV<sup>2</sup>, $`s_{23}^2=0.51`$,$`s_{12}^2=0.04`$ and $`s_{13}^2=0.06`$ This solution is allowed at 53.28% C.L. The solid(dashed) lines in fig. 2 present the variation of the $`\mathrm{\Delta }\chi ^2=\chi ^2\chi _{min}^2`$ for the SK data, with respect to one of the parameters keeping the other three unconstrained, when we include(exclude) the matter effect. In fig. 2(a) as we go towards smaller values of $`\mathrm{\Delta }_{23}`$ around $`10^3`$ eV<sup>2</sup> the effect of matter starts becoming important as the matter term is now comparable to the mass term. If matter effects are not there then for values of $`\mathrm{\Delta }_{23}`$ $`\stackrel{<}{}10^4`$ eV<sup>2</sup> the $`S_{23}`$ term in eq.(6) is very small and there is no up-down asymmetry resulting in very high values of $`\chi ^2`$ as is evident from the dashed curve. If the matter effects are included, then in the limit of very low $`\mathrm{\Delta }_{23}`$ the matter term dominates and $`\mathrm{\Delta }_{23}^M`$ is given by eq.(16). Since this term $`10^3`$ eV<sup>2</sup> there can be depletion of the neutrinos passing through the earth causing an updown asymmetry. For $`\mathrm{\Delta }_{23}`$ around $`10^4`$ eV<sup>2</sup>, there is cancellation between the two comparable terms in the numerator of eq. (12) and the mixing angle becomes very small and hence the $`\chi ^2`$ around these values of $`\mathrm{\Delta }_{23}`$ comes out to be very high. Fig 2(b) illustrates the corresponding variation of $`\mathrm{\Delta }\chi ^2`$ with $`s_{23}^2`$ while the other three parameters are allowed to vary arbitrarily. For small and large values of $`s_{23}^2`$ the inclusion of matter effect makes a difference. For $`s_{23}^2`$ either very small or large ($`\mathrm{sin}^22\theta _{23}0`$) the overall suppression of the $`\nu _\mu `$ flux is less than that required by the data if vacuum oscillation is operative and so it is ruled out. If we include matter effects then in the limit of $`s_{23}^2=0`$ and $`s_{23}^2=1`$ the matter mixing angle is given by eqs. (17) and (18), which can be large for suitable values of $`s_{13}^2`$ and $`s_{12}^2`$ and hence one gets lower $`\chi ^2`$ even for these values of $`s_{23}^2`$. In figs 2(c) and 2(d) we show the effect of $`s_{12}^2`$ and $`s_{13}^2`$ respectively on $`\mathrm{\Delta }\chi ^2`$. From the solid and the dashed lines it is clear that matter effects do not vary much the allowed ranges of $`s_{12}^2`$ and $`s_{13}^2`$. The dashed-dotted line in the figure shows the 99% C.L. (= 13.28 for 4 parameters) limit. In Table 1 we give the allowed ranges of the mixing parameters, inferred from fig. 2 at 99% C.L. for the SK atmospheric data, with and without matter effects. Table 1: The allowed ranges of parameters for the SK data. | | $`\mathrm{\Delta }_{23}`$ in eV<sup>2</sup> | $`s_{23}^2`$ | $`s_{12}^2`$ | $`s_{13}^2`$ | | --- | --- | --- | --- | --- | | with | $`1.6\times 10^4\mathrm{\Delta }_{23}7.0\times 10^3`$ | $`0.26s_{23}^20.77`$ | $`s_{12}^20.21`$ | $`s_{13}^20.55`$ | | matter effects | $`\mathrm{\Delta }_{23}6.5\times 10^5`$ | $`s_{23}^20.85`$ | | | | without | $`5\times 10^4\mathrm{\Delta }_{23}7.0\times 10^3`$ | $`0.27s_{23}^20.74`$ | $`s_{12}^20.21`$ | $`s_{13}^20.6`$ | | matter effects | | | | | ### 4.1 Zenith-Angle distribution Since the probabilities in our case are in general governed by two mass scales and all three mixing angles it is difficult to understand the allowed regions. To facilitate the qualitative understanding we present in fig. 3 the histograms which describe the zenith angle distribution. The event distributions in these histograms are approximately given by, $$\frac{N_\mu }{N_{\mu _0}}P_{\nu _\mu \nu _\mu }+\frac{N_{e_0}}{N_{\mu _0}}P_{\nu _e\nu _\mu }$$ (22) $$\frac{N_e}{N_{e_0}}P_{\nu _e\nu _e}+\frac{N_{\mu _0}}{N_{e_0}}P_{\nu _\mu \nu _e}$$ (23) where the quantities with suffix 0 indicates the no-oscillation values. For the sub-GeV data $`N_{\mu _0}/N_{e_0}2`$ to a good approximation however for the multi-GeV data this varies in the range 2 (for $`\mathrm{cos}\mathrm{\Theta }`$ =0) to 3 (for $`\mathrm{cos}\mathrm{\Theta }=\pm `$1) . In fig. 3a we study the effect of varying $`s_{12}^2`$ and $`s_{13}^2`$ for fixed values of $`\mathrm{\Delta }_{23}`$ = 0.002 eV<sup>2</sup> and $`s_{23}^2`$ = 0.5. From eq. (12), (13) and from fig. 2 we see that for the values of the $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$ considered in this figure the matter effects are small and we can understand the histograms from the vacuum oscillation probabilities. The thick solid line shows the event distribution for $`s_{12}^2=0`$ and $`s_{13}^2=0.1`$. As $`s_{13}^2`$ increases from 0, keeping $`s_{12}^2`$ as 0, from eqs. (9) $`P_{\nu _e\nu _e}`$ decreases from 1 and $`P_{\nu _e\nu _\mu }`$ increases from zero resulting in a net electron depletion according to eq. (23). The long dashed line corresponds to $`s_{13}^2`$ = 0.3 for which the electron depletion is too high as compared to data. The muon events are also affected as $`P_{\nu _\mu \nu _e}`$ increases with increasing $`s_{13}^2`$ even though $`P_{\nu _\mu \nu _\mu }`$ is independent of $`s_{13}^2`$. On the other hand for $`s_{13}^2`$ = 0.0, the effect of increasing $`s_{12}^2`$ is to increase the number of electron events and decrease the number of muon events according to eqs. (10), (23) and (22). This is shown by the short-dashed and dotted lines in fig. 3a. For $`s_{12}^2`$ = 0.2 the electron excess and muon depletion both becomes too high as compared to the data. For the case when both $`s_{12}^2`$ and $`s_{13}^2`$ are 0.1 the electron depletion caused by increasing $`s_{12}^2`$ and the excess caused by increasing $`s_{13}^2`$ gets balanced and the event distributions are reproduced quite well, shown by the dashed-dotted line. In fig. 3b we study the effect of varying $`s_{23}^2`$ and $`\mathrm{\Delta }_{23}`$ in the limit of $`s_{12}^2=0`$ with $`s_{13}^2`$ fixed at 0.1. Although we use the full probabilities including the matter effect, for 0.004 eV<sup>2</sup> this is not so important and one can understand the histograms from the vacuum oscillation probabilities. For fixed $`\mathrm{\Delta }_{23}`$ as $`s_{23}^2`$ increases, $`P_{\nu _\mu \nu _\mu }`$ decreases, making the muon depletion higher. This is shown in the figure for two representative values of $`\mathrm{\Delta }_{23}`$. The electron events are not affected much by change of $`s_{23}^2`$. The slight increase with $`s_{23}^2`$ is due to increase of both $`P_{\nu _e\nu _e}`$ and $`P_{\nu _\mu \nu _e}`$. To understand the dependence on $`\mathrm{\Delta }_{23}`$ we note that for $`s_{23}^2=0.2`$, if one looks at the vacuum oscillation probabilities, $`N_\mu /N_{\mu 0}10.65S_{23}`$. For 0.004 eV<sup>2</sup> the contribution of $`S_{23}`$ is more resulting in a lower number of muon events. For the electron events however the behavior with $`\mathrm{\Delta }_{23}`$ is opposite, with $`N_e/N_{e0}=0.82+0.12S_{23}`$. Thus with increasing $`\mathrm{\Delta }_{23}`$ the number of electron events increase. Also note that since the contribution of $`S_{23}`$ comes with opposite sign the zenith-angle distribution for a fixed $`\mathrm{\Delta }_{23}`$ is opposite for the muon and the electron events. In fig. 3c we show the histograms in the limit of $`s_{13}^2`$ = 0.0, keeping $`s_{12}^2`$ as 0.1 and varying $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$. As $`s_{23}^2`$ increases all the relevant probabilities decrease and therefore both $`N_\mu /N_{\mu 0}`$ and $`N_e/N_{e0}`$ decrease giving less number of events for both. For this case the $`S_{23}`$ term comes with the same sign (negative) in both $`N_\mu /N_{\mu 0}`$ and $`N_e/N_{e0}`$. Therefore the depletion is more for higher $`\mathrm{\Delta }_{23}`$ for both muon and electron events. Finally, the long dashed line in fig. 3d represent the histograms for the best-fit value for two-generation $`\nu _\mu \nu _\tau `$ oscillations, for which $`P_{\nu _e\nu _e}=1`$. The short dashed line gives the histograms for the three-generation best-fit values. Both give comparable explanation for the zenith angle distribution of the data. The dotted line gives the event distribution for $`\mathrm{\Delta }_{23}=10^5`$ eV<sup>2</sup>. As discussed in section 3.2 even for such low value of $`\mathrm{\Delta }_{23}`$, we find that due to the unique feature of the beyond OMSD neutrino mass spectrum, earth matter effects ensure that both the sub-GeV as well as the multi-GeV upward muon events are very well reproduced, as are the electron events. But since $`s_{12}^2`$ is high, the downward $`\nu _\mu `$ are depleted more than the data requires (eq. (6)). ### 4.2 Allowed parameter region In fig. 4a the solid lines give the 99% C.L. allowed area from SK data in the $`\mathrm{\Delta }_{23}`$-$`s_{23}^2`$ plane keeping the values of $`s_{13}^2`$ and $`s_{12}^2`$ fixed in the allowed range from fig. 2 and Table 1. The first panel represents the two-generation $`\nu _\mu \nu _\tau `$ oscillation limit modulo the difference in the definition of the C.L. limit as the number of parameters are different. We have seen from the histograms in fig. 3a that raising $`s_{12}^2`$ results in electron excess and muon depletion. On the other hand increase in $`s_{13}^2`$ causes electron depletion. The above features are reflected in the shrinking and disappearance of the allowed regions in the first row and column. In the panels where both $`s_{12}^2`$ and $`s_{13}^2`$ are nonzero one may get allowed regions only when the electron depletion due to increasing $`s_{13}^2`$ is replenished by the increase in $`s_{12}^2`$. In fig. 4b we present the 99% C.L. allowed areas in the bilogarithmic $`\mathrm{tan}^2\theta _{12}\mathrm{tan}^2\theta _{13}`$ plane for various fixed values of the parameters $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$. We use the $`\mathrm{log}(\mathrm{tan})`$ representation which enlarges the allowed regions at the corners and the clarity is enhanced. The four corners in this plot refer to the two-generation limits discussed in section 3. The extreme left corner ($`\theta _{12}0,\theta _{13}0`$) correspond to the two generation $`\nu _\mu \nu _\tau `$ oscillation limit. As we move up increasing $`\theta _{13}`$, one has $`\nu _e\nu _\mu `$ and $`\nu _e\nu _\tau `$ mixing in addition and for $`s_{13}^21`$ one goes to the two generation $`\nu _\mu \nu _e`$ oscillation region. For the best-fit values of $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$ if we take $`s_{12}^2`$ and $`s_{13}^2`$ to be 0 and $`1`$ respectively, then the $`\chi _{min}^2`$ is 66.92 which is therefore ruled out. Both the right hand corners in all the panels refer to pure $`\nu _e\nu _\tau `$ oscillations and therefore there are no allowed regions in these zones. For the panels in the first row, $`\mathrm{\Delta }_{23}=0.006`$ eV<sup>2</sup> and the 2-generation $`\nu _\mu \nu _\tau `$ oscillation limit is just disallowed (this can also be seen in the first panel of fig. 4a). The small area allowed for the middle panel of first row (between the solid lines) is due to the fact that for non-zero $`s_{12}^2`$ and $`s_{13}^2`$ the electron events are better reproduced, while $`s_{23}^2=0.5`$ takes care of the muon events. Hence for this case slight mixture of $`\nu _\mu \nu _e`$ and $`\nu _e\nu _\tau `$ oscillations is favoured. This feature was also reflected in the fact that in the fig. 4a, the panel for $`s_{12}^2=0.1`$ and $`s_{13}^2=0.3`$ has more allowed range for $`\mathrm{\Delta }_{23}`$ than the panel for the 2-generation $`\nu _\mu \nu _\tau `$ limit. For the panels with $`\mathrm{\Delta }_{23}=0.002`$ eV<sup>2</sup>, both the pure $`\nu _\mu \nu _\tau `$ limit as well as full three-generation oscillations, give good fit. For the last two rows with $`\mathrm{\Delta }_{23}=0.0007`$ eV<sup>2</sup> and $`0.0004`$ eV<sup>2</sup> the matter effects are important in controlling the shape of the allowed regions. Infact the allowed region that one gets for $`0.0004`$ eV<sup>2</sup> and $`s_{23}^2`$ = 0.5 is the hallmark of the matter effect in this particular three-generation scheme. As can be seen from fig. 2a and Table 1, if one does not include the matter effect, then there are no allowed regions below $`\mathrm{\Delta }_{23}`$ = 0.0005 eV<sup>2</sup> for any arbitrary combination of the other three parameters. Even for the first and the last panels with $`\mathrm{\Delta }_{23}=0.0007`$ eV<sup>2</sup>, one gets allowed areas solely due to matter effects. In fig. 4c the solid lines show the 99% C.L. allowed regions from SK data in the $`s_{23}^2s_{12}^2`$ plane for fixed values of $`\mathrm{\Delta }_{23}`$ and $`s_{13}^2`$. In contrast to the previous figure, here (and in the next figure) we use the $`\mathrm{sin}\mathrm{sin}`$ representation because the allowed regions are around $`\theta _{23}=\pi /4`$ and this region gets compressed in the $`\mathrm{log}(\mathrm{tan})\mathrm{log}(\mathrm{tan})`$ representation. For explaining the various allowed regions we separate the figures in two sets * For $`s_{13}^2`$ = 0.0, the four corners of the panels represent the no-oscillation limits inconsistent with the data. Also as discussed in section 3 for $`s_{23}^2`$ = 0.0 or 1.0 one goes to the limit of pure $`\nu _\mu \nu _e`$ conversions driven by $`\mathrm{\Delta }_{LSND}`$, which is not consistent with data. One obtains allowed regions only when $`s_{23}^2`$ is close to 0.5 with $`s_{12}^2`$ small, so that $`\nu _\mu \nu _\tau `$ conversions are dominant. The allowed range of $`s_{12}^2`$ is controlled mainly by the electron excess as has been discussed before while the allowed range of $`s_{23}^2`$ is determined mostly by the muon depletion. * For $`s_{13}^20`$, the four corners represent the two-generation $`\nu _e\nu _\tau `$ oscillation limit discussed in section 3 and hence these corners are not allowed. For $`s_{23}^2=0.0`$ or 1.0 and $`s_{12}^2`$ 0 or 1 one has $`\mathrm{\Delta }_{LSND}`$ driven $`\nu _\mu \nu _e`$ and $`\nu _\mu \nu _\tau `$ conversion and $`\mathrm{\Delta }_{ATM}`$ driven $`\nu _e\nu _\tau `$ conversions. This scenario is not allowed as it gives excess of electron events and also fails to reproduce the correct zenith angle dependence. For a fixed $`\mathrm{\Delta }_{23}`$ as $`s_{13}^2`$ increases the electron depletion increases which can be balanced by increasing $`s_{12}^2`$ which increases the number of electron events. Hence for a fixed $`\mathrm{\Delta }_{23}`$ the allowed regions shift towards higher $`s_{12}^2`$ values. As in fig. 4b the allowed area in the middle panel of the last row is due to the inclusion of the matter effect. In fig. 4d the solid contours refer to the 99% C.L. allowed areas from SK atmospheric neutrino data in the $`s_{13}^2s_{23}^2`$ plane for various values of $`\mathrm{\Delta }_{23}`$ and $`s_{12}^2`$. * For $`s_{12}^2`$ = 0.0 the corners represent no oscillation limits. In the limit $`s_{23}^20`$ or 1, one gets $`\nu _e\nu _\tau `$ oscillation driven by $`\mathrm{\Delta }_{LSND}`$ which is also not allowed. For $`s_{13}^2`$ = 0.0 and $`s_{23}^20.5`$ one has maximal two-flavour $`\nu _\mu \nu _\tau `$ oscillation limit which is therefore allowed (not allowed for $`\mathrm{\Delta }_{23}=0.006`$ eV<sup>2</sup> as discussed before). As $`s_{13}^2`$ increases the electron depletion becomes higher and that restricts higher $`s_{13}^2`$ values. * For $`s_{12}^20`$, the four corners represent two-generation limits driven by $`\mathrm{\Delta }_{LSND}`$. This is the regime of average oscillations and cannot explain the zenith angle dependence of the data. For a fixed $`\mathrm{\Delta }_{23}`$ the allowed region first expands and then shrinks in size and also shifts towards higher $`s_{13}^2`$ values as $`s_{12}^2`$ increases just as in fig. 4c. Matter effect is important for the last two rows and the increase in the allowed areas for the last two panels of $`\mathrm{\Delta }_{23}=0.0004`$ eV<sup>2</sup> are typical signatures of matter effect. In fig. 4e we present the allowed range in the $`\mathrm{\Delta }_{23}s_{23}^2`$ plane with $`\mathrm{\Delta }_{23}`$ in the $`10^510^4`$ eV<sup>2</sup> range and $`s_{12}^2`$, $`s_{13}^2`$ fixed at 0.185 and 0.372 respectively. We get allowed regions in this range of small $`\mathrm{\Delta }_{23}`$ and small mixing due to matter effects – a feature unique to the mass spectrum considered in this paper. ## 5 $`\chi ^2`$ analysis of the SK + CHOOZ data The CHOOZ experiment can probe upto $`10^3`$ eV<sup>2</sup> and hence it can be important to cross-check the atmospheric neutrino results. In particular a two-generation analysis shows that CHOOZ data disfavours the $`\nu _\mu \nu _e`$ solution to the atmospheric neutrino problem. The general expression for the survival probability of the electron neutrino in presence of three flavours is $$P_{\nu _e\nu _e}=14U_{e1}^2(1U_{e1}^2)\mathrm{sin}^2(\pi L/\lambda _{12})4U_{e2}^2U_{e3}^2\mathrm{sin}^2(\pi L/\lambda _{23})$$ (24) This is the most general expression without the one mass scale dominance approximation. We now minimize the $`\chi ^2`$ defined as $$\chi ^2=\chi _{ATM}^2+\chi _{CHOOZ}^2$$ (25) where we define $`\chi ^2_{CHOOZ}`$ as $$\chi _{CHOOZ}^2=\underset{j=1,15}{}(\frac{x_jy_j}{\mathrm{\Delta }x_j})^2$$ (26) where $`x_j`$ are the experimental values, $`y_j`$ are the corresponding theoretical predictions and the sum is over 15 energy bins of data of the CHOOZ experiment . For the CHOOZ experiment the $`\mathrm{sin}^2(\pi L/\lambda _{12})`$ term does not always average out to 0.5 (for SK this term always averages to 0.5) and one has to do the energy integration properly. For our analysis we keep the $`\mathrm{\Delta }_{12}`$ fixed at 0.5 eV<sup>2</sup> and do a four parameter fit as in SK. The $`\chi _{min}^2`$ and the best-fit values of parameters that we get are * $`\chi _{min}^2/d.o.f.=42.22/51`$, $`\mathrm{\Delta }_{23}`$ = 0.0023 eV<sup>2</sup>, $`s_{23}^2=0.5`$, $`s_{12}^2=0.0022`$ and $`s_{13}^2=0.0`$. Thus the best-fit values shift towards the two-generation limit when we include the CHOOZ result. This provides a very good fit to the data being allowed at 80.45% C.L. The dotted lines in fig. 2 give the combined SK+CHOOZ $`\mathrm{\Delta }\chi ^2(=\chi ^2\chi _{min}^2)`$ given by eq. (25), as a function of one of the parameters, keeping the other three unconstrained. We find that the CHOOZ data severely restricts the allowed ranges for the parameters $`s_{12}^2`$ and $`s_{13}^2`$ to values $`\stackrel{<}{}0.047`$, while $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$ are left almost unaffected. Since CHOOZ is consistent with no oscillation one requires $`P_{\nu _e\nu _e}`$ close to 1. So the second and the third terms in eq. (24) should separately be very small. The second term implies $`U_{e1}^2`$ to be close to either 0 or 1. $`U_{e1}^2`$ close to zero implies either $`s_{12}^2`$ or $`s_{13}^2`$ close to 1 which is not consistent with SK. Therefore $`U_{e1}^2`$ is close to 1. Then from unitarity both $`U_{e2}^2`$ and $`U_{e3}^2`$ are close to 0 and so the third term goes to zero irrespective of the value of $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$. Hence contrary to expectations, CHOOZ puts almost no restriction on the allowed values of $`s_{23}^2`$ and $`\mathrm{\Delta }_{23}`$, although $`\mathrm{\Delta }_{23}10^3`$ eV<sup>2</sup> – in the regime in which CHOOZ is sensitive. On the other hand it puts severe constraints on the allowed values of $`s_{12}^2`$ and $`s_{13}^2`$ in order to suppress the average oscillations driven by $`\mathrm{\Delta }_{12}`$. Because of such low values of $`s_{12}^2`$ and $`s_{13}^2`$ the matter effects for the atmospheric neutrinos are not important and the additional allowed area with low $`\mathrm{\Delta }_{23}`$ and high $`s_{23}^2`$ obtained in the SK analysis due to matter effects are no longer allowed. The 99% C.L. regions allowed by a combined analysis of SK and CHOOZ data is shown by the dotted lines in figs. 4a-d. It is seen that most of the regions allowed by the three-flavour analysis of the SK data is ruled out when we include the CHOOZ result. None of the allowed regions shown in fig. 4a are allowed excepting the two-generation $`\nu _\mu \nu _\tau `$ oscillation limit because CHOOZ does not allow such high values of either $`s_{13}^2`$ or $`s_{12}^2`$. Hence we present again in fig. 5 the allowed regions in the $`\mathrm{\Delta }_{23}s_{23}^2`$ plane for various fixed values of $`s_{12}^2`$ and $`s_{13}^2`$, determined from the dotted lines in fig. 2. The solid lines in fig. 5 give the 99% C.L. area allowed by the SK data while the dotted lines give the corresponding allowed region from the combined analysis of SK+CHOOZ. We find that for the combined analysis we get allowed regions in this plane only for much smaller values of $`s_{12}^2`$ and $`s_{13}^2`$, which ensures that the electron events are neither less nor more than expectations. ## 6 Combined allowed area from short baseline accelerator and reactor experiments As mentioned earlier the higher mass scale of this scenario can be explored in the accelerator based neutrino oscillation search experiments. For the mass-pattern considered the most constraining accelerator experiments are LSND , CDHSW , E531 and KARMEN . Among these only LSND reported positive evidence of oscillation. Other experiments are consistent with no-oscillation hypothesis. Also important in this mass range are the constraints from the reactor experiment Bugey . The relevant probabilities are * Bugey $$P_{\overline{\nu }_e\overline{\nu }_e}=14c_{13}^2c_{12}^2\mathrm{sin}^2(\pi L/\lambda _{12})+4c_{13}^4c_{12}^4\mathrm{sin}^2(\pi L/\lambda _{12})$$ (27) * CDHSW $$P_{\overline{\nu }_\mu \overline{\nu }_\mu }=14c_{12}^2s_{12}^2\mathrm{sin}^2(\pi L/\lambda _{12})$$ (28) * LSND and KARMEN $$P_{\overline{\nu }_\mu \overline{\nu }_e}=4c_{12}^2s_{12}^2c_{13}^2\mathrm{sin}^2(\pi L/\lambda _{12})$$ (29) * E531 $$P_{\nu _\mu \nu _\tau }=4c_{12}^2s_{12}^2s_{13}^2\mathrm{sin}^2(\pi L/\lambda _{13})$$ (30) We note that the probabilities are functions of one of the mass scales and two mixing angles. Thus the one mass scale dominance approximation applies. There are many analyses in the literature of the accelerator and reactor data including LSND under this one mass scale dominance assumption . These analyses showed that when one considers the results from the previous (prior to LSND) accelerator and reactor experiments there are three allowed regions in the $`\theta _{12}\theta _{13}`$ plane * low $`\theta _{12}`$ \- low $`\theta _{13}`$ * low $`\theta _{12}`$ \- high $`\theta _{13}`$ * high $`\theta _{12}`$ \- $`\theta _{13}`$ unconstrained When the LSND result was combined with these results then only the first and the third zones remained allowed in the mass range $`0.5\mathrm{\Delta }_{12}2`$ eV<sup>2</sup>. In these earlier analyses of the accelerator and reactor data E776 was more constraining than KARMEN. But with the new data KARMEN2 gives stronger constraint than E776. Also the results from the KARMEN2 experiment now rule out most of the region allowed by the LSND experiment above 1 eV<sup>2</sup> . The LSND collaboration has also now reduced the value of the transition probability that they see . We have repeated the analysis with the latest LSND and KARMEN results for one representative value of $`\mathrm{\Delta }_{12}=0.5`$ eV<sup>2</sup> and present the allowed region in fig. 6. The light-shaded area in fig. 6 shows the 90% C.L. allowed area in the bilogarithmic $`\mathrm{tan}^2\theta _{12}\mathrm{tan}^2\theta _{13}`$ plane from the observance of no-oscillation in all the other above mentioned accelerator and reactor experiments except KARMEN2. The inclusion of the KARMEN2 results as well gives the 90% C.L. region shown by the area shaded by asterix. The 90% allowed region by the LSND experiment is within the dashed lines. The KARMEN2 data severely restricts the LSND allowed regions. The solid line shows the 90% C.L. ($`\chi ^2\chi _{min}^2+7.78`$) region allowed by the combined $`\chi ^2`$ analysis of the SK+CHOOZ data keeping $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$ at 0.002 eV<sup>2</sup> and 0.5 respectively. The combined SK atmospheric and the CHOOZ reactor data rule out the third zone (high $`\theta _{12}`$ with $`\theta _{13}`$ unconstrained ) allowed from LSND and other accelerator and reactor experiments. Thus if one takes into account constraints from all experiments only a small region in the first zone (small $`\theta _{12},\theta _{13}`$) remains allowed. This common allowed region is shown as a dark-shaded area in the fig. 6. As evident from the expression of the probabilities for the accelerator and reactor experiments the combined allowed area of all the accelerator reactor experiments remains the same irrespective of the value of $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$. Even though the combined area in fig. 6 shows that in the first zone (small $`\theta _{12},\theta _{13}`$), SK+CHOOZ data allows more area in the $`\theta _{12}\theta _{13}`$ plane for $`\mathrm{\Delta }_{23}=0.002`$ eV<sup>2</sup> and $`s_{23}^2=0.5`$, from fig. 4b we see that for some other combinations of $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$ one does not find any allowed zones from the SK+CHOOZ analysis, even at 99% C.L.. For those sets of values of $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$ the SK+CHOOZ analysis is more restrictive than the LSND and other accelerator reactor data. ## 7 Implications From our analysis of the SK atmospheric data the explicit form for the $`3\times 3`$ mixing matrix $`U`$ at the best-fit values of parameters is $$U=\left(\begin{array}{ccc}0.95& 0.039& 0.31\\ 0.2& 0.686& 0.7\\ 0.24& 0.727& 0.644\end{array}\right)$$ (31) From the combined SK+CHOOZ analysis the mixing matrix at the best-fit values of the parameters is $$U=\left(\begin{array}{ccc}0.999& 0.033& 0.033\\ 0.047& 0.706& 0.706\\ 0.0& 0.707& 0.707\end{array}\right)$$ (32) From the combined allowed area of fig. 6 the mixing matrix at $`\mathrm{\Delta }_{12}`$ = 0.5 eV<sup>2</sup>, $`\mathrm{\Delta }_{23}=0.0028`$ eV<sup>2</sup>, $`s_{12}^2=0.005`$, $`s_{13}^2`$ = 0.001 and $`s_{23}^2`$ = 0.5, is $$U=\left(\begin{array}{ccc}0.997& 0.028& 0.072\\ 0.071& 0.705& 0.705\\ 0.032& 0.708& 0.705\end{array}\right)$$ (33) Thus the allowed scenario corresponds to the one where $`\nu _1|\nu _e`$ is close to 1 while the states $`\nu _2`$ and $`\nu _3`$ are combinations of nearly maximally mixed $`\nu _\mu `$ and $`\nu _\tau `$ <sup>5</sup><sup>5</sup>5 Thus this scenario is the same as the one termed 3a in Table VI in the pre-SK analysis of . In their notation the states 2 and 3 were 1 and 2. It was disfavoured from solar neutrino results.. Long baseline (LBL) experiments can be useful to confirm if the atmospheric neutrino anomaly is indeed due to neutrino oscillations, using well monitored accelerator neutrino beams. Some of the important LBL experiments are K2K<sup>6</sup><sup>6</sup>6K2K has already presented some preliminary results. (KEK to SK, L $``$ 250 km), MINOS (Fermilab to Soudan, L $``$ 730 km ) and the proposed CERN to Gran Sasso experiments (L $``$ 730 km) . In this section we explore the sensitivity of the LBL experiment K2K in probing the parameter spaces allowed by the SK+CHOOZ and other accelerator and reactor experiments including LSND. K2K will look for $`\nu _\mu `$ disappearance as well as $`\nu _e`$ appearance. In fig. 7 we show the regions in the $`\mathrm{\Delta }_{23}s_{23}^2`$ plane that can be probed by K2K using their projected sensitivity from . The top left panel is for the two-generation $`\nu _\mu \nu _\tau `$ limit. The other panels are for different fixed values of $`s_{12}^2`$ and $`s_{13}^2`$ while $`\mathrm{\Delta }_{12}`$ is fixed at 0.5 eV<sup>2</sup>. For LBL experiments the term containing $`\mathrm{\Delta }_{12}`$ averages to 0.5 as in the atmospheric case. The solid lines in the panels show the region that can be probed by K2K using the $`\nu _\mu `$ disappearance channel while the dotted lines give the 90% C.L. contours allowed by SK+CHOOZ. One finds that for for $`\mathrm{\Delta }_{23}2\times 10^3`$ eV<sup>2</sup>, the whole region allowed by SK+CHOOZ can be probed by the $`\nu _\mu `$ disappearance channel in K2K. The dashed lines show the 90% C.L. area that K2K can probe by the $`\nu _e`$ appearance mode. As $`s_{12}^2`$ increases the constraint from the $`P_{\nu _\mu \nu _e}`$ channel becomes important as is seen in the top right panel of fig. 7. However such high values of $`s_{12}^2`$, although allowed by SK+CHOOZ, is not favoured when one combines LSND and other accelerator and reactor results. For lower $`s_{12}^2`$ values allowed by all the accelerator, reactor and SK atmospheric neutrino experiment the projected sensitivity in the $`\nu _\mu \nu _e`$ channel of K2K is not enough to probe the allowed regions in the $`\mathrm{\Delta }_{23}s_{23}^2`$ plane as is shown by the absence of the dashed curves in the lower panels. In fig. 8 we show the regions in the bilogarithmic $`\mathrm{tan}^2\theta _{12}\mathrm{tan}^2\theta _{13}`$ plane which can be probed by K2K. For drawing these curves we fix $`\mathrm{\Delta }_{23}=0.002`$ eV<sup>2</sup>, $`s_{23}^2=0.5`$ and $`\mathrm{\Delta }_{12}=0.5`$ eV<sup>2</sup>. Shown is the area that can be explored by the $`\nu _\mu \nu _\mu `$ (left of the solid line) and $`\nu _\mu \nu _e`$ (hatched area) channels in K2K at 90% C.L.. The light-shaded area is allowed by SK+CHOOZ and the dark shaded area is allowed by the combination of all the accelerator, reactor and SK atmospheric neutrino data at 90% C.L.. It is clear from the figure that even though the sensitivity of the $`\nu _e`$ appearance channel is not enough, K2K can still probe the combined allowed region in the $`\theta _{12}\theta _{13}`$ plane from $`\nu _\mu `$ disappearance. The projected sensitivities of MINOS and the CERN to ICARUS proposals are lower than K2K and it will be interesting to check if one can probe the regions allowed in this picture better in these experiments. However since in our case the OMSD approximation is not applicable one has to do the energy averaging properly to get the corresponding contours in the three-generation parameters space, and one cannot merely scale the allowed regions from the two-generation plots. For K2K we could use the fig. 5 of to circumvent this problem. However since the analogous information for MINOS and CERN-Gran Sasso proposals is not available to us we cannot check this explicitly. An important question in this context is whether one can distinguish between the OMSD three generation and this mass scheme. In both pictures the SK atmospheric neutrino data can be explained by the dominant $`\nu _\mu \nu _\tau `$ oscillations mixed with little amount of $`\nu _e\nu _\mu (\nu _\tau )`$ transition. However the mixing matrix $`U`$ is different. A distinction can be done if one can measure the mixing angles very accurately. What is the prospect in LBL experiments to distinguish between these pictures? We give below a very preliminary and qualitative discussion on this. If we take $`s_{12}^2`$ = 0.02, $`s_{13}^2`$ = 0.02 and $`s_{23}^2`$ = 0.5, $`P_{\nu _\mu \nu _e}`$ would be (0.038 + 0.0004 $`S_{23}`$). As the second term is negligible one has average oscillations. This is different from the OMSD limit where $`P_{\nu _\mu \nu _e}=4U_{\mu 3}^2U_{e3}^2S_{23}`$ is energy dependent. If one combines the other accelerator and reactor experiments including LSND then the allowed values of of $`s_{12}^2`$ and $`s_{13}^2`$ are even less and choosing $`s_{12}^2`$ = 0.005, $`s_{13}^2`$ = 0.001 and $`s_{23}^2`$ = 0.5 we get $`P_{\nu _e\nu _\mu }=0.010.004S_{23}`$. Here also the term involving $`S_{23}`$ is one order of magnitude smaller and the oscillations will be averaged. Thus this channel has different predictions for the OMSD limit and beyond the OMSD limit. ## 8 Discussions and Conclusions In this paper we have done a detailed $`\chi ^2`$ analysis of the SK atmospheric neutrino data going beyond the OMSD approximation. The mass spectrum chosen is such that $`\mathrm{\Delta }_{12}=\mathrm{\Delta }_{13}`$ eV<sup>2</sup> to explain the LSND data and $`\mathrm{\Delta }_{23}`$ is in the range suitable for the atmospheric neutrino problem. We study in details the implications of the earth matter effects and bring out the essential differences of our mass pattern with the OMSD scenario and the two-generation limits. We first examine in detail what are the constraints obtained from only SK data considering its overwhelming statistics. The allowed regions include * the two-generation $`\nu _\mu \nu _\tau `$ limit (both $`s_{12}^2`$ and $`s_{13}^2`$ zero) * regions where either $`s_{12}^2`$ or $`s_{13}^2`$ is zero; in this limit the probabilities are functions in general of two mixing angles and two mass scales. * the three-generation regions with all three mixing angles non-zero and the probabilities governed by both mass scales. The last two cases correspond to dominant $`\nu _\mu \nu _\tau `$ oscillation with small admixture of $`\nu _\mu \nu _e`$ and $`\nu _e\nu _\tau `$ oscillation. * regions with very low $`\mathrm{\Delta }_{23}`$ ($`<10^4`$ eV<sup>2</sup>) and $`s_{23}^2`$ close to 1, for which the earth matter effects enhance the oscillations of the upward neutrinos and cause an up-down flux asymmetry. This region is peculiar to the mass spectrum considered by us and is absent in the two-generation and the OMSD pictures. We present the zenith angle distributions of the events in these cases. With the inclusion of the CHOOZ result the allowed ranges of the mixing angles $`s_{12}^2`$ and $`s_{13}^2`$ is constrained more ($`\stackrel{<}{}0.047`$), however the allowed ranges of $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$ do not change much (see fig. 2) except that the low $`\mathrm{\Delta }_{23}`$ region allowed by SK due to matter effects is now disallowed. The inclusion of the constraints from LSND and other accelerator and reactor experiments may restrict the allowed area in the $`\theta _{12}\theta _{13}`$ plane for certain values of $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$, but for some other combinations of $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$, SK+CHOOZ turns out to be more constraining. We have included the latest results from LSND and KARMEN2 in our analysis. In order to explain the solar neutrino problem in this picture one has to add an extra light sterlie neutrino. With the new LSND results the allowed 4 neutrino scenarios are * the (2+2) picture where two degenerate mass states are separated by the LSND gap . * the (3+1) scheme with three neutrino states closely degenerate in mass and the fourth one separated from these by the LSND gap . In the separated state is predominantly a sterile state. In , on the other hand, the state sepaerated by the LSND gap has a very small sterile component. The extension of our scenario to the 2+2 picture is straight forward. One has to add an extra sterile state 4 close to the state 1 such that $`\mathrm{\Delta }_{14}`$ is in the solar range. Then we have two almost decoupled two-generation pictures in which the atmospheric neutrino problem is mainly due to $`\nu _\mu \nu _\tau `$ oscillation and the solar neutrino problem is explained by $`\nu _e\nu _s`$ oscillation. The SMA MSW solution for two -generation $`\nu _e\nu _s`$ picture is allowed at $``$ 15% C.L. . A detailed global fit of solar and atmospheric neutrino data under this picture would tell us how much this will change due to the small admixture with the other generations. If on the other hand we assume the 4th state to be close to the 2nd and the third state then we will have a (3+1) picture where the 1 state, separated by the LSND gap, is predominantly $`\nu _e`$. This picture will have difficulties in solving the solar neutrino problem as because of the CHOOZ constraints $`U_{ei}`$ (i=2,3,4) are small so that $`P_{\nu _e\nu _e}`$ $``$ 1 indicating very small suppression of the solar neutrino flux. To conclude, one can get allowed regions from the SK atmospheric neutrino data where both the mass scales and all the three mixing angles are relevant. The beyond one mass scale dominance spectrum considered in this paper allows new regions in the low mass – low mixing regime due to the earth matter effects. With the inclusion of the CHOOZ, LSND and other accelerator reactor results, the allowed regions are constrained severely. It is, in principle, possible to get some signatures in the LBL experiments to distinguish this picture from the OMSD limit. The authors wish to thank Kenji Kaneyuki for sending them the 1144 days SK atmospheric data and the detection efficiencies. They also wish to thank E. Lisi, N. Fornengo and S. Uma Sankar for useful correspondences. Figure Captions Fig. 1: The two possible neutrino mass spectra in a three generation scheme. Fig. 2: The variation of $`\mathrm{\Delta }\chi ^2=\chi ^2\chi _{min}^2`$ with one of the parameters keeping the other three unconstrained. The solid (dashed) line corresponds to only SK data when matter effects are included (excluded) while the dotted curve gives the same for SK+CHOOZ. The dashed-dotted line shows the 99% C.L. limit for 4 parameters. Fig. 3a: The zenith angle distribution of the lepton events with $`\mathrm{\Delta }_{23}=0.002`$ eV<sup>2</sup> and $`s_{23}^2`$ = 0.5 for various combinations of $`s_{12}^2`$ and $`s_{13}^2`$. $`N`$ is the number of events as given by eq. (1) and $`N_0`$ is the corresponding number with survival probability 1. The panels labelled $`SG_\alpha `$ and $`MG_\alpha `$ ($`\alpha `$ can be e or $`\mu `$) give the histograms for the sub-GeV and multi-GeV $`\alpha `$-events respectively. Also shown are the SK experimental data points with $`\pm `$ 1$`\sigma `$ error bars. Fig. 3b: Same as in fig. 3a for fixed $`s_{12}^2=0.1`$ and $`s_{13}^2=0.0`$ varying $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$. Fig. 3c: Same as in fig. 3a fixing $`s_{12}^2=0.0`$ and $`s_{13}^2=0.1`$ for different $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$ values. Fig. 3d: The long-dashed (short-dashed) line gives the zenith angle distribution of the lepton events for the best-fit cases of the two-generation (three-generation) oscillation solutions for SK. The dotted line gives the corresponding distribution for $`\mathrm{\Delta }_{23}=10^5`$ eV<sup>2</sup>, $`s_{12}^2=0.2`$, $`s_{13}^2=0.4`$ and $`s_{23}^2=1.0`$. Fig. 4a: The allowed parameter regions in the $`\mathrm{\Delta }_{23}s_{23}^2`$ plane for various fixed values of $`s_{12}^2`$ and $`s_{13}^2`$, shown at the top of each panel. The solid lines corresponds to the 99% C.L. contours from the SK data alone, while the dotted line gives the 99% contour from the combined analysis of the SK+CHOOZ data. Fig. 4b: Same as 4a but in the bilogarithmic $`\mathrm{tan}^2\theta _{12}\mathrm{tan}^2\theta _{13}`$ plane for fixed values of $`\mathrm{\Delta }_{23}`$ and $`s_{23}^2`$. Fig. 4c: Same as 4a but in the $`s_{12}^2s_{23}^2`$ plane for fixed values of $`s_{13}^2`$ and $`\mathrm{\Delta }_{23}`$. Fig. 4d: Same as 4a but in the $`s_{13}^2s_{23}^2`$ plane for various fixed values of $`s_{12}^2`$ and $`\mathrm{\Delta }_{23}`$. Fig. 4e: The allowed parameter space in the $`\mathrm{\Delta }_{23}s_{23}^2`$ plane with $`\mathrm{\Delta }_{23}`$ in the range $`10^510^4`$ eV<sup>2</sup> and with fixed values of $`s_{12}^2=0.185`$ and $`s_{13}^2=0.372`$. Fig. 5: Same as 4a but for smaller values of $`s_{12}^2`$ and $`s_{13}^2`$, chosen from the range determined by the SK+CHOOZ dashed line in fig. 3. Fig. 6: The area between the dashed lines is the 90% C.L. region allowed by LSND while the light shaded zone gives the 90% C.L. allowed region from the non-observance of neutrino oscillation in the other short baseline accelerator and reactor experiments except KARMEN2. The corresponding area which includes KARMEN2 as well is shown by the region shaded by asterix. The 90% C.L. allowed region from SK+CHOOZ analysis is within the dotted line. The dark shaded area corresponds to the combined allowed region. Fig. 7: 90% C.L. regions in the $`\mathrm{\Delta }_{23}s_{23}^2`$ plane that can be explored by the $`\nu _\mu \nu _\mu `$ (solid line) and $`\nu _\mu \nu _e`$ (dashed line) oscillation channels in the K2K experiment. The area inside the dotted line shows the 90% C.L. region allowed by SK+CHOOZ. The curves are presented for fixed values of $`s_{12}^2`$ and $`s_{13}^2`$ with $`\mathrm{\Delta }_{12}=0.5`$ eV<sup>2</sup>. Fig. 8: Sensitivity of the K2K experiment in the $`\mathrm{tan}^2\theta _{12}\mathrm{tan}^2\theta _{13}`$ plane for $`\mathrm{\Delta }_{23}=0.002`$ eV<sup>2</sup>, $`s_{23}^2=0.5`$ and $`\mathrm{\Delta }_{12}=0.5`$ eV<sup>2</sup>. The area that can be explored by the $`\nu _\mu \nu _\mu `$ (left of solid line) and $`\nu _\mu \nu _e`$ (hatched area) channels in K2K at 90% C.L. is shown. The light-shaded area is allowed by SK+CHOOZ and the dark-shaded region is the combined area allowed by all accelerator and reactor data at 90% C.L..
warning/0004/nucl-th0004044.html
ar5iv
text
# LBNL-45199 McGill/00-12 Baryonic contributions to the dilepton spectra in relativistic heavy ion collisions ## I Introduction One of the main goals of relativistic heavy ion physics is to explore the possibility of studying strongly interacting matter in a highly exotic form: the quark gluon plasma. In this context the production of dileptons in high energy heavy ion collisions has received particular attention in recent years. This interest is due to the fact that dileptons, once produced, essentially decouple from the strongly interacting system and will reach the detectors mostly unscathed. Hence the dileptons serve as a promising probe, as their production rate is biased towards regions of high densities and high temperatures. Information on those is only available indirectly if one is limited to hadronic observables. It is possible to roughly separate the dilepton invariant mass region in three parts: the “soft” dileptons essentially lie below the $`\varphi `$ peak and the “hard” dileptons lie beyond the $`J/\psi `$. The region in-between has been coined the “intermediate mass region”. It is important to realize that different physics can be at work in those different regimes, mainly reflecting the different epochs of the relativistic nuclear collisions. For example, the hard dileptons will receive an important contribution from the Drell-Yan process which will happen in the first instants of the interaction. In this work, we shall restrict our attention to the soft part. This region is of interest as it has been the focus of detailed recent experimental measurements. Those were done by the HELIOS/3 and CERES collaborations, respectively. Those heavy ion experiments have reported an excess of lepton pairs over measurements involving proton-nucleus collisions. This has stimulated a great deal of theoretical activity which has been summarized recently . The theoretical interpretation of the dilepton excess has mostly concentrated on possible in-medium effects. One school of thought attributes the low mass abundance to a dropping vector meson mass, precursor of a chiral symmetry restoration in the dense medium . Another evaluates the in-medium $`\rho `$ spectral function and finds it considerably broadened (its peak value is not shifted appreciably), owing mainly to the coupling of the $`\rho `$ with baryonic resonances , with the N$`{}_{}{}^{}(1520)`$ playing a dominant role as first pointed out in . To be precise, the effect is distributed over several channels . However, the current experimental data are also consistent with a scenario without in-medium modifications . Certainly, data with better statistics and an improved mass resolution are needed to extract any possible in-medium effects as well as to distinguish between the different in-medium modifications. Several issues deserve further investigations in this matter. The dilepton calculations based on in-medium spectral functions receive a substantial contribution from the baryonic sector . This feature is at first view puzzling, since the baryons are a minor component of the hadron population at CERN energies . A rough estimate of the contribution of baryons to the dilepton channel had been provided in , and turned out to be about a factor of two below the contribution of the $`\omega `$-Dalitz decay. With this in mind, we have set out to explicitly investigate the role of baryons in dilepton production in relativistic nuclear collisions. Our paper is organized as follows: we first explicitly examine the role of the $`N^{}(1520)`$ resonance, which will turn out to be the largest baryon dilepton channel. We also examine a few subtleties associated with its off-shell behaviour and its coupling to electromagnetic radiation. We then compare rate calculations with each other in order to get a feeling for the magnitude of the different contributions in a somewhat idealized thermal environment. We then model the dynamics of the full nuclear collision in the UrQMD approach . The transport model will cover aspects that are related to the possible importance of pre-equilibrium dynamics, while insuring a proper reproduction of the hadronic observables. It is important here to insist on the following: throughout this work the effective spectral function of the vector mesons are the vacuum ones , i.e. we will work with an unmodified pion form-factor. Of course the spectral function associated with the electromagnetic current-current correlator will receive contribution from the medium, such as the Dalitz decays of mesons and baryons inside the fireball . Therefore in this language the low mass lepton pairs come from hadron reactions and Dalitz decays. While formally the lepton pair production process can be linked to the imaginary part of the vector meson self-energy , its connection to Dalitz decay is clear . Bear in mind that a consistent transport treatment of in-medium spectral functions is a topic that still requires development. In keeping with our focus on the role of baryons we then formulate a prediction for the dilepton spectra measured in recent low-energy runs of the CERN SPS. We finally conclude. ## II Baryonic Interactions Since one of the main issues we wish to address in this work is the role played by baryons in the production of low mass lepton pairs, this section first deals with technical issues that arise in such calculations. The baryons will mostly manifest themselves through their radiative decay channel into a dilepton. To fix the ideas, we explicitly consider the case of spin 3/2 baryon resonances which will turn out to be the most important contribution in any case. The channel we are considering is thus $`RNe^+e^{}`$, where R denotes the baryon resonance. The interaction Lagrangian is $`_{RN\gamma }={\displaystyle \frac{ieg_1}{2M}}\overline{\psi }_R^\mu \mathrm{\Theta }(z_1)_{\mu \nu }\gamma _\lambda \mathrm{\Gamma }T_3\psi _NF^{\nu \lambda }{\displaystyle \frac{eg_2}{4M^2}}\overline{\psi }_R^\alpha \mathrm{\Theta }_{\alpha \mu }(z_2)T_3\mathrm{\Gamma }_\nu \psi _NF^{\nu \mu }+h.c.,`$ (1) where $`\mathrm{\Theta }_{\mu \nu }(z)=g_{\mu \nu }1/2(1+2z)\gamma _\mu \gamma _\nu `$. $`\mathrm{\Gamma }`$ is either 1 or $`\gamma _5`$ depending upon the parity of the resonances and $`T`$ is a $`3/21/2`$ isospin transition operator or a $`2\times 2`$ Pauli matrix, depending on the isospin of the resonance. $`\mathrm{\Psi }_R^\mu `$ and $`\psi `$ correspond to Rarita-Schwinger and nucleon spinors respectively, $`F^{\mu \nu }`$ is the electromagnetic field tensor, and $`M`$ is the nucleon mass. The influence of the off-shell parameter $`z`$ is seen in calculations of electromagnetic transitions of nucleon resonances into different multipolarities . It should be mentioned here that when the spin 3/2 particle is on-shell the results are independent of parameter $`z`$. Explicit calculation shows that the Rarita-Schwinger projection operator $`\mathrm{\Delta }_{\mu \nu }(q)=(q/+M_R)(g_{\mu \nu }+{\displaystyle \frac{2}{3}}{\displaystyle \frac{q_\mu q_\nu }{M_R^2}}+{\displaystyle \frac{1}{3}}\gamma _\mu \gamma _\nu {\displaystyle \frac{1}{3}}{\displaystyle \frac{q_\mu \gamma _\nu q_\nu \gamma _\mu }{M_R}})`$ (2) which appears in the squared matrix element for the decay of the 3/2 resonance contracted with $`z`$-dependent vertex terms $`\gamma _\mu `$ vanishes for on-shell particles. In other words, only the first term of the $`\mathrm{\Theta }_{\mu \nu }(z)`$ remains operative for vertex involving on-shell 3/2 resonances. The presence of the second term in the Lagrangian is important in order to keep the electric quadrupole (E2) and magnetic dipole (M1) transitions independent . Quantitatively however, we find that the contributions from the second term to the dilepton yield are smaller than that from the first one by an order of magnitude. Before going into more detailed discussions, we point out some differences between this work and some current popular approaches that rely on a nonrelativistic (NR) reduction of $`={\displaystyle \frac{f_{RN\rho }}{m_\rho }}\overline{\mathrm{\Psi }_R^\mu }\gamma ^\nu TF_{\mu \nu }\psi _N`$ (3) to calculate the $`\rho `$ spectral function. To cast it into the form relevant for the NR case, one neglects the lower component of the Dirac spinor and upon simplification one obtains $`={\displaystyle \frac{f_{RN\rho }}{m_\rho }}\mathrm{\Psi }_R^{}(S_k\rho _k\omega \rho _0S_kq_k)\psi _N`$ (4) where $`S_k`$ ($`k`$ = 1, 2, 3) is the 3/2 spin transition operator, $`\rho ^\mu `$ = ($`\rho _0`$, $`\rho _k`$) is the $`\rho `$ meson field, and $`q^\mu =(\omega ,q_k)`$ is its four momentum. The approach represented by Eq. (1) enables a study of the influence of the off-shell parameter $`z`$. This is done later on. By setting $`z=1/2`$ and $`g_2=0`$ and using the the Vector Meson Dominance model (VMD), Eq.(1) can be cast in the form of the Lagrangian of Eq. (3). We note that it is satisfying to have a context where the importance of relativistic effects can be appreciated quantitatively. Those are found to be of some importance in the lower invariant mass region, as we will see shortly. Let us consider the $`N^{}(1520)`$ resonance. In the interaction Lagrangian of Eq. (1) we use parameters obtained in Ref. , where a fit to pion photoproduction multipoles was performed. To be quantitative, we take $`g_1=1.839`$, $`g_2=0.018`$, $`z_1=0.092`$ and $`z_2=0.024`$. There exists other parameter sets which yield comparable $`\chi ^2`$ as far as the photoproduction data is concerned, but in addition we will require sensible results for the calculation of radiative decay channels. The above quoted values of the parameter set yields the following widths: $`\mathrm{\Gamma }_{N^{}(1520)N\gamma }=0.78`$ MeV and $`\mathrm{\Gamma }_{N^{}(1520)N\rho }=22.35`$ MeV. The experimental values are 0.55$`\pm `$0.1 MeV and 24.0 $`\pm `$ 6.0 MeV respectively. Using VMD with the nonrelativistic interaction Lagrangian Eq. (4) and fitting its only constant in order to reproduce $`\mathrm{\Gamma }_{N^{}N\rho }`$, one obtains $`f_\rho =5.5`$ and $`\mathrm{\Gamma }_{N\gamma }`$ = 0.88 MeV. Using the same $`f_\rho `$ in the relativistic Lagrangian, Eq. (3), produces $`\mathrm{\Gamma }_{N\gamma }`$ = 1.13 MeV and $`\mathrm{\Gamma }_{N\rho }`$ = 30 MeV. Given a Lagrangian, the differential partial width into a lepton pair of invariant mass $`M`$ is calculated by standard techniques . We first investigate the effects of the different interactions. We fix the mass of the $`N^{}(1520)`$ at its central value (i.e. at the peak of the vacuum spectral function). The results are shown in Fig. 1. The solid and dashed curves labeled Non-Relativistic and Relativistic I respectively correspond to the cases where the Lagrangian (3) and its NR reduction (4) have been used with the same coupling constant, $`f_\rho `$ = 5.5. Comparing those two curves thus permits a direct assessment of relativistic effects. Those are found to be largest at the photon point and they become smaller at larger invariant masses, where the nucleon has a smaller momentum. The relativistic and NR treatments should then be in agreement at the larger invariant masses and this is indeed the case. The dashed-dotted curve (Relativistic II) represents the differential width obtained using Eq. (1). The difference between the two relativistic results appear to be a simple overall scaling. We have verified that this overall shift in fact corresponds to the ratio of the two radiative decay widths in the two models. We recall the versatility and the success of the relativistic approach represented by Lagrangian (1) in terms of its capability to reproduce pion photoproduction multipoles and decay widths into $`\gamma N`$ and $`\rho N`$ final states. However, the quantitative differences with a nonrelativistic treatment are overall not large. The relativistic approach can also lend itself to a study of off-shell effects (owing to its $`z`$-dependence): a topic towards which we now turn. We have calculated $`d\mathrm{\Gamma }/dM^2`$ for $`N^{}(1520)Ne^+e^{}`$ in three different ways. First using the on-shell mass (“On-Shell”, in Fig. 2), then integrating over the vacuum spectral function of this resonant state with (“Off-Shell I”), and without (“Off-Shell II”) the $`z`$-dependent interaction. This latter case was achieved by setting $`z=1/2`$. It is clear that off-shell effects can arise not only from the mass distribution but also from the associated vertex containing the factor $`\mathrm{\Theta }_{\mu \nu }(z)`$. A comparison of those three approaches is shown in Fig. 2. It is seen that off-shellness can cause differences of at most a few percent. Summarizing this section, one concludes that in this case off-shell effects are not numerically important and that relativistic effects are slightly larger. However, a nonrelativistic treatment remains a sensible approximation. ## III Rates Before going into the transport calculation and thereby dealing with possible complications owing to the dynamics, we present thermal rate calculations of various lepton pair sources and compare them with with each other. As discussed above, the relativistic effects are not large although not necessarily negligible. Note that the difference between the relativistic and non-relativistic results might become substantial, if evaluated in a different frame of reference . However, in the present work we will employ the calculated decay within a transport model, where the widths are always evaluated in the rest frame of the resonance and then simply boosted to the appropriate matter frame. Thus, in this context using the non-relativistic expression for the widths is a good approximation. Using Eq. (4) and VMD, the differential partial width for the decay of baryonic resonance $`R_i`$ into a nucleon and a lepton pair of invariant mass $`M`$ can be written as $$\frac{d\mathrm{\Gamma }_{R_i\rho }}{dM}=\left(\frac{f_{R_iN\rho }}{m_\rho }\right)^2\frac{1}{\pi }S_\mathrm{\Gamma }M\frac{m_N}{m_{R_i}}k_\rho AF_\rho (M)F(k_\rho ^2),$$ (5) where $`A`$ $`=`$ $`k_\rho ^2\mathrm{for}p\mathrm{wave}`$ (6) $`=`$ $`2k_\rho ^2+3M^2\mathrm{for}s\mathrm{wave},`$ (7) and $`F(k_\rho ^2)=\mathrm{\Lambda }^2/(\mathrm{\Lambda }^2+k_\rho ^2)`$, with $`\mathrm{\Lambda }=1.5`$ GeV. In the above, $`f_{R_iN\rho }`$ is the coupling coupling constant, $`S_\mathrm{\Gamma }`$ is the spin sum, and $`k_\rho `$ is the momentum of the rho meson in the rest frame of the baryon resonance. We should note that the vector dominance coupling used here for the $`N^{}(1720)`$ and $`\mathrm{\Delta }(1905)`$ over-estimates the measured photon decay width . Thus our results for the dilepton rates due to these resonances are in fact upper limits. The $`\rho `$ mass distribution is $$F_\rho (M)=\frac{1}{\pi }\frac{m_\rho \mathrm{\Gamma }_\rho (M)}{(M^2m_\rho ^2)^2+(m_\rho \mathrm{\Gamma }_\rho (M))^2}.$$ (8) One uses the differential decay width to derive the differential dilepton production rate. For a general process $`ab+e^+e^{}`$, this is $`{\displaystyle \frac{dR_{ab+e^+e^{}}}{dM^2}}={\displaystyle \frac{Nm_a}{(2\pi )^2}}{\displaystyle \frac{d\mathrm{\Gamma }_{ab+e^+e^{}}}{dM^2}}{\displaystyle _{m_a}^{\mathrm{}}}𝑑E_ap_af_a(E_a){\displaystyle _1^1}𝑑x[1+f_b(e_b)],`$ (9) where $`E_b=(E_aE_b^{}+p_ap_b^{}x)/m_a`$ and $`E_b^{}=(m_a^2+m_b^2M^2)/(2m_a)`$ . In the above we have assumed that $`a`$ and $`b`$ were fermions and thus the $`f`$’s are Fermi-Dirac distribution functions. We first calculate and compare dilepton rates associated with the radiative decay of several hadronic resonances. The species we consider and the associated rates are shown in Fig. 3. We show representative results obtained at normal nuclear matter density. Adding to the sum of baryonic contributions the signal from $`\pi \pi `$ annihilation and that from $`\omega `$ decay one obtains the net rates shown in Fig. 4. The two mesonic sources we have considered are actually the largest ones in the invariant mass range we have chosen to consider. Also shown on Fig. 4 are the rates obtained by Steele, Yamagishi, and Zahed (SYZ) and by Rapp, Urban, Buballa, and Wambach (RUBW) . At very low baryonic densities, the three rates are quite close to each other. As the density grows our rates remain quite close to that obtained in the SYZ formalism. At the highest baryonic density we have considered, our results approximately lie in between those of SYZ and RUBW, for low invariant masses. A marked difference exists in the results of RUBW, that can be understood in terms of the spectral function re-summation technique leading to a suppression at the $`\rho `$ peak . To provide a reference point, the analysis of particle ratios for SPS-collision at 160 GeV per nucleon gives a chemical freeze-out temperature of about $`T=170\mathrm{MeV}`$ and a baryon density of about $`\rho =\rho _0`$ (see e.g. ). Finally, the relative importance of our mesonic and baryonic rates is shown in Fig. 5. The baryons dominate the low invariant mass region while the $`\pi \pi `$ channel picks up near the $`\rho `$ peak. Representative results at normal nuclear matter density are shown. The relative importance of mesons and baryons does not change considerably if one uses the chemical freeze-out temperature of $`170\mathrm{MeV}`$. It is finally important to realize that the actual contribution from the $`\omega `$ Dalitz decay to the final measurements is much higher than it appears from Fig. 5: the dominant part is due to the decay of the final state $`\omega `$’s. ## IV Transport calculation The nuclear dynamics and dilepton production is studied here within the framework of the Ultra-Relativistic Quantum Molecular Dynamics model, UrQMD . In addition to earlier molecular dynamics models its collision term describes the production of all established meson and baryon resonances up to about 2 GeV with all corresponding isospin projections and antiparticle states. In the present model, light meson formation at low energies is modeled by multi-step processes that proceed via intermediate heavy meson and baryon resonances and their subsequent decay . The pole masses, decay widths and branching ratios of all resonances are taken from . However, due to the experimental uncertainties a certain range of parameters might be used to obtain a consistent fit to cross section data. As an example, the production of $`\omega `$ mesons is described in the UrQMD model by the formation and the decay of the $`N^{}(1900)`$ resonance. It decays in 35% of the cases into $`N\pi `$ and 55% into $`N\omega `$. In line with data at SIS energies (1 AGeV), the $`\eta `$ production proceeds not only via $`N^{}(1535)`$, but invokes also nucleon resonances with masses from 1650 MeV to 2080 MeV. A detailed description of the resonance formation cross sections and comparisons to data is give in Ref. . For earlier works on dileptons within the same framework see e.g. . Dilepton production is calculated in the following way. We consider two distinct classes of processes. First, direct decays of vector mesons, i.e. the $`\rho `$ and $`\omega `$ mesons. Second, the Dalitz decay of mesons and baryons during the reaction, specifically the $`\omega `$ and $`a_1`$ meson as well as the $`N^{}(1520)`$. The Dalitz decay of the pion and eta is treated at the end of the simulation according to Ref. . By treating the direct decay of the $`\rho `$ we implicitly also include the contribution form the pion annihilation channel (see e.g. ). In order to avoid double counting, we do not allow $`\rho `$ mesons created via the decay of the $`N^{}(1520)`$ or $`a_1`$-meson to decay into dileptons, since we treat the Dalitz decay of these states explicitly. Since in UrQMD the $`\omega `$-meson does not have a $`\rho \pi `$ decay channel, there is no double counting in this channel. One should note at this point that treating the Dalitz decays as a two-step process via an intermediate $`\rho `$ meson can lead to erroneous results in a transport description: in this case lepton pairs below the minimum mass of $`m<2m_\pi `$ cannot be produced while this is of course kinematically allowed for a Dalitz decay. In order to extract the dilepton yield from UrQMD, we extract for each of the aforementioned channels the lifetime and four-momentum. The properties of the individual hadrons are given by the collision dynamics. If a meson is produced in an elementary collision (may it be due to string fragmentation, meson or baryon resonance decay or annihilation) it gets a mass according to a Breit-Wigner distribution and a 4-momentum vector according to the kinematics of this single reaction. The meson is then assumed to travel on a straight line trajectory until it (i) decays or (ii) collides with another particle and its 4-momentum changes. Any change of the 4-momentum of the meson is considered as a destruction of the original meson and the creation of a new meson with new quantum numbers and a new 4-momentum. Thus for each hadron under consideration UrQMD provides a list of four-momenta $`q`$ and lifetimes $`\mathrm{\Delta }t`$, $`\{\{q,\mathrm{\Delta }t\}_i\}_h`$, where the index $`h`$ labels the hadron, and $`i`$ labels the different four-momenta and time-intervals for which the hadrons exists during the course of the collision.The resulting dilepton spectrum is then obtained from $$\frac{dN_{l^+l^{}}}{dM}=\underset{i}{}\left(\mathrm{\Delta }t_i\mathrm{\Gamma }_{Vl^+l^{}}(M)\delta \left(\sqrt{q_i^2}M\right)\right)_{h=V}$$ (10) for the the direct decay of vector-mesons. For the Dalitz decays, on the other hand we have $$\frac{dN_{l^+l^{}}^{\mathrm{Dalitz}}}{dM}=\underset{i}{}\left(\mathrm{\Delta }t_i\frac{d\mathrm{\Gamma }(q_i^2)_{Rl^+l^{}X}}{dM}\right).$$ (11) If absorption is negligible, this approach is equivalent to the method of adding one dilepton (with appropriate normalization) at each decay vertex. However, this “shining” method gives better numerical statistics compared to a single decay of the vector meson. To get an upper limit on the possible dilepton radiation, particles are allowed to couple to the dilepton channel also during their formation time. If we insist that the particles shine only after their formation time is over, we have verified that the net rate is only reduced by $``$ 25%. Since we are specifically concerned with the role of the baryons in the production of lepton pair, we present a set of predictions for recent CERN low-energy runs where baryon density effects should be larger than for the high energy runs. Those predictions appear in Figs. 6, 7 and 8. To provide a visual scale reference, we also show data by the CERES collaboration for 160 AGeV collisions . Aside from the clearly visible $`\omega `$-peak, we predict that there will be little difference between the “high-energy” (160 AGeV) and “low-energy” (40 AGeV) dilepton invariant mass spectra. At an invariant mass of 400 MeV, the 160 AGeV calculation is 20% higher (including the CERES normalization) than the 40 AGeV one . Here, we have worked with a mass resolution of $`\mathrm{\Delta }M/M=1\%`$. If this resolution can be achieved in experiment, the omega-peak should be clearly visible, providing a strong constraint for presently available model calculations. Furthermore, we find that the contribution of the $`N^{}(1520)`$-Dalitz (thick dashed curve) is very small, even at these low bombarding energy. We finally note, that a calculation based on the model of with the addition of the $`N^{}(1520)`$ channel and constrained by the final state hadronic spectra as predicted by the RQMD model for a 40 AGeV collision, gives virtually the same results. This model also is in fair agreement with the dilepton spectra measured in the higher energy collisions . ## V Conclusion We see that our transport results lead to a very small baryonic contribution to the net dilepton yield for the recently done low energy heavy ion runs of the CERN SPS. This is to be contrasted with the baryon influence on the net rate, as shown in Fig. 5. The reason for this difference is two-fold. First, the baryonic contribution are severely cut down by the CERES acceptance which we apply to our transport results in Figs. 6 and 7. Second, the baryon content of the central rapidity region is smaller than its mesonic content, owing to the dynamics of the stopping power systematics and of the phase space for particle creation. The small baryon contributions obtained here can be compared with similar results obtained with a hydrodynamic model . Therefore it really does appear that the ideal test to pin down the relevant dilepton production scenario would involve a low energy run. A high resolution measurement at the positions of the vector meson vacuum masses would highlight the high baryon density. In such an environment the collective behaviour of the many-body calculations leads to a suppression of the vector meson peak . In this context we have generated a set of predictions and the corresponding dilepton results of the CERN low energy runs are eagerly awaited. We are also looking forward to HADES measurements at the GSI where it has been shown that the dilepton contribution from N(1520) Dalitz decay is most important in the invariant mass region 0.35 $`<M<`$ 0.75 GeV/c<sup>2</sup> . ###### Acknowledgements. CG, CMK, and VK wish to acknowledge the hospitality of the Theory Institute of the University of Giessen, where this work was started. We also would like to thank P. Huovinen for furnishing the SYZ and RUBW rates used in Fig. 4. The work of CG and AKDM is supported in part by the Natural Sciences and Engineering Research Council of Canada and in part the Fonds FCAR of the Québec Government. The work of CMK is supported by the National Science Foundation under Grant No. PHY-9870038, the Welch Foundation under Grant No. A-1358, the Texas Advanced Research Program under Grants No. FY97-010366-0068 and FY99-010366-0081, and the Alexander von Humboldt Foundation. VK and MB are supported by the Director, Office of Energy Research, division of Nuclear Physics of the Office of High Energy and Nuclear Physics of the U.S. Department of Energy under Contract No. DE-AC03-76SF00098. MB is supported by a Feodor-Lynen fellowship of the Alexander von Humboldt Foundation.
warning/0004/hep-ph0004029.html
ar5iv
text
# 1 Introduction ## 1 Introduction The total hadronic cross section in $`\gamma ^{}\gamma ^{}`$ scattering at electron–positron colliders is considered to be a suitable observable for studying the interesting dynamics of QCD at small $`x`$. For sufficiently large photon virtualities one expects this process to be the optimal test of the perturbative (BFKL) Pomeron . The process is being studied by the L3 and the OPAL collaborations at LEP. First analytic calculations based upon the leading order (LO) BFKL approximation have been compared to measurements by the L3 collaboration . In the meantime various aspects of this process have been considered in more detail . In summary, the present situation is similar to that of forward jets at HERA . The data points for the total $`\gamma ^{}\gamma ^{}`$ cross section lie above the two–gluon exchange approximation but clearly below the LO BFKL prediction. First attempts to include the NLO corrections (see also ) to the BFKL approximation are encouraging but not yet conclusive: for a consistent NLO calculation of the $`\gamma ^{}\gamma ^{}`$ cross section one also needs the NLO corrections to the photon impact factors which have not yet been calculated. But even the LO calculations still suffer from several theoretical uncertainties which we would like to discuss in this short note. The LO calculation of the $`\gamma ^{}\gamma ^{}`$ cross section which has been compared with L3 data uses four massless quarks. Due to the charge $`+2/3`$ of the charm quark the cross section is multiplied by a factor of $`2.8`$ when going from three to four massless flavors. It is obviously very desirable to eliminate this big uncertainty. Effects of the charm mass have so far only been considered in where a formula was given in $`x`$-space and its effect was not considered separately. We give a formula for the cross section for non–zero charm quark mass in Mellin space and perform a numerical study which shows that the charm quark mass leads to a considerable reduction of the cross section expected for LEP. The theoretical prediction of the cross section also depends on the value of the strong coupling constant $`\alpha _s`$ which is not too well known at the (comparatively small) momentum scales dominating the kinematics at LEP, especially at $`\sqrt{s}=91\text{GeV}`$. Our numerical study shows that the dependence on $`\alpha _s`$ is rather strong. Further, we briefly discuss the effect of taking into account contributions from all photon polarizations as well as the uncertainty associated with the BFKL energy scale. ## 2 Cross section formula In the events of interest the scattered electron as well as the scattered positron are tagged (‘double–tag events’), and we define useful scaling variables (we use the notation of ) $$x_1=\frac{Q_1^2}{2q_1k_2},x_2=\frac{Q_2^2}{2q_2k_1}$$ (1) and $$y_1=\frac{q_1k_2}{k_1k_2},y_2=\frac{q_2k_1}{k_1k_2},$$ (2) where $`k_1`$ and $`k_2`$ are the momenta of the electron and positron, respectively. We have $$y_i=1\frac{E_{tag}^i}{E_b}\mathrm{cos}^2\left(\frac{\theta _{tag}^i}{2}\right),$$ (3) where $`E_b`$ is the beam energy, and $`E_{tag}^i`$ and $`\theta _{tag}^i`$ are the energy of the tagged lepton and its angle with respect to the beam axis, respectively. The virtualities of the photons are ($`i=1,2`$) $$Q_i^2=q_i^2=2E_bE_{tag}^i(1\mathrm{cos}\theta _{tag}^i),$$ (4) and they are required to be large to make perturbation theory applicable. The squared center–of–mass energy of the $`e^+e^{}`$ collisions is $`s=(k_1+k_2)^2`$, whereas for the underlying $`\gamma ^{}\gamma ^{}`$ process the squared energy is given by $$\widehat{s}=(q_1+q_2)^2sy_1y_2.$$ (5) We consider the limit where $`Q_1^2,Q_2^2`$ and $`\widehat{s}`$ are large and $$Q_1^2,Q_2^2\widehat{s}.$$ (6) The differential $`e^+e^{}`$ cross section has been calculated in and . The total $`\gamma ^{}\gamma ^{}`$ cross section is given by $$\sigma _{\gamma ^{}\gamma ^{}}=\sigma _{\gamma ^{}\gamma ^{}}^{TT}+ϵ_2\sigma _{\gamma ^{}\gamma ^{}}^{TL}+ϵ_1\sigma _{\gamma ^{}\gamma ^{}}^{LT}+ϵ_1ϵ_2\sigma _{\gamma ^{}\gamma ^{}}^{LL}$$ (7) with $$ϵ_i=\frac{2(1y_i)}{1+(1y_i)^2}.$$ (8) The labels $`T`$ and $`L`$ refer to the transverse and longitudinal polarization of the incoming photon, respectively. In it has been argued that in the L3 region the $`ϵ_i`$ have values larger than $`0.97`$. For simplicity we take them as 1 which leads to a deviation of less than 3 percent. For the polarizations $`i,j\{L,T\}`$ of the initial photons one finds $$\sigma _{\gamma ^{}\gamma ^{}}^{ij}=\frac{\alpha _{em}}{16Q_1^2}\frac{\alpha _{em}}{16Q_2^2}\frac{\text{d}\nu }{2\pi ^2}\mathrm{exp}\left[\mathrm{ln}\left(\frac{\widehat{s}}{s_0}\right)\chi (\nu )\right]W_i(\nu ,\frac{m^2}{Q_1^2})W_j(\nu ,\frac{m^2}{Q_2^2}),$$ (9) where $`s_0`$ is the typical BFKL energy scale which we choose as $$s_0=\sqrt{Q_1^2Q_2^2}.$$ (10) Further, $$\chi (\nu )=N_c\alpha _s/\pi [2\psi (1)\psi (1/2+i\nu )\psi (1/2i\nu )],$$ (11) and $`\psi `$ denotes the digamma function. For massless quarks the functions $`W_i`$ are given in . For massive quarks we can use the results found in to obtain $`W_L(\nu ,{\displaystyle \frac{m_c^2}{Q_i^2}})`$ $`=`$ $`{\displaystyle \underset{f=u,d,s}{}}8\sqrt{2}q_f^2\alpha _s\pi ^2{\displaystyle \frac{\nu ^2+\frac{1}{4}}{\nu ^2+1}}{\displaystyle \frac{\mathrm{sinh}\pi \nu }{\nu \mathrm{cosh}^2\pi \nu }}(Q_i^2)^{\frac{1}{2}+i\nu }`$ (12) $`+{\displaystyle \frac{128}{15}}\sqrt{2}q_c^2\alpha _s\pi \sqrt{\pi }{\displaystyle \frac{\mathrm{\Gamma }(\frac{3}{2}+i\nu )}{\mathrm{\Gamma }(2+i\nu )\mathrm{cosh}\pi \nu }}(Q_i^2)^{\frac{1}{2}+i\nu }\left({\displaystyle \frac{Q_i^2+4m_c^2}{Q_i^2}}\right)^{\frac{3}{2}i\nu }\times `$ $`\times {}_{2}{}^{}F_{1}^{}({\displaystyle \frac{3}{2}}+i\nu ,{\displaystyle \frac{1}{2}};{\displaystyle \frac{7}{2}};{\displaystyle \frac{Q_i^2}{Q_i^2+4m_c^2}})`$ and $`W_T(\nu ,{\displaystyle \frac{m_c^2}{Q_i^2}})={\displaystyle \underset{f=u,d,s}{}}4\sqrt{2}q_f^2\alpha _s\pi ^2{\displaystyle \frac{\nu ^2+\frac{9}{4}}{\nu ^2+1}}{\displaystyle \frac{\mathrm{sinh}\pi \nu }{\nu \mathrm{cosh}^2\pi \nu }}(Q_i^2)^{\frac{1}{2}+i\nu }`$ (13) $`+\mathrm{\hspace{0.17em}16}\sqrt{2}q_c^2\alpha _s\pi \sqrt{\pi }{\displaystyle \frac{\mathrm{\Gamma }(\frac{1}{2}+i\nu )}{\mathrm{\Gamma }(2+i\nu )\mathrm{cosh}\pi \nu }}(Q_i^2)^{\frac{1}{2}+i\nu }\left({\displaystyle \frac{Q_i^2+4m_c^2}{Q_i^2}}\right)^{\frac{1}{2}i\nu }\times `$ $`\times \left\{\right[{\displaystyle \frac{(1+3i\nu )Q_i^2+(3+2i\nu )m_c^2}{Q_i^2+4m_c^2}}i\nu \left({\displaystyle \frac{Q_i^2}{Q_i^2+4m_c^2}}\right)^2\left]_2F_1\right({\displaystyle \frac{1}{2}}+i\nu ,{\displaystyle \frac{1}{2}};{\displaystyle \frac{3}{2}};{\displaystyle \frac{Q_i^2}{Q_i^2+4m_c^2}})`$ $`+[(1i\nu ){\displaystyle \frac{Q_i^2}{Q_i^2+4m_c^2}}{\displaystyle \frac{3+2i\nu }{4}}]{}_{2}{}^{}F_{1}^{}({\displaystyle \frac{1}{2}}+i\nu ,{\displaystyle \frac{1}{2}};{\displaystyle \frac{3}{2}};{\displaystyle \frac{Q_i^2}{Q_i^2+4m_c^2}})\}`$ where $`q_f`$ denotes the electric charge of the quark flavor $`f`$. The terms containing $`q_c=2/3`$ correspond to the contribution of the charm quark. In the limit $`m_c0`$ they can be simplified in such a way that the sum over flavours in the first terms is extended to include four flavours. The bottom quark — relevant to experiments at a future Linear Collider — can be included in the same way as the charm quark. (The contribution of the bottom quark is very small at LEP and will be neglected in the following.) ## 3 Numerical results For our numerical analysis we use the kinematics of the L3 detector (see also ). At LEP91 the outgoing electrons close to the forward direction can be detected within the angles $`30\text{mrad}<\theta _{tag}<66\text{mrad}`$, and the electron energy required for tagging is $`E_{tag}>30\text{GeV}`$. This leads to a range of possible photon virtualities $$1.2\text{GeV}^2<Q_{1,2}^2<9\text{GeV}^2,$$ (14) and the cross section is evaluated at the mean value $`Q^2=3.5\text{GeV}^2`$. At LEP183 the energy of the tagged electrons is restricted by $`E_{tag}>40\text{GeV}`$, and the angular range is the same as at LEP91. The virtualities are $$2.5\text{GeV}^2<Q_{1,2}^2<35\text{GeV}^2,$$ (15) with the mean value $`Q^2=14\text{GeV}^2`$. Note that with these mean values of $`Q^2`$ the available range for the rapidity $`Y=\mathrm{ln}(\widehat{s}/s_0)`$ is the same for LEP91 and LEP183. In the BFKL formalism the strong coupling $`\alpha _s`$ is kept fixed, and the natural scale is $`Q^2`$. In our numerical study, however, we use a running coupling in order to arrive at an improved prediction for the cross section. Specifically, we use a running $`\alpha _s`$ in NNLO with four active flavours. The variation of $`\alpha _s`$ at a given scale $`Q^2`$ (see below) is then obtained by varying $`\mathrm{\Lambda }_{\text{QCD}}`$ (approximately between $`100\text{MeV}`$ and $`350\text{MeV}`$). The results of our numerical study are illustrated in the following figures. Fig. 1 shows a comparison between the saddle point approximation to the BFKL cross section and the exact calculation of the LO BFKL Pomeron, here for LEP183 and for transversely polarized photons only. In order to allow for a direct comparison with the theoretical curve used in , the curves in this and in the next figure are calculated for four massless quarks (mass effects are discussed separately below). In the measured data have been compared to the saddle point approximation. We find that this approximation overestimates the exact cross section by 20% to 30%. Fig. 1 also shows the cross section obtained from two–gluon exchange which approximates the full DGLAP formalism if the two photon virtualities are of the same order. Next we show in Fig. 2 the contributions of different polarizations of the two incoming photons, again for LEP183 and with four massless quarks. Clearly the sum of all polarization is substantially larger than the transverse contribution alone which was compared with data in . Now we address the question how the cross section depends on the value of $`\alpha _s`$ chosen at the scale $`Q^2`$. Here and in the following we include the charm quark as a massive quark and sum over all photon polarizations, i. e. we use eq. (7) with (9), (12), (13). As can be seen from eqs. (9) and (11), $`\alpha _s`$ enters in the exponent and thus we expect a strong dependence of the cross section on its particular value. This is confirmed by the numerical results presented in Figs. 3 and 4. Fig. 3 shows the cross section for $`Q^2=3.5\text{GeV}^2`$ (LEP91), and we study a spread $`0.20<\alpha _s<0.28`$ of possible values. In Fig. 3 we have plotted the cross section for $`Q^2=14\text{GeV}^2`$ (LEP183), now varying $`\alpha _s`$ between $`0.18`$ and $`0.24`$. In the theoretical curves were obtained using $`\alpha _s=0.20`$ for both energies. For LEP91 this choice appears very low. To obtain more realistic predictions (see below), we choose $`\alpha _s=0.268`$ (LEP91) and $`\alpha _s=0.208`$ (LEP183), respectively. A further uncertainty is the choice of the BFKL energy scale $`s_0`$. Strictly speaking, its value is not determined in the leading order BFKL formalism. In this sense, the choice (10) is only an educated guess. It is well possible that the true value differs from that choice by a factor of two, for example. Choosing $`s_0`$ as twice (half) the value given in (10) results in a shift of the BFKL prediction to the right (left) by $`\mathrm{ln}20.7`$ units in rapidity. Shifting the curves in our figures by this value amounts to a considerable change in the LO BFKL prediction. However, the uncertainty in $`s_0`$ is less serious in the NLO BFKL calculation since there the typical curves are less steep and a horizontal shift results in a smaller absolute change of the cross section. Finally, we study the effect of the charm quark mass on the cross section. Our results (including also the improvements discussed above) are shown in Fig. 5 for LEP91, and in Fig. 6 for LEP183. For LEP91 the charm quark mass reduces the cross section by about 30 % (compared to a massless charm quark), for LEP183 the cross section is reduced by about 20 %. The effect of the charm mass is different in the two cases due to the different mean values $`Q^2`$. It can be seen immediately from eqs. (12) and (13) that the charm mass $`m_c`$ enters in such a combination with $`Q^2`$ that for increasing $`Q^2`$ the effect of a finite $`m_c`$ becomes weaker. Also shown in Figs. 5 and 6 are the data points measured by the L3 collaboration . One should note that in these data the contribution of the quark box diagram — or quark parton model (QPM) graph — has already been subtracted. This diagram is also not included in our calculation. Compared with the theoretical prediction used in we have made the following changes to arrive at the cross sections in Figs. 5 and 6: exact BFKL is used instead of the saddle point approximation, all photon polarizations are included, and a nonzero charm quark mass is used. Further we have chosen a higher (and more realistic) value for $`\alpha _s`$ in the case of LEP91. Roughly speaking, the first three changes lead to a change in the normalization, whereas the last one results in a different energy dependence. For LEP91 we find a significantly larger cross section than the one given by the theoretical curve in . Here the main effect is due to the choice of $`\alpha _s`$. For LEP183 our prediction is very close to the one used in because the changes almost compensate each other. However, one should have in mind that already a small change in $`\alpha _s`$ leads to a visible modification of the energy dependence. Our predictions for the cross section are higher and exhibit a much faster rise with energy than the measured data points (especially in the case of LEP91). This was to be expected since the NLO corrections to the BFKL equation are known to be sizable and lead to a slower rise with energy. Certainly there is a good chance to successfully describe the data by a NLO BFKL calculation. However, also there one should keep in mind the uncertainties discussed above. ## 4 Conclusions Our study of the BFKL prediction shows that there are substantial theoretical uncertainties even at the LO level. However, also an improved treatment does not change the main conclusion: the LEP data for the total hadronic $`\gamma ^{}\gamma ^{}`$ cross section clearly indicate that a simple two–gluon exchange is not sufficient to describe the data. The LO BFKL prediction, on the other hand, lies well above the data. A consistent NLO calculation including the impact factors is therefore urgently needed. ## Acknowledgements CE would like to thank G. Salam, M. Wadhwa and M. Wüsthoff for helpful discussions.
warning/0004/math-ph0004024.html
ar5iv
text
# 1 Introduction ## 1 Introduction Let $`YX`$ be a real smooth fibre bundle coordinated by $`(x^\lambda ,y^i)`$, and $`\mathrm{\Omega }_{\mathrm{}}^{}`$ the algebra of exterior forms on all finite order jet manifolds of $`YX`$ modulo the pull-back identification. Put dim$`X=n`$. We have the variational complex $$0𝐑\mathrm{\Omega }_{\mathrm{}}^0\stackrel{d_H}{}\mathrm{\Omega }_{\mathrm{}}^{0,1}\stackrel{d_H}{}\mathrm{}\stackrel{d_H}{}\mathrm{\Omega }_{\mathrm{}}^{0,n}\stackrel{\epsilon _1}{}E_1\stackrel{\epsilon _2}{}E_2\mathrm{},$$ (1) where $`\mathrm{\Omega }_{\mathrm{}}^{0,}`$ is a subalgebra of horizontal (semibasic) exterior forms, $`d_H`$ is the horizontal exterior differential, $`\epsilon _1`$ is the Euler–Lagrange map, $`\epsilon _2`$ is the Helmholtz–Sonin map, and so on . This complex provides the algebraic approach to the calculus of variations in field theory. The well-known theorem states the local exactness of this complex as follows. THEOREM 1. If a fibre bundle $`Y`$ is $`𝐑^{n+m}𝐑^n`$, the variational complex (1) except the first term is exact, i.e. $`(i)\mathrm{Ker}d_H=\mathrm{Im}d_H,(ii)\mathrm{Ker}\epsilon _1=\mathrm{Im}d_H,(iii)\mathrm{Ker}\epsilon _{k+1}=\mathrm{Im}\epsilon _k.`$ REMARK 1. As in the De Rham complex on a connected manifold, one puts $`H^0(d_H)=𝐑`$ because, if $`d_Hf=0`$, $`f\mathrm{\Omega }_{\mathrm{}}^0`$, then $`f=`$const. From now on, by $`H^{}`$ we will mean the cohomology groups of order $`k>0`$. Restricted to $`\mathrm{\Omega }_{\mathrm{}}^{0,}`$, the variational complex (1) yields the so called horizontal complex $$0𝐑\mathrm{\Omega }_{\mathrm{}}^0\stackrel{d_H}{}\mathrm{\Omega }_{\mathrm{}}^{0,1}\stackrel{d_H}{}\mathrm{}\stackrel{d_H}{}\mathrm{\Omega }_{\mathrm{}}^{0,n}\stackrel{d_H}{}0.$$ (2) Then a corollary of Theorem 1 is the algebraic Poincaré lemma on the local exactness of the complex (2). PROPOSITION 2. If a fibre bundle $`Y`$ is $`𝐑^{n+m}𝐑^n`$, the horizontal complex has the cohomology groups $$H^{k<n}(d_H)=0,H^n(d_H)=\epsilon _1(\mathrm{\Omega }_{\mathrm{}}^{0,n}).$$ (3) This work is devoted to the obstruction to the (global) exactness of the complexes (1) and (2). It should be emphasized that we consider the variational bicomplex in the class of smooth exterior forms $`\mathrm{\Omega }_{\mathrm{}}^{}`$ on the infinite-order jet space $`J^{\mathrm{}}Y`$ which are the pull-back of exterior forms on finite-order jets. This is not the case of . In these works smooth forms on $`J^{\mathrm{}}Y`$ which are locally the pull-back of forms on finite-order jet manifolds are considered. The key point is that $`J^{\mathrm{}}Y`$ admits a partition of unity with respect to this class of smooth functions. Then, applying the well-known Mayer-Vietoris sequence , one can show that contact $`d_H`$-closed forms are necessarily $`d_H`$-exact. This may not be true for the class of exterior forms $`\mathrm{\Omega }_{\mathrm{}}^{}`$ which usually are utilized in field theory. We show that, if a fibre bundle $`YX`$ admits a global section (e.g., if it is an affine bundle), there is a monomorphism of the De Rham homology groups $`H^{}(X)`$ of the base manifold $`X`$ to the cohomology groups $`H_{\mathrm{var}}^{}`$ of the variational complex (1) and to those $`H^{}(d_H)`$ of the horizontal complex (2). When $`YX`$ is a vector bundle, we find the obstruction to the exactness of the variational complex (1) at $`E_k`$. We show that, in this case, the cohomology group of the variational complex (1) at $`\mathrm{\Omega }_{\mathrm{}}^{0,n}`$ is equal to the cohomology group $`H^n(X)`$ of the base manifold. It states the following theorem (see also ). THEOREM 3. If $`YX`$ is an affine bundle, an $`r`$-order Lagrangian $`L\mathrm{\Omega }_r^{0,n}`$ is variationally trivial iff it takes the form $`L=d_H\sigma `$ where $`\sigma \mathrm{\Omega }_{r1}^{0,n1}`$ is a horizontal $`(n1)`$-form on the jet manifold $`J^{r1}Y`$. Recall that the familiar result states that $`L`$ is locally $`d_H`$-exact. ## 2 Preliminaries Unless otherwise stated, manifolds throughout are assumed to be real, smooth, finite-dimensional, Hausdorff, paracompact and connected. Let $`J^rY`$ be the $`r`$-order jet manifold of sections of a fibre bundle $`YX`$. It is endowed with the adapted coordinates $`(x^\lambda ,y_\mathrm{\Lambda }^i)`$, $`0\mathrm{\Lambda }r`$, where $`\mathrm{\Lambda }=(\lambda _k\mathrm{}\lambda _1)`$, $`|\mathrm{\Lambda }|=k`$, is a symmetric multi-index. By $`\mathrm{\Lambda }+\mathrm{\Sigma }`$ we will denote the symmetric multi-index $`(\lambda _k\mathrm{}\lambda _1\sigma _m\mathrm{}\sigma _1)`$. The transition functions of these coordinates read $$y_{}^{}{}_{\lambda +\mathrm{\Lambda }}{}^{i}=\frac{x^\mu }{x^\lambda }d_\mu y_\mathrm{\Lambda }^i,$$ (4) where by $`d_\lambda `$ are meant the total derivatives $`d_\lambda =_\lambda +{\displaystyle \underset{|\mathrm{\Lambda }|=0}{}}y_{\mathrm{\Lambda }+\lambda }^i_i^\mathrm{\Lambda }.`$ We will use the notation $`_\mathrm{\Lambda }=_{\lambda _k}\mathrm{}_{\lambda _1},d_\mathrm{\Lambda }=d_{\lambda _k}\mathrm{}d_{\lambda _1},\mathrm{\Lambda }=(\lambda _k\mathrm{}\lambda _1).`$ There is the inverse system $$X\stackrel{\pi }{}Y\stackrel{\pi _0^1}{}\mathrm{}J^{r1}Y\stackrel{\pi _{r1}^r}{}J^rY\mathrm{}$$ (5) of fibrations (surjective submersions) of finite order jet manifolds. This inverse system has a projective limit $`J^{\mathrm{}}Y`$. It is a paracompact Fréchet manifold which admits a smooth partition of unity . Differential objects on $`J^{\mathrm{}}Y`$ are introduced as operations on the $`𝐑`$-ring $`C^{\mathrm{}}(J^{\mathrm{}}Y`$) of smooth locally cylindrical functions. A real function $`f`$ on $`J^{\mathrm{}}Y`$ is said to be a smooth (locally cylindrical) function if it is locally a pull-back of a smooth function on some finite-order jet manifold $`J^kY`$ with respect to the surjection $`\pi _k^{\mathrm{}}:J^{\mathrm{}}YJ^kY`$. Vector fields on $`J^{\mathrm{}}Y`$ are defined as derivations of this ring. They make up the left locally free $`C^{\mathrm{}}(J^{\mathrm{}}Y)`$-module Der$`(C^{\mathrm{}}(J^{\mathrm{}}Y))`$. The $`C^{\mathrm{}}(J^{\mathrm{}}Y)`$-dual of Der$`(C^{\mathrm{}}(J^{\mathrm{}}Y))`$ is the $`C^{\mathrm{}}(J^{\mathrm{}}Y)`$-module of exterior 1-forms on $`J^{\mathrm{}}Y`$. In field theory, one usually considers the direct limit $`\mathrm{\Omega }_{\mathrm{}}^{}`$ in the category of $`𝐑`$-modules of the direct system $$\mathrm{\Omega }^{}(X)\stackrel{\pi ^{}}{}\mathrm{\Omega }^{}(Y)\stackrel{\pi _0^1^{}}{}\mathrm{\Omega }_1^{}\stackrel{\pi _1^2^{}}{}\mathrm{}\stackrel{\pi _{r1}^r^{}}{}\mathrm{\Omega }_r^{}\mathrm{}.$$ (6) of the $`𝐑`$-modules $`\mathrm{\Omega }_k^{}`$ of exterior forms on finite-order jet manifolds $`J^kY`$. It consists of pull-backs of exterior forms on finite-order jet manifolds. The $`𝐑`$-module $`\mathrm{\Omega }_{\mathrm{}}^{}`$ possesses the structure of a filtered module given by the $`\mathrm{\Omega }_k^0`$-submodules $`\mathrm{\Omega }_k^{}`$ of $`\mathrm{\Omega }_{\mathrm{}}^{}`$ . An endomorphism $`\gamma `$ of $`\mathrm{\Omega }_{\mathrm{}}^{}`$ is called a filtered morphism if there exists $`i𝐍`$ such that the restriction of $`\gamma `$ to $`\mathrm{\Omega }_k^{}`$ is the homomorphism of $`\mathrm{\Omega }_k^{}`$ into $`\mathrm{\Omega }_{k+i}^{}`$ over the injection $`\mathrm{\Omega }_k^0\mathrm{\Omega }_{k+i}^0`$ for all $`k`$. THEOREM 4. . Every direct system of endomorphisms $`\{\gamma _k\}`$ of $`\mathrm{\Omega }_k^{}`$ such that $`\pi _k^j{}_{}{}^{}\gamma _k=\gamma _j\pi _k^j^{}`$ for all $`j>k`$ has the direct limit $`\gamma _{\mathrm{}}`$ in filtered endomorphisms of $`\mathrm{\Omega }_{\mathrm{}}^{}`$. If all $`\gamma _k`$ are monomorphisms \[epimorphisms\], then $`\gamma _{\mathrm{}}`$ is also a monomorphism \[epimorphism\]. By virtue of this theorem, the standard algebraic operations on exterior forms on finite-order jet manifolds have the direct limits on $`\mathrm{\Omega }_{\mathrm{}}^{}`$, and make $`\mathrm{\Omega }_{\mathrm{}}^{}`$ an exterior algebra. Since $`\mathrm{\Omega }_{\mathrm{}}^{}`$ consists of pull-back forms, its elements can be written in the familiar coordinate form. In particular, the basic 1-forms $`dx^\lambda `$ and the contact 1-forms $`\theta _\mathrm{\Lambda }^i=dy_\mathrm{\Lambda }^iy_{\lambda +\mathrm{\Lambda }}^idx^\lambda `$ provide the local generators of $`\mathrm{\Omega }_{\mathrm{}}^{}`$, and define the canonical splitting of the space of $`m`$-forms $$\mathrm{\Omega }_{\mathrm{}}^m=\mathrm{\Omega }_{\mathrm{}}^{0,m}\mathrm{\Omega }_{\mathrm{}}^{1,m1}\mathrm{}\mathrm{\Omega }_{\mathrm{}}^{m,0},$$ (7) in the spaces $`\mathrm{\Omega }^{k,mk}`$ of $`k`$-contact forms. We have the corresponding $`k`$-contact projections $`h_k:\mathrm{\Omega }_{\mathrm{}}^m\mathrm{\Omega }_{\mathrm{}}^{k,mk},1km,`$ and the horizontal projection $`h_0:\mathrm{\Omega }_{\mathrm{}}^m\mathrm{\Omega }_{\mathrm{}}^{0,m}.`$ Accordingly, the exterior differential on $`\mathrm{\Omega }_{\mathrm{}}^{}`$ is decomposed into the sum $`d=d_H+d_V`$ of horizontal and vertical differentials $`d_H:\mathrm{\Omega }_{\mathrm{}}^{k,s}\mathrm{\Omega }_{\mathrm{}}^{k,s+1},d_H(\varphi )=dx^\lambda d_\lambda (\varphi ),\varphi \mathrm{\Omega }_{\mathrm{}}^{},`$ $`d_V:\mathrm{\Omega }_{\mathrm{}}^{k,s}\mathrm{\Omega }_{\mathrm{}}^{k+1,s},d_V(\varphi )=\theta _\mathrm{\Lambda }^i_\mathrm{\Lambda }^i\varphi .`$ They obey the nilpotency rule $$d_Hd_H=0,d_Vd_V=0,d_Vd_H+d_Hd_V=0$$ (8) and the relations $`h_0d=d_Hh_0,`$ (9) $`h_kdh_k=d_Hh_k,`$ (10) ## 3 Cohomology of the infinite-order De Rham complex The exterior algebra $`\mathrm{\Omega }_{\mathrm{}}^{}`$ provides the infinite-order De Rham complex $$0𝐑\mathrm{\Omega }_{\mathrm{}}^0\stackrel{d}{}\mathrm{\Omega }_{\mathrm{}}^1\stackrel{d}{}\mathrm{}.$$ (11) PROPOSITION 5. The cohomology $`H^{}(\mathrm{\Omega }_{\mathrm{}}^{})`$ of the infinite-order De Rham complex (11) coincides with the De Rham cohomology $`H^{}(Y)`$ of the fibre bundle $`Y`$. The result is a corollary of the following two assertions. PROPOSITION 6. . The operation of taking cohomology groups of a cochain complex commutes with the passage to the direct limit. It is a corollary of Theorem 2. PROPOSITION 7. . The De Rham cohomology $`H^{}(J^kY)`$ of any finite-order jet manifold $`J^kY`$ is equal to that of the fibre bundle $`Y`$. It is based on the fact that jet bundles $`J^kYJ^{k1}Y`$ are affine, while the De Rham cohomology of an affine bundle is equal to that of its base. Recall that, given a fibre bundle $`\pi :YX`$, there is a homomorphism $$\pi ^{}:H^{}(X)H^{}(Y)$$ (12) of the De Rham cohomology groups of its base $`X`$ to those of $`Y`$. LEMMA 8. If a fibre bundle $`YX`$ admits a global section $`s`$, there is the sequence of homomorphisms of cohomology groups $$H^{}(X)\stackrel{\pi ^{}}{}H^{}(Y)\stackrel{s^{}}{}H^{}(X),$$ (13) where $`s^{}`$ is an epimorphism and, consequently, $`\pi ^{}`$ (12) is a monomorphism. If $`YX`$ admits a global section, we have the corresponding monomorphism $$\pi ^{}:H^{}(X)H^{}(\mathrm{\Omega }_{\mathrm{}}^{})$$ (14) of the De Rham cohomology groups of the base $`X`$ to the cohomology groups of the infinite-order De Rham complex (11). If $`YX`$ is an affine bundle, the monomorphism (14) is an isomorphism $$H^{}(\mathrm{\Omega }_{\mathrm{}}^{})=H^{}(X).$$ (15) In this case, any closed form $`\varphi \mathrm{\Omega }_{\mathrm{}}^{}`$ is decomposed into the sum $$\varphi =\phi +d\xi $$ (16) where $`\phi \mathrm{\Omega }^{}(X)`$ is a closed form on $`X`$. ## 4 Cohomology of the horizontal complex Due to the nilpotency rule (8), the vertical and horizontal differentials $`d_V`$ and $`d_H`$ define the bicomplex $$\begin{array}{ccccccccc}& & & \text{}\hfill & & & \text{}\hfill & & \\ \hfill \mathrm{}& & & \mathrm{\Omega }_{\mathrm{}}^{k+1,m}\hfill & \stackrel{d_H}{}& & \mathrm{\Omega }_{\mathrm{}}^{k+1,m+1}\hfill & & \hfill \mathrm{}\\ & & \hfill _{d_V}& \text{}\hfill & & \hfill _{d_V}& \text{}\hfill & & \\ \hfill \mathrm{}& & & \mathrm{\Omega }_{\mathrm{}}^{k,m}\hfill & \stackrel{d_H}{}& & \mathrm{\Omega }_{\mathrm{}}^{k,m+1}\hfill & & \hfill \mathrm{}\\ & & & \text{}\hfill & & & \text{}\hfill & & \end{array}$$ (17) The rows and columns of these bicomplex are horizontal and vertical complexes. Let us start from the vertical ones. PROPOSITION 9. If $`YX`$ is a vector bundle, then any $`d_V`$-closed form $`\varphi `$ is the sum $`\varphi =d_V\sigma +\phi `$ of a $`d_V`$-exact form and an exterior form on $`X`$, i.e., vertical complexes are exact. Proof. Local exactness of a vertical complex on a coordinate chart $`(x^\lambda ,y^i)`$ follows from a version of the Poincaré lemma with parameters. We have the the corresponding homotopy operator $$\sigma =_0^1t^k[\overline{y}\varphi (x^\lambda ,ty_\lambda ^i)]dt,\varphi \mathrm{\Omega }^{k,},$$ (18) where $`\overline{y}=y_\mathrm{\Lambda }^i_i^\mathrm{\Lambda }`$. Since $`YX`$ is a vector bundle, it is readily observed that the homotopy operator is globally defined, and so is the form $`\sigma `$. $`\mathrm{}`$ Turn now to the rows of the diagram (17). We will start from the horizontal complex (2). Due to the relation (9), the horizontal projection $`h_0`$ defines a homomorphism of the infinite-order De Rham complex (11) to the horizontal complex (2) which sends closed and exact forms to $`d_H`$-closed and $`d_H`$-exact horizontal forms, respectively. Accordingly, it yields the homomorphism of the cohomology groups of these complexes $$h_0^{}:H^{}(\mathrm{\Omega }_{\mathrm{}}^{})H^{}(d_H),$$ (19) which is neither monomorphism nor epimorphism in general. By virtue of Proposition 3, the homomorphism (19) is the homomorphism of the cohomology groups $$h_0^{}:H^{}(Y)H^{}(d_H).$$ (20) Then we also have the homomorphism $$h_0^{}\pi ^{}:H^{}(X)H^{}(d_H).$$ (21) PROPOSITION 10. If a fibre bundle $`YX`$ admits a global section $`s`$, the homomorphism (21) is a monomorphism. Proof. Bearing in mind the monomorphism (14), let $`\varphi \mathrm{\Omega }_{\mathrm{}}^{}`$ be a closed form whose cohomology class belongs to $`H^{}(X)H^{}(\mathrm{\Omega }_{\mathrm{}}^{})`$. We will show that, if $`h_0\varphi `$ is $`d_H`$-exact, $`\varphi `$ is exact. The form $`\varphi `$ is represented by the sum (16), and we have $`h_0\varphi =\phi +d_H(h_0\xi ).`$ Let $`\phi =d_H\sigma `$ where $`\sigma `$ is an exterior form on some finite-order jet manifold $`J^kY`$. Given a global section $`s`$ of the fibre bundle $`YX`$, let $`J^{k+1}s`$ be its $`(k+1)`$-order jet prolongation to a section of the jet bundle $`J^{k+1}YX`$. Then it is readily observed that $`\phi =(J^{k+1}s)^{}(d_H\sigma )=(J^{k+1}s)^{}(d\sigma )=d((J^{k+1}s)^{}\sigma ).`$ $`\mathrm{}`$ Since $`h_0`$ is a homomorphism of the infinite-order De Rham complex (11) to the horizontal complex (2), we have the simple exact sequence of the cochain complexes $$\begin{array}{cccccccccccccccc}& & & 0\hfill & & & 0\hfill & & & & & 0\hfill & & & & \\ & & & \text{}\hfill & & & \text{}\hfill & & & & & \text{}\hfill & & & & \\ \hfill \mathrm{}& \stackrel{d}{}& & \mathrm{Ker}h_0\hfill & \stackrel{d}{}& & \mathrm{Ker}h_0\hfill & \stackrel{d}{}& \mathrm{}& \stackrel{d}{}& & \mathrm{Ker}h_0\hfill & \stackrel{d}{}& \mathrm{Ker}h_0\hfill & & \\ & & \hfill _{\mathrm{in}}& \text{}\hfill & & \hfill _{\mathrm{in}}& \text{}\hfill & & & & \hfill _{\mathrm{in}}& \text{}\hfill & & ||\hfill & & \\ \hfill \mathrm{}& \stackrel{d}{}& & \mathrm{\Omega }_{\mathrm{}}^{k1}\hfill & \stackrel{d}{}& & \mathrm{\Omega }_{\mathrm{}}^k\hfill & \stackrel{d}{}& \mathrm{}& \stackrel{d}{}& & \mathrm{\Omega }_{\mathrm{}}^n\hfill & \stackrel{d}{}& \mathrm{\Omega }_{\mathrm{}}^{n+1}\hfill & & \mathrm{}\\ & & \hfill _{h_0}& \text{}\hfill & & \hfill _{h_0}& \text{}\hfill & & & & \hfill _{h_0}& \text{}\hfill & & \text{}\hfill & & \\ \hfill \mathrm{}& \stackrel{d_H}{}& & \mathrm{\Omega }_{\mathrm{}}^{0,k1}\hfill & \stackrel{d_H}{}& & \mathrm{\Omega }_{\mathrm{}}^{0,k}\hfill & \stackrel{d_H}{}& \mathrm{}& \stackrel{d_H}{}& & \mathrm{\Omega }_{\mathrm{}}^{0,n}\hfill & & 0\hfill & & \\ & & & \text{}\hfill & & & \text{}\hfill & & & & & \text{}\hfill & & & & \\ & & & 0\hfill & & & 0\hfill & & & & & 0\hfill & & & & \end{array}$$ (22) Its first row is a subcomplex of the infinite-order De Rham complex (11). There is the corresponding exact sequence of cohomology groups $$\mathrm{}H^k(\mathrm{Ker}h_0)\stackrel{\mathrm{in}^{}}{}H^k(\mathrm{\Omega }_{\mathrm{}}^{})\stackrel{h_0^{}}{}H^k(d_H)\stackrel{d^{}}{}H^{k+1}(\mathrm{Ker}h_0)\mathrm{}.$$ (23) We can say something more on this exact sequence when $`YX`$ is an affine bundle. PROPOSITION 11. If $`YX`$ is an affine bundle, the exact sequence of cohomology groups (23) takes the form $`\mathrm{}H^{k1}(d_H)/H^{k1}(X)\stackrel{\mathrm{in}^{}}{}H^k(X)\stackrel{h_0^{}}{}H^k(d_H)\stackrel{d^{}}{}H^k(d_H)/H^k(X)`$ (24) $`\mathrm{}\stackrel{\mathrm{in}^{}}{}H^n(X)\stackrel{h_0^{}}{}H^n(d_H)\stackrel{d^{}}{}H^n(d_H)/H^n(X)0,`$ where $`\mathrm{in}^{}`$ is a homomorphism to $`0H^{}(X)`$ and $`h_0^{}`$ is a monomorphism. Proof. Firstly, let us describe the cohomology group $`H^k(\mathrm{Ker}h_0)`$ of the first row in the diagram (22). In this complex, the cocycles are closed forms $`\varphi \mathrm{Ker}h_0`$, while the coboundaries are exact forms $`d\xi `$ where $`\xi \mathrm{Ker}h_0`$. Because of the isomorphism (15), such a $`k`$-cocycle is a coboundary in the infinite-order De Rham complex and is decomposed into the sum $$\varphi =d\sigma +d\xi ,$$ (25) where $`\sigma \mathrm{\Omega }_{\mathrm{}}^{0,k1}`$ is a horizontal form such that $`h_0d\sigma =0`$. It follows that $`d_H\sigma =0`$, i.e., $`\sigma `$ is a $`d_H`$-closed form. At the same time, if $`\sigma `$ is a $`d_H`$-exact form $`\sigma =d_H\psi `$, then $`d\sigma =dd_H\psi =dd_V\psi `$ is a coboundary in $`\mathrm{Ker}h_0`$. Therefore, we have an epimorphism of cohomology groups $`d^{}:H^{k1}(d_H)H^k(\mathrm{Ker}h_0).`$ Obviously, its kernel $`\mathrm{Ker}d^{}`$ contains the cohomology group $`H^{k1}(X)H^{k1}(d_H)`$. We show that $`\mathrm{Ker}d^{}=H^{k1}(X)`$. Let $`\sigma \mathrm{Ker}d^{}`$, i.e., $`d_H\sigma =0,d\sigma =\psi ,\psi \mathrm{Ker}h_0.`$ Then we have $`d(\sigma \psi )=0`$, and $`\sigma =\phi +d\xi +\psi ,\phi \mathrm{\Omega }^{}(X),`$ in accordance with the decomposition (16). Moreover, being horizontal, $`\sigma =\phi +h_0d\xi =\phi +d_Hh_0\xi ,`$ i.e., it belongs to the $`d_H`$-cohomology class of the form $`\phi `$ on $`X`$. At last, since all cocycles in $`\mathrm{Ker}h_0`$ are coboundaries in $`\mathrm{\Omega }_{\mathrm{}}^{}`$, the morphism $`\mathrm{in}^{}`$ sends the cohomology group $`H^k(\mathrm{Ker}h_0)`$ to zero. $`\mathrm{}`$ Turn now to cohomology of the $`k`$-contact horizontal complex $$\mathrm{}\stackrel{d_H}{}\mathrm{\Omega }_{\mathrm{}}^{k,m1}\stackrel{d_H}{}\mathrm{\Omega }_{\mathrm{}}^{k,m}\stackrel{d_H}{}\mathrm{}.$$ (26) It is readily observed that, because of the relation (9), the contact projection $`h_k`$ fails to be a homomorphism of the infinite-order De Rham complex (11) to the horizontal complex (26). Accordingly, we have no homomorphism of the corresponding cohomology groups. At the same time, due to the nilpotency rule (8), the horizontal differential yields a homomorphism of the $`k`$-contact horizontal complex (26) to the $`(k+1)`$-contact one, together with the corresponding homomorphism of their cohomology groups $`d_V^{}:H^{}(k,d_H)H^{}(k+1,d_H).`$ Moreover, we have the complex of the cohomology groups $$0H^{}(X)\stackrel{h_0^{}}{}H^{}(d_H)\stackrel{d_V^{}}{}\mathrm{}\stackrel{d_V^{}}{}H^{}(k,d_H)\stackrel{d_V^{}}{}\mathrm{}.$$ (27) Its cohomology is called the second cohomology . Similarly, the horizontal differential $`d_H`$ defines homomorphisms of the vertical complexes. Let us show that, if the fibre bundle $`YX`$ admits a global section, the complex (27) is exact at the second term, i.e. the kernel of $`d_V`$ in $`H^{}(d_H)`$ coincides with the De Rham cohomology $`H^{}(X)`$. LEMMA 12. If $`\varphi \mathrm{Ker}d_Hd_V`$, then $$\varphi =\sigma +\xi +\phi ,$$ (28) where $`\sigma `$ is a $`d_H`$-closed form, $`\xi `$ is a $`d_V`$-closed form and $`\phi \mathrm{\Omega }^{}(X)`$ (see for the local case). PROPOSITION 13. If the fibre bundle $`YX`$ admits a global section, the complex (27) is exact at the second term. Proof. Let $`\varphi \mathrm{\Omega }_{\mathrm{}}^{0,}`$ be a $`d_H`$-closed form, i.e., $`d_H\varphi =0`$. It belongs to $`\mathrm{Ker}d_V^{}`$ if $`d_V\varphi =d_H\epsilon `$. In accordance with the relation (28), it follows that $`\varphi =d_H\epsilon +\phi `$, i.e. belongs to $`H^{}(X)H^{}(d_H)`$. $`\mathrm{}`$ ## 5 Cohomology of the variational complex Obviously, the cohomology groups $`H_{\mathrm{var}}^{<n}`$ of the variational complex (1) coincide with those the horizontal complex (2). To say something on other cohomology groups $`H_{\mathrm{var}}^n`$, let us consider the simple exact sequence of cochain complexes $$\begin{array}{ccccccccccccccc}& & & 0\hfill & & & 0\hfill & & & 0\hfill & & & 0\hfill & & \\ & & & \text{}\hfill & & & \text{}\hfill & & & \text{}\hfill & & & \text{}\hfill & & \\ \hfill \mathrm{}& & & \mathrm{Ker}h_0\hfill & \stackrel{d}{}& & \mathrm{Ker}h_0\hfill & \stackrel{d}{}& & \mathrm{Ker}e_1\hfill & \stackrel{d}{}& & \mathrm{Ker}e_2\hfill & & \mathrm{}\hfill \\ & & & \text{}\hfill & & & \text{}\hfill & & & \text{}\hfill & & & \text{}\hfill & & \\ \hfill \mathrm{}& & & \mathrm{\Omega }_{\mathrm{}}^{n1}\hfill & \stackrel{d}{}& & \mathrm{\Omega }_{\mathrm{}}^n\hfill & \stackrel{d}{}& & \mathrm{\Omega }_{\mathrm{}}^{n+1}\hfill & \stackrel{d}{}& & \mathrm{\Omega }_{\mathrm{}}^{n+2}\hfill & & \mathrm{}\hfill \\ & & \hfill _{h_0}& \text{}\hfill & & \hfill _{h_0}& \text{}\hfill & & \hfill _{e_1}& \text{}\hfill & & \hfill _{e_2}& \text{}\hfill & & \\ \hfill \mathrm{}& & & \mathrm{\Omega }_{\mathrm{}}^{0,n1}\hfill & \stackrel{d_H}{}& & \mathrm{\Omega }_{\mathrm{}}^{0,n}\hfill & \stackrel{\epsilon _1}{}& & E_1\hfill & \stackrel{\epsilon _2}{}& & E_2\hfill & & \mathrm{}\hfill \\ & & & \text{}\hfill & & & \text{}\hfill & & & \text{}\hfill & & & \text{}\hfill & & \\ & & & 0\hfill & & & 0\hfill & & & 0\hfill & & & 0\hfill & & \end{array}$$ (29) where $`e_k=\tau _kh_k`$ and $`\tau _k`$ is the projection map providing the decomposition $`\mathrm{\Omega }_{\mathrm{}}^{k,n}=\tau _k(\mathrm{\Omega }_{\mathrm{}}^{k,n})+d_H(\mathrm{\Omega }_{\mathrm{}}^{k,n1}).`$ We have the corresponding decomposition $$\mathrm{\Omega }_{\mathrm{}}^{n+k}=\mathrm{Ker}e_kE_k,$$ (30) where $$\mathrm{Ker}e_k=d_H(\mathrm{\Omega }_{\mathrm{}}^{k,n1})(\underset{m>0}{}\mathrm{\Omega }^{k+m,nm}).$$ (31) It is readily observed that $`\epsilon _k=\tau _kd`$ on $`\mathrm{\Omega }_{\mathrm{}}^{k1,n}`$. The diagram (29) is derived from the variational bicomplex (see, e.g., ). Its first row is a subcomplexes of the infinite-order De Rham complex (11). Since the diagram (29) coincides with the diagram (22) at exterior forms of degree $`n`$, the corresponding exact sequence of cohomology groups for the diagram (29) differs from the exact sequence (23) starting from the terms after $`H^n(\mathrm{\Omega }_{\mathrm{}}^n)`$. If $`YX`$ is an affine bundle, then $`H^{>n}(\mathrm{\Omega }_{\mathrm{}}^{})=0`$ and the exact sequence at these terms breaks into the short sequences $`0H^n(X)\stackrel{h_0^{}}{}H_{\mathrm{var}}^n\stackrel{d^{}}{}H(\mathrm{Ker}e_1)\stackrel{\mathrm{in}^{}}{}0,`$ (32) $`0H_{\mathrm{var}}^{n+k}=H(e_{k+1})0,k>0.`$ (33) Using these exact sequences and the definition (31) of $`\mathrm{Ker}e_k`$, one can find the cohomology groups $`H_{\mathrm{var}}^n`$ of the variational complex (1). Recall that cocycles in $`\mathrm{Ker}e_k`$ are closed forms $`\varphi \mathrm{Ker}e_k`$, while the coboundaries are exact forms $`d\xi `$ where $`\xi \mathrm{Ker}e_{k1}`$. LEMMA 14. The cohomology group $`H(e_1)`$ is trivial. Proof. The obstruction for the cohomology group $`H(e_1)`$ to be trivial are the elements of the overlap $`d\mathrm{\Omega }_{\mathrm{}}^{0,n}d_H\mathrm{\Omega }_{\mathrm{}}^{k,n1}`$, i.e., the forms $`\varphi \mathrm{\Omega }_{\mathrm{}}^{0,n}`$ such that $`d_V\varphi =d_H\epsilon `$, i.e., $`\varphi \mathrm{Ker}d_Hd_V`$. By virtue of Lemma 4, such a form is given by the sum (28) which reads $`\varphi =d_H\psi +\phi .`$ Then, $`d\varphi =dd_V\psi `$ is an exact form. $`\mathrm{}`$ It follows that the short exact sequence (32) leads to the isomorphism $$H^n(X)=H_{\mathrm{var}}^n.$$ (34) In other words, the cohomology group of variationally trivial Lagrangians on an affine bundle modulo $`d_H`$-exact forms coincides with the cohomology group $`H^n(X)`$ of the base manifold $`X`$. This states Theorem 1. The isomorphism (34) also leads to the isomorphism $`H^n(d_H)/H^n(X)=\epsilon _1(\mathrm{\Omega }_{\mathrm{}}^{0,n}),`$ where $`H^n(d_H)`$ is the $`n`$th-cohomology group of the horizontal complex (2). It generalizes the local equality (3). One can extend Lemma 5 to higher cohomology groups $`H(e_k)`$ as follows. PROPOSITION 15. The obstruction for the cohomology group $`H(e_{k+1})`$ and, consequently, the cohomology group $`H_{\mathrm{var}}^{n+k}`$ to be trivial is the second cohomology group of the complex (27) at the term $`H^k(d_H)`$. The result follows from the decomposition (30).
warning/0004/cond-mat0004094.html
ar5iv
text
# Absence of residual quasiparticle conductivity in the underdoped cuprate YBa2Cu4O8 ## Abstract We report here measurements of the in-plane thermal conductivity $`\kappa `$ of the stoichiometric underdoped cuprate YBa<sub>2</sub>Cu<sub>4</sub>O<sub>8</sub> (Y124) below 1K. $`\kappa `$($`T`$) is shown to follow a simple, phononic $`T^3`$ dependence at the lowest $`T`$ for both current directions, with a negligible linear, quasiparticle contribution. This observation is in marked contrast with behavior reported in optimally doped cuprates, and implies that extended zero-energy (or low-energy) quasiparticles are absent in Y124. Experimental evidence for $`d`$-wave superconductivity in high-$`T_c`$ cuprates is now well established. The presence of nodes in the gap is expected to produce a finite density of well-defined quasiparticle (QP) excitations at low energies that dominate the low-$`T`$ physics. For a $`\mathrm{𝑝𝑢𝑟𝑒}`$ $`d`$-wave superconductor with line nodes on the Fermi surface, for example, the density of states (DOS) is linear in the excitation energy, giving rise to a $`T^2`$ dependence of the low-$`T`$ specific heat and thermal conductivity (assuming a constant scattering rate). This excitation spectrum, however, is altered significantly in the presence of impurities. For an anisotropic two-dimensional (2D) superconductor with scattering in the unitary limit, a band of impurity states is expected to develop whose width $`\gamma `$ grows with increasing impurity concentration $`n_{imp}`$, leading to a finite zero-energy DOS. Lee showed that the residual conductivity is independent of $`n_{imp}`$, the result of compensation between the increased DOS and a reduction in the associated transport lifetime. This residual, or ”universal” conductivity develops at low $`T`$ in the so-called ”dirty” limit, $`k_BT\gamma `$. Similarly, the low-$`T`$ $`\kappa (T)`$ will be dominated by QP states in the vicinity of the line nodes, and is given by the expression $$\kappa _{res}/T(n/d)(k_B^2/3)(v_F/v_2)$$ (1) where $`n/d`$ is the stacking density of CuO<sub>2</sub> planes and $`v_F`$ and $`v_2`$ are the energy dispersions (QP velocities) perpendicular and tangential to the Fermi surface respectively. The issue of localization of electronic states in $`d`$-wave superconductors, however, is a complex problem and many competing viewpoints prevail. Balatsky and Salkola, for example, argue that the long-range nature of hopping between impurity states along the nodal directions leads to strong overlap of the impurity wave functions along the diagonals of the square lattice, and ultimately to ”extended” impurity states. Senthil and co-workers, on the other hand, argue that quantum interference effects destabilize the extended QP states and lead to a vanishing DOS at zero energy (See also for a similar conclusion). They coined the term ”superconducting insulator” to describe such a superconductor with localized states. Experimentally, the observation of a finite linear term in $`\kappa (T)`$ in pure and Zn-doped YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7-δ</sub> (Y123) by Taillefer $`\mathrm{𝑒𝑡}`$ $`\mathrm{𝑎𝑙}.`$ appeared to confirm the existence of zero-energy quasiparticles. Moreover, the size of this term was indeed found to be ”universal”, i.e. independent of Zn concentration, in agreement with Lee’s prediction. Similar behavior was later reported for Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> (Bi2212) by Behnia $`\mathrm{𝑒𝑡}`$ $`\mathrm{𝑎𝑙}.`$, using irradiated crystals. More recently, the magnitude of $`\kappa _{res}/T`$ in Bi2212 ($``$ 0.15 mW/cm.K<sup>2</sup>) was shown to be consistent with absolute values of $`v_F/v_2`$ estimated from angle-resolved photoemission (ARPES). Despite this apparent consistency between theory and experiment, $`\kappa _{res}/T`$ has only been reported for two compounds, both at their optimum doping level, and it is not immediately obvious how $`\kappa _{res}/T`$ will vary across the phase diagram. ARPES and penetration depth measurements support claims that $`v_F/v_2`$, and thereby $`\kappa _{res}/T`$, increase in the underdoped regime. On the other hand, in certain underdoped cuprates, where $`T_c`$ has been suppressed in high magnetic fields, there is a marked tendency towards localization below $`T_c`$, suggesting a $`\mathrm{𝑣𝑎𝑛𝑖𝑠ℎ𝑖𝑛𝑔}`$ QP contribution at low $`T`$. Clearly, low-$`T`$ $`\kappa (T)`$ measurements on underdoped cuprates are important to help clarify this seemingly contradictory behavior. With this in mind, we have carried out the first low-$`T`$ $`\kappa (T)`$ measurements on the underdoped cuprate Y124 ($`T_c`$ = 80K), which is a self-doped, stoichiometric cuprate and therefore relatively free of disorder. Below 0.25K, $`\kappa (T)T^3`$ for both $`a`$\- and $`b`$-axis currents, consistent with a phonon heat conduction in the ballistic regime. The residual linear QP term, however, is either absent or is negligibly small, with an upper bound of 0.02 mW/cm.K<sup>2</sup>. This result reveals that the universal conductivity scenario breaks down dramatically in underdoped Y124. One compelling possibility is that the QP states in Y124 are localized at low $`T`$, due to the proximity to the superconductor/insulator (S/I) boundary, and therefore do not contribute to the low-$`T`$ heat transport. The Y124 crystals were grown by a flux method described elsewhere. For this particular study, three plate-like crystals were selected, two with their longest dimension along the $`b`$-axis, the other along the $`a`$-axis. Approximate dimensions were 0.25 x 0.16 x 0.015 mm<sup>3</sup> for the $`\kappa _a`$ crystal (labelled hereafter as a$`\mathrm{}`$1) and 1 x 0.09 x 0.05 mm<sup>3</sup> and 0.8 x 0.25 x 0.06 mm<sup>3</sup> for the two $`\kappa _b`$ crystals, b$`\mathrm{}`$1 and b$`\mathrm{}`$2. $`T_c`$ = 80K for all crystals, with a transition width, measured resistively, of less than 1K. $`\kappa (T)`$ for each crystal was measured between 0.14K and 1K using a conventional steady-state four-probe technique that allowed the electrical resistivity $`\rho _{a,b}(T)`$ of each sample to be measured in situ without changing the contact configuration. Gold wires were attached as electrical contacts using Dupont 6838 silver paint. The $`\rho _{a,b}(T)`$ behavior was found to be in excellent agreement with previous measurements, with room temperature values, $`\rho _a`$ = 350 $`\mu \mathrm{\Omega }`$cm and $`\rho _b`$ = 90 $`\mu \mathrm{\Omega }`$cm (for both $`b`$-axis crystals). This large in-plane anisotropy arises from the high conductivity of carriers on the quasi-1D CuO chains that run parallel to the $`b`$-axis (see schematic inset to Figure 1) and confirms not only the high quality of the crystals used in this study, but also that current flow in each case is uniaxial. The temperature gradient was measured by two RuO<sub>2</sub> thermometers connected to the ”voltage” contacts through the gold wires, and the thermometers were supported by long, thin superconducting Nb-Ti wires to minimize heat losses. Uncertainties in the absolute magnitudes of $`\kappa (T)`$ ($`\rho (T)`$), mainly due to the finite contact dimensions on these small crystals, are estimated to be around 15$`\%`$ for $`\kappa _b`$ ($`\rho _b`$) and around 25$`\%`$ for $`\kappa _a`$ ($`\rho _a`$). The $`\kappa (T)`$ data for all crystals are shown on double-logarithmic axes in Figure 1. The variation of $`\kappa _b(T)`$ for the two $`b`$-axis crystals is almost identical over the whole temperature range studied, giving us confidence in the reproducibility of our data. Below 0.25K, $`\kappa _a`$ and $`\kappa _b`$ both vary approximately as $`T^3`$, consistent with phonon heat transport in the boundary-scattering limit. Above 0.25K, $`\kappa _a(T)`$ deviates more strongly from a $`T^3`$ dependence. The origin of the enhancement of $`\kappa _b`$ over $`\kappa _a`$ above $`T`$ = 0.25K is not understood at present, though we assume it reflects an additional contribution to $`\kappa `$ from the CuO chains; either QP conductivity on the chains develops swiftly above 0.25K (note that the CuO chains, being quasi-1D, may be susceptible to charge ordering at very low $`T`$), or there exists an additional channel for phonon heat propagation along the chains that reduces the effects of phonon scattering beyond the ballistic regime. Further measurements in a magnetic field are envisaged to clarify the origin of this anisotropy. In order to look for evidence of a linear $`\kappa _{res}`$, we have re-plotted the low $`T`$ data in Figure 2 as $`\kappa _{a,b}/T`$ versus $`T^2`$ and fitted each data set below 0.25K to the expression $`\kappa _{a,b}=AT+BT^3`$. The coefficients for each fit are $`A`$ = 0.011, - 0.007 and 0.006 ($`\pm `$ 0.02) mW/cm.K<sup>2</sup> and $`B`$ = 6.50, 7.00 and 10.67 ($`\pm `$ 0.50) mW/cm.K<sup>4</sup> for a$`\mathrm{}`$1, b$`\mathrm{}`$1 and b$`\mathrm{}`$2 respectively. In the boundary-scattering limit, $`\kappa _{ph}`$ is given by $$\kappa _{ph}=1/3\beta <v_{ph}>l_0T^3$$ (2) where $`\beta `$ is the phonon specific heat coefficient, $`<v_{ph}>`$ the average acoustic sound velocity and $`l_0=2w/\sqrt{\pi }`$ is the maximum phonon mean free path. Here, $`w`$ represents a mean width of the rectangular-shaped crystal. Taking the dimensions of our crystals and suitable values for $`\beta `$ (= 0.5 $`\pm `$ 0.1 mJ/mol.K<sup>4</sup>) and $`<v_{ph}>`$ (= 5 $`\pm `$ 1 x 10<sup>5</sup> cm/s), we obtain estimates for $`\kappa _{ph}`$/$`T^3`$ = 4.1 $`\pm `$ 1.2, 5.25 $`\pm `$ 1.5 and 9.55 $`\pm `$ 2.0 mW/cm.K<sup>4</sup> for a$`\mathrm{}`$1, b$`\mathrm{}`$1 and b$`\mathrm{}`$2 respectively. Given the uncertainties in measuring dimensions and contact distances, we believe these values compare favourably with the experimental values. More importantly, the size of the $`T^3`$ term for the two $`b`$-axis crystals scales well with $`w`$ and we conclude that the $`T^3`$ contribution is indeed simply the phonon contribution in the ballistic regime. The most striking result here is the complete absence (to within our experimental accuracy) of the residual linear term in the low-$`T`$ $`\kappa (T)`$ for $`\mathrm{𝑏𝑜𝑡ℎ}`$ chain and plane current directions. It should be emphasized, of course, that a zero linear term in $`\kappa _b`$ also implies a negligible $`\kappa _{res}`$ within the planes, meaning we have effectively confirmed the absence of the universal QP term in Y124 in all three samples. Moreover, for there to be any finite zero-$`T`$ intercept in $`\kappa /T`$, it would require $`\kappa _{ph}(T)`$ below 0.15K to vary as $`T^{\mathit{3}+n}`$ with $`n`$ $`>`$ 0, which is simply not physical, given that the lattice heat capacity is strictly cubic below 1K. Thus, we are confident that the main result of this Letter, namely the absence of $`\kappa _{res}`$ in Y124, is robust. For comparison, we also show in the inset to Figure 2, $`\kappa _{ab}/T`$ for optimally doped Bi2212 $`\mathrm{𝑚𝑒𝑎𝑠𝑢𝑟𝑒𝑑}`$ $`\mathrm{𝑤𝑖𝑡ℎ}`$ $`\mathrm{𝑡ℎ𝑒}`$ $`\mathrm{𝑠𝑎𝑚𝑒}`$ $`\mathrm{𝑒𝑥𝑝𝑒𝑟𝑖𝑚𝑒𝑛𝑡𝑎𝑙}`$ $`\mathrm{𝑠𝑒𝑡}`$-$`\mathrm{𝑢𝑝}`$. In Bi2212, we can clearly distinguish a finite $`\kappa _{res}/T`$ 0.15 mW/cm.K<sup>2</sup> (shown by a dotted line), that is $`\mathrm{𝑎𝑛}`$ $`\mathrm{𝑜𝑟𝑑𝑒𝑟}`$ $`\mathrm{𝑜𝑓}`$ $`\mathrm{𝑚𝑎𝑔𝑛𝑖𝑡𝑢𝑑𝑒}`$ $`\mathrm{𝑙𝑎𝑟𝑔𝑒𝑟}`$ than the upper limits for $`\kappa _{res}/T`$ in Y124. Despite the overwhelming case for $`d`$-wave pairing in high-$`T_c`$ cuprates, there is still limited, direct evidence for a $`d_{x^2y^2}`$ order parameter in Y124. Thus, before discussing our result in terms of nodal QP states, we should first examine the possibility that there is a finite gap everywhere on the Fermi surface in Y124. First of all, the orthorhombic distortion in Y124, induced by the chains, introduces some $`d`$ \+ $`s`$ admixture in the gap function. As the $`s`$-component is increased from zero, the position of the nodal lines are first shifted away from ($`\pi ,\pi `$), but as the $`s`$-component becomes comparable with the $`d`$-component, a nodeless gap may form. Secondly, magnetic impurities are thought to induce a local $`\mathrm{𝑖𝑚𝑎𝑔𝑖𝑛𝑎𝑟𝑦}`$ component that could also give rise to a fully gapped state and a suppression of low $`T`$ thermal transport. This latter possibility, however, is not supported by specific heat data taken on crystals from similar batches to those studied here, which show no sign of a low $`T`$ Schottky anomaly arising from such magnetic impurities. In addition, power law penetration depths have now been observed down to 2K for both the $`a`$\- and $`b`$-axes in Y124, suggesting a simple nodal gap picture is equally applicable to Y124. Of course, we cannot rule out completely the possibility of a finite gap in Y124, but if it does exist, it would have to be vanishingly small. In what follows, therefore, we assume that nodal lines are present in Y124 and turn to consider what might be happening to the low-energy QP states in their vicinity. From (1), we see that $`\kappa _{res}/T`$ is directly proportional to the ratio $`v_F`$/$`v_2`$ at the nodal positions, so a negligible $`\kappa _{res}/T`$ may indicate a sharp gap feature at the nodes at very low energies, induced either by doping, impurities or structural modifications. However, as noted above, $`v_F/v_2`$ is expected to $`\mathrm{𝑖𝑛𝑐𝑟𝑒𝑎𝑠𝑒}`$ as we move into the underdoped regime, and indeed, independent estimates of $`v_F/v_2`$ from the slope of the low-$`T`$ penetration depth in Y124 yield an estimate for $`\kappa _{res}/T`$ that is $`\mathrm{𝑙𝑎𝑟𝑔𝑒𝑟}`$ than those for both optimally doped Bi2212 and Y123. Moreover, given that band structure estimates for $`v_F`$ are similar for Y123 and Y124, even with our upper bound estimate for $`\kappa _{res}/T`$ (= 0.02 mW/cm.K<sup>2</sup>), we obtain a physically unrealistic value ($`v_F/v_2`$ $``$ 2) for the gap slope within the nodes in Y124. It appears unlikely, therefore, that the gap structure can itself explain a value of $`\kappa _{res}/T`$ one order of magnitude lower than in Y123. Another important consideration is the size of the impurity band, $`\gamma `$. We recall that in the unitary limit, the universal conductivity regime develops below $`k_BT\gamma `$. Thus, the observation of $`\kappa _{res}`$ depends not only on the temperature range of the experiment, but also on the energy scale of the impurity band, imposed by the scattering phase shift. However, taking values for the $`b`$-axis residual resistivity $`\rho _0`$ = 0.5 $`\mu \mathrm{\Omega }`$cm and plasma frequency $`\omega _p`$ = 2.5eV, we obtain a $`\mathrm{𝑙𝑜𝑤𝑒𝑟}`$ $`\mathrm{𝑏𝑜𝑢𝑛𝑑}`$ estimate of $`\gamma `$ in the unitary limit of 14K, some two orders higher than the base temperature of our measurements. (Unfortunately, similar analysis for $`J//a`$ cannot be performed due to the difficulties in estimating $`\rho _0`$ from $`\rho _a(T)`$). Even in the opposite (Born) limit, where $`\gamma `$ becomes exponentially small and therefore, may fall below our measurement range, the product of the DOS and the lifetime is also energy independent above $`k_BT\gamma `$. Hence, $`\kappa _{res}/T`$ should still be constant and finite and the same arguments still apply. Given these simple yet rather compelling arguments, we are left to consider how QP localization might account for the absence of $`\kappa _{res}/T`$ in underdoped Y124. As mentioned above, high magnetic field measurements on the underdoped cuprates La<sub>2-x</sub>Sr<sub>x</sub>CuO<sub>4</sub> and Pr<sub>2-x</sub>Ce<sub>x</sub>CuO<sub>4</sub> revealed that $`\rho _{ab}(T)`$, though ”metallic” at high $`T`$, tends towards localization as $`T`$ 0K. The origin of this localization phenomenon is unknown at present. However, if the field-induced destruction of superconductivity leads directly to an insulating phase, then it is not unreasonable to assume that this transition is from a superconductor with already $`\mathrm{𝑙𝑜𝑐𝑎𝑙𝑖𝑧𝑒𝑑}`$ QP excitations. The crossover from metallic to insulating behavior in the normal state (i.e. above $`H_{c2}`$) as we approach the Mott insulator, suggests an increasing role of long-range interactions on the mobility of low-energy quasiparticles. The superfluid condensate, suppressed in the vicinity of an impurity, becomes ineffective in screening completely the Coulomb repulsion between quasiparticles in the bound state. As we approach the parent insulator, we expect such interactions to grow, leading to insulating behavior of the quasiparticles above the superconducting condensate and a negligible QP contribution to the low-$`T`$ heat transport. Such localization in a nominally clean superconductor is an exciting prospect, and measurements on Zn-doped or irradiated Y124 are envisaged to investigate this possibility further. We note here that most impurity models for $`d`$-wave (cuprate) superconductivity fail to take into account the developing role of long-range interactions between quasiparticles as the S/I boundary is approached. We hope, therefore, that this work stimulates renewed theoretical efforts to understand the nature of QP excitations in the CuO<sub>2</sub> planes, deep inside the superconducting state, and in particular in stoichiometric crystals on the underdoped side of the phase diagram. In conclusion, we have measured the low-$`T`$ $`\kappa (T)`$ of stoichiometric, underdoped Y124 and have found that, in marked contrast to optimally doped Y123 and Bi2212, the ”universal” QP conductivity term is absent. We have considered several interpretations of this intriguing result, including localization of the quasiparticles themselves, due to enhanced long-range interactions as we approach the S/I boundary. Prior to this work, the observation of the universal conductivity in Y123 and Bi2212 had been widely regarded as solid support for the picture of long-lived quasiparticles above the superconducting ground state of high-$`T_c`$ cuprates. Our surprising result offers important and timely counter evidence that the residual conductivity term is non-universal, and we hope it encourages further debate and investigation into this critical, and still controversial, issue. We acknowledge enlightening discussions with A.S. Alexandrov, J.F. Annett, A.V. Balatsky, D.M. Broun, J.R. Cooper, P.J. Hirschfeld, F.V. Kusmartsev, A.P. Mackenzie, A.J. Schofield, L. Taillefer, I. Vekhter and V.W. Wittorff. S.N. acknowledges support from the National Science Foundation under Grant No.INT-9901436. This work was also partly supported by CREST, a Grant in Aid for Scientific Research from the Ministry of Education, Science and Culture, Japan and New Energy and Industrial Technology Department Organization (NEDO). Figure Captures Fig.1. $`\kappa _a(T)`$ and $`\kappa _b(T)`$ of Y124, plotted on double-logarithmic axes. The dashed line represents the $`T^3`$ dependence expected for phonon heat transport in the ballistic regime. The inset shows a simple schematic of the crystal structure of Y124. Fig.2. $`\kappa /T`$ versus $`T^2`$ below 0.4K for a$`\mathrm{}`$1 (closed circles), b$`\mathrm{}`$1 (open circles) and b$`\mathrm{}`$2 (open squares). Fits to the expression $`\kappa =AT+BT^3`$ below 0.25K are indicated by dashed lines for b$`\mathrm{}`$1 and b$`\mathrm{}`$2 and by a solid line for a$`\mathrm{}`$1. The dotted line represents the universal conductivity limit for Bi2212. Inset: Comparison of $`\kappa _a/T`$ for Y124 (closed circles) and $`\kappa _{ab}/T`$ of Bi2212 (open diamonds), measured in the same apparatus.
warning/0004/hep-th0004137.html
ar5iv
text
# Symmetry content of a generalized 𝑝-form model of Schwarz-type in 𝑑 dimensions ## 1 General setup BF models and their properties are a widely investigated area in the literature . In particular in ref. BF models in arbitrary dimensions are considered in great detail. In ref. a conclusive formalism in order to derive the minimal action in the framework of Batalin and Vilkovisky (BV) is presented. Based on these concepts ref. introduces a possible generalization to more generic models of Schwarz-type. In this paper we present the BRST-variations, the vector supersymmetry (VSUSY) as well as a scalar supersymmetry, denoted as Ł-symmetry, of the proposed model. The classical, invariant action of such a generalized $`p`$-form model in $`d`$ space-time dimensions is give by<sup>1</sup><sup>1</sup>1In the following we omit the wedge product sign $``$. $`S_{inv}={\displaystyle __d}\left\{B_pF_2+{\displaystyle \underset{i}{}}X_{p_i}DY_{r_i}\right\},`$ (1) where $`p=d2`$, $`r_i=dp_i1`$ and $`_i`$ denotes the summation over an arbitrary set of pairs of $`p_i`$\- and $`r_i`$-form fields. $`F_2=dA+AA`$ is the usual curvature of the connection one-form $`A`$ and the covariant derivative is $`D=d+[A,.]`$. All commutators $`[.,.]`$ are understood in a graded sense. The fields $`B_p`$, $`X_{p_i}`$ and $`Y_{r_i}`$ exhibit more or less reducible gauge symmetries, e. g. $`\delta _{\lambda _{p1}}B_p=D\lambda _{p1}`$. The algebra of gauge-transformations closes on-shell. The gauge-invariant equations of motion imply the zero-curvature conditions. Following the guideline of we define pairs of generalized forms which are called dual to each other<sup>2</sup><sup>2</sup>2Upper indices label the ghost-number and lower ones denote the form-degree. $$\begin{array}{cccccc}\hfill \stackrel{~}{B}_p& =& B_d^2+B_{d1}^1+B_p+\mathrm{}+B^p,\hfill & \hfill \stackrel{~}{A}& =& A_d^{1d}+\mathrm{}+A_2^1+A+c,\hfill \\ \hfill \stackrel{~}{X}_{p_i}& =& X_d^{p_id}+\mathrm{}+X_{p_i}+\mathrm{}+X^{p_i},\hfill & \hfill \stackrel{~}{Y}_{r_i}& =& Y_d^{r_id}+\mathrm{}+Y_{r_i}+\mathrm{}+Y^{r_i}.\hfill \end{array}$$ For later convenience we cast all fields of e. g. $`\stackrel{~}{X}_{p_i}`$ with negative Faddeev-Popov ($`\mathrm{\Phi }\mathrm{\Pi }`$) charge into $`\stackrel{ˇ}{X}_{p_i}`$, whereas the contributions with positive or zero $`\mathrm{\Phi }\mathrm{\Pi }`$-charge are collected in $`\widehat{X}_{p_i}`$. Hence, a total expansion is $`\stackrel{~}{X}_{p_i}=\stackrel{ˇ}{X}_{p_i}+\widehat{X}_{p_i}`$. The denotation of dual fields becomes more evident if one observes that the fields with negative $`\mathrm{\Phi }\mathrm{\Pi }`$-charge, serve as antifields in the sense of Batalin and Vilkovisky for the elements of the dual partner with positive ghost-degree, i. e. $`\stackrel{ˇ}{X}_{p_i}=\pm (\widehat{Y}_{r_i})^{}`$, and vice versa<sup>3</sup><sup>3</sup>3The superscript denotes antifields.. The classical gauge and ghost fields can be addressed by $`\mathrm{\Phi }^a`$, whereas the corresponding antifields are collected in $`\mathrm{\Phi }_a^{}`$. ## 2 BRST-symmetry and BV-action With the definition of a generalized exterior derivative $`\stackrel{~}{d}=d+s`$, with $`s`$ denoting the BRST-differential, we can define a generalized covariant derivative $`\stackrel{~}{D}=\stackrel{~}{d}+[\stackrel{~}{A},.]`$ and construct the curvatures $`\stackrel{~}{F}_2=\stackrel{~}{d}\stackrel{~}{A}+\stackrel{~}{A}\stackrel{~}{A},\stackrel{~}{G}_{p+1}=\stackrel{~}{D}\stackrel{~}{B}_p,\stackrel{~}{H}_{p_i+1}=\stackrel{~}{D}\stackrel{~}{X}_{p_i},\stackrel{~}{I}_{r_i+1}=\stackrel{~}{D}\stackrel{~}{Y}_{r_i}.`$ (2) The BRST-transformations of the fields are determined through so-called horizontality conditions which read $$\begin{array}{cccc}\stackrel{~}{F}_2=0,\hfill & \stackrel{~}{G}_{p+1}=_i(1)^{p_i}[\stackrel{~}{X}_{p_i},\stackrel{~}{Y}_{r_i}],\hfill & \stackrel{~}{H}_{p_i+1}=0,\hfill & \stackrel{~}{I}_{r_i+1}=0.\hfill \end{array}$$ (3) With the definitions $`F_2^{\stackrel{~}{A}}=d\stackrel{~}{A}+\stackrel{~}{A}\stackrel{~}{A}`$ and $`D^{\stackrel{~}{A}}=d+[\stackrel{~}{A},.]`$ this yields $$\begin{array}{cccccc}\hfill s\stackrel{~}{A}& =& F_2^{\stackrel{~}{A}},\hfill & \hfill s\stackrel{~}{B}_p& =& D^{\stackrel{~}{A}}\stackrel{~}{B}_p+_i(1)^{p_i}[\stackrel{~}{X}_{p_i},\stackrel{~}{Y}_{r_i}],\hfill \\ \hfill s\stackrel{~}{X}_{p_i}& =& D^{\stackrel{~}{A}}\stackrel{~}{X}_{p_i},\hfill & \hfill s\stackrel{~}{Y}_{r_i}& =& D^{\stackrel{~}{A}}\stackrel{~}{Y}_{r_i},\hfill \end{array}$$ The above BRST-transformations admit the cocycle equation $`\stackrel{~}{d}\mathrm{tr}\left(\stackrel{~}{B}_pF_2^{\stackrel{~}{A}}+{\displaystyle \underset{i}{}}\stackrel{~}{X}_{p_i}D^{\stackrel{~}{A}}\stackrel{~}{Y}_{r_i}\right)=0.`$ (4) This leads to the BRST-invariant, minimal BV-action $`S_{min}={\displaystyle __d}\left\{\stackrel{~}{B}_pF_2^{\stackrel{~}{A}}+{\displaystyle \underset{i}{}}\stackrel{~}{X}_{p_i}D^{\stackrel{~}{A}}\stackrel{~}{Y}_{r_i}\right\}|_d^0.`$ (5) An expansion in the generic fields yields $`S_{min}`$ $`=`$ $`S_{inv}+{\displaystyle __d}\{\stackrel{ˇ}{B}_p(s\widehat{A})+\stackrel{ˇ}{A}(s\widehat{B}_p)+{\displaystyle \underset{i}{}}(\stackrel{ˇ}{X}_{p_i}(s\widehat{Y}_{r_i})+(1)^{d(p_i+1)}\stackrel{ˇ}{Y}_{r_i}(s\widehat{X}_{p_i}))`$ (6) $`+`$ $`{\displaystyle \underset{i}{}}(1)^{p_i}\stackrel{ˇ}{A}([\widehat{X}_{p_i},\stackrel{ˇ}{Y}_{r_i}]+[\stackrel{ˇ}{X}_{p_i},\widehat{Y}_{r_i}]){\displaystyle \frac{1}{2}}\widehat{B}_p[\stackrel{ˇ}{A},\stackrel{ˇ}{A}]\left\}\right|_d^0.`$ The latter action induces the antifield identification $`\stackrel{ˇ}{B}_p=(\widehat{A})^{},\stackrel{ˇ}{A}=(\widehat{B}_p)^{},\stackrel{ˇ}{X}_{p_i}=(\widehat{Y}_{r_i})^{},\stackrel{ˇ}{Y}_{r_i}=(1)^{d(p_i+1)}(\widehat{X}_{p_i})^{},`$ (7) in order to ensure coincidence with the BRST-transformations obtained from $`s\mathrm{\Phi }_a^{}={\displaystyle \frac{\delta S_{min}}{\delta \mathrm{\Phi }^a}},s\mathrm{\Phi }^a={\displaystyle \frac{\delta S_{min}}{\delta \mathrm{\Phi }_a^{}}}.`$ (8) ## 3 Gauge-fixing Similarly to the BF model we need to introduce for each classical gauge field a BV-pyramid (c. f. Table 2 and 2). From the gauge-fixing point of view the dual fields $`B_p`$ and $`A`$ can be considered as an ordinary $`p`$\- and one-form field. Hence, we define $`\stackrel{~}{X}_{d2}\stackrel{~}{B}_p`$ and $`\stackrel{~}{Y}_1\stackrel{~}{A}`$ and according definitions for the antighost and multiplier fields. Moreover, for the sake of a compact notation when giving the gauge-fermion, we let the antighost fields to the lowest order $`n`$ be the classical gauge and ghost fields, i. e. $`\overline{v}_{p_iq}^qX_{p_iq}^q`$ and $`\overline{w}_{r_iq}^qY_{r_iq}^q`$. The gauge-fixing fermion $`\mathrm{\Psi }_{gf}`$ then looks like $`\mathrm{\Psi }_{gf}`$ $`=`$ $`{\displaystyle __d}\{{\displaystyle \underset{i}{}}{\displaystyle \underset{n=1}{\overset{p_i}{}}}{\displaystyle \underset{q=0}{\overset{p_in}{}}}\overline{v}_{p_iqn}^{\gamma (n)}({}_{}{}^{n}\alpha _{(q)}^{}\kappa _{p_iqn}^{\gamma (n+1)}+d\overline{v}_{p_iqn+1}^{\gamma (n+1)})`$ (9) $`+`$ $`{\displaystyle \underset{i}{}}{\displaystyle \underset{n=1}{\overset{r_i}{}}}{\displaystyle \underset{q=0}{\overset{r_in}{}}}\overline{w}_{r_iqn}^{\gamma (n)}({}_{}{}^{n}\beta _{(q)}^{}\lambda _{r_iqn}^{\gamma (n+1)}+d\overline{w}_{r_iqn+1}^{\gamma (n+1)})\},`$ where $``$ is the Hodge-star operator, $`{}_{}{}^{n}\alpha _{(q)}^{},{}_{}{}^{n}\beta _{(q)}^{}𝐑`$ are some arbitrary parameters and $`\gamma (n=2k)=q`$ or $`\gamma (n=2k+1)=q1`$. The implementation of external sources $`\rho _a^{}`$ implies a further contribution to the gauge-fermion $`\mathrm{\Psi }_{ext}`$ $`=`$ $`{\displaystyle __d}{\displaystyle \underset{i}{}}(1)^{(d+1)p_i}\left({\displaystyle \underset{q=0}{\overset{p_i}{}}}(1)^{d+1}X_{p_iq}^q\tau _{dp_i+q}^{q1}+{\displaystyle \underset{q=0}{\overset{r_i}{}}}Y_{r_iq}^q\eta _{dr_i+q}^{q1}\right).`$ The total gauge-fermion is $`\mathrm{\Psi }=\mathrm{\Psi }_{gf}+\mathrm{\Psi }_{ext}`$. The corresponding multiplier fields to the antighosts $`\overline{v}_p^q`$ and $`\overline{w}_p^q`$ are $`\kappa _p^{q+1}`$ and $`\lambda _p^{q+1}`$ respectively. The auxiliary contribution is given by $`S_{aux}`$ $`=`$ $`{\displaystyle __d}{\displaystyle \underset{i}{}}\left({\displaystyle \underset{n=1}{\overset{p_i}{}}}{\displaystyle \underset{q=0}{\overset{p_in}{}}}\left(\overline{v}_{p_iqn}^{\gamma (n)}\right)^{}\kappa _{p_iqn}^{\gamma (n)+1}+{\displaystyle \underset{n=1}{\overset{r_i}{}}}{\displaystyle \underset{q=0}{\overset{r_in}{}}}\left(\overline{w}_{r_iqn}^{\gamma (n)}\right)^{}\lambda _{r_iqn}^{\gamma (n)+1}\right).`$ The BRST-doublet fields are collected in $`(\overline{C}^\alpha ,\mathrm{\Pi }^\alpha )`$ and together with the fields of $`\mathrm{\Phi }^a`$ they may by addressed by $`\mathrm{\Phi }^A`$. The non-minimal solution of the BV-masters equation is given by $`S_{nm}=S_{min}+S_{aux}`$. The antifields can be expressed as functionals of the fields via the equation $`\mathrm{\Phi }_A^{}[\mathrm{\Phi }^A]=(1)^{(d+1)|\mathrm{\Phi }^A|+d}{\displaystyle \frac{\delta \mathrm{\Psi }}{\delta \mathrm{\Phi }^A}}.`$ (10) This admits the elimination of the antifields of the action in order to get $`\mathrm{\Gamma }^{(0)}=S_{nm}|_{\mathrm{\Phi }_A^{}[\mathrm{\Phi }^A]}=S_{inv}+S_{gf}+S_{ext},`$ (11) where $`S_{gf}=s\mathrm{\Psi }_{gf}+S_{gf}^{mod}`$ and $`S_{ext}=s\mathrm{\Psi }_{ext}+S_{ext}^{mod}`$. The additional contribution $`S^{mod}=S_{gf}^{mod}+S_{ext}^{mod}`$ appears since the BRST-transformations of the fields (2) exhibit an antifield-dependency. The structure of $`S^{mod}`$ can be seen from the last line in (6). Although, $`S^{mod}`$ can not be written as a BRST-exact expression, it however, does not spoil the topological character of the model, due to its metric independence. With equation (10) the antifields can also be eliminated in the generalized field expansions (1). ## 4 Vector supersymmetry ### 4.1 Derivation The above concepts provide a neat formalism to derive the VSUSY , $`\delta _\tau =\tau ^\mu \delta _\mu ^1`$, where $`\tau `$ is a constant, BRST-invariant, even graded vector-field<sup>4</sup><sup>4</sup>4Henceforth $`_d`$ is considered as a flat space-time manifold.. The VSUSY-transformations satisfy the algebra $`[s,\delta _\tau ]=_\tau =[d,i_\tau ],`$ (12) where $`_\tau `$ is the Lie-derivative and $`i_\tau `$ is the interior product along $`\tau `$. The algebra (12) suggests an equivalence between $`i_\tau `$ and $`\delta _\tau `$, yielding for the $`\delta _\tau `$-transformations $`\delta _\tau \stackrel{~}{X}_{p_i}=i_\tau \stackrel{~}{X}_{p_i},\delta _\tau \stackrel{~}{Y}_{r_i}=i_\tau \stackrel{~}{Y}_{r_i}.`$ (13) This determines the VSUSY-transformations of the classical gauge and ghost fields but also of the antighost fields at the first reducibility level. However, in order to describe the $`\delta _\tau `$-transformations of the higher reducibility antighost fields we need to collect also the antighosts with positive $`\mathrm{\Phi }\mathrm{\Pi }`$-charge together with their corresponding antifields in generalized forms<sup>5</sup><sup>5</sup>5The reasoning is only given for the $`X_{p_i}`$-sector, but the same arguments hold for the antighost fields $`\overline{w}_{r_inq}^q`$. $`\widehat{\overline{v}}_{p_in}={\displaystyle \underset{q=0}{\overset{p_in}{}}}\overline{v}_{p_inq}^q,(\widehat{\overline{v}}_{p_in})^{}={\displaystyle \underset{q=0}{\overset{p_in}{}}}(\overline{v}_{p_inq}^q)^{},n=2k.`$ (14) Although only the antighosts with positive $`\mathrm{\Phi }\mathrm{\Pi }`$-charge are cast into this scheme, the antighost fields with negative ghost-degree come into play automatically via the elimination of the antifields (10). The VSUSY-transformations now follow from the proposed equivalence of the $`\delta _\tau `$-operation and the interior product $`i_\tau `$ in the space of generalized forms and their duals, thus we get $`\delta _\tau \widehat{\overline{v}}_{p_in}=i_\tau \widehat{\overline{v}}_{p_in},\delta _\tau (\widehat{\overline{v}}_{p_in})^{}=i_\tau (\widehat{\overline{v}}_{p_in})^{}.`$ (15) This determines the VSUSY-transformations of the remaining antighost fields, but also the gauge-parameters $`{}_{}{}^{n}\alpha _{(q)}^{}={}_{}{}^{n}\beta _{(q)}^{}=(1)^d`$ for $`n=2k`$. ### 4.2 Explicit results The detailed results for the elements of $`\mathrm{\Phi }^a`$ yield from (13) $$\begin{array}{cccc}\hfill \delta _\tau X_{p_i}& =& i_\tau \left(\eta _{p_i+1}^1(1)^{(d+1)p_i+d}d\overline{w}_{r_i1}^1\right),\hfill & \\ \hfill \delta _\tau X_{p_iq}^q& =& i_\tau X_{p_iq+1}^{q1},\hfill & q=1,\mathrm{},p_i,\hfill \\ \hfill \delta _\tau Y_{r_i}& =& i_\tau \left(\tau _{r_i+1}^1(1)^{d+p_i+1}d\overline{v}_{p_i1}^1\right),\hfill & \\ \hfill \delta _\tau Y_{r_iq}^q& =& i_\tau Y_{r_iq+1}^{q1},\hfill & q=1,\mathrm{},r_i.\hfill \end{array}$$ (16) The antighost field transformations are determined through (15) $$\begin{array}{cccc}\hfill \delta _\tau \overline{v}_{p_in}& =& 0,\hfill & \\ \hfill \delta _\tau \overline{v}_{p_iqn}^q& =& i_\tau \overline{v}_{p_iqn+1}^{q1},\hfill & q=1,\mathrm{},p_in,\hfill \\ \hfill \delta _\tau \overline{v}_{p_iqn}^{q1}& =& (1)^{d+p_i+q+1}g(\tau )\overline{v}_{p_iqn1}^{q2},\hfill & q=0,\mathrm{},p_in1,\hfill \\ \hfill \delta _\tau \overline{v}_0^{p_i+n1}& =& 0,\hfill & \end{array}$$ (17) and $$\begin{array}{cccc}\hfill \delta _\tau \overline{w}_{r_in}& =& 0,\hfill & \\ \hfill \delta _\tau \overline{w}_{r_iqn}^q& =& i_\tau \overline{w}_{r_iqn+1}^{q1},\hfill & q=1,\mathrm{},r_in,\hfill \\ \hfill \delta _\tau \overline{w}_{r_iqn}^{q1}& =& (1)^{d+r_i+q+1}g(\tau )\overline{w}_{r_iqn1}^{q2},\hfill & q=0,\mathrm{},r_in1,\hfill \\ \hfill \delta _\tau \overline{w}_0^{r_i+n1}& =& 0,\hfill & \end{array}$$ (18) where the Hodge-star $``$ intertwines between the interior product and the one-form $`g(\tau )=\tau _\mu dx^\mu `$ in the way $`i_\tau \alpha _p=(1)^pg(\tau )\alpha _p`$. The corresponding multiplier field transformations can be obtained from the algebra (12) through $`\delta _\tau \mathrm{\Pi }^\alpha =_\tau \overline{C}^\alpha s\delta _\tau \overline{C}^\alpha ,`$ (19) for some element of the BRST-doublets $`(\overline{C}^\alpha ,\mathrm{\Pi }^\alpha )`$. Furthermore, the $`\delta _\tau `$-variations of the external sources are $$\begin{array}{cccc}\hfill \delta _\tau \tau _{dp_i+q}^{q1}& =& i_\tau \tau _{dp_i+q+1}^{q2},\hfill & q=0,\mathrm{},p_i1,\hfill \\ \hfill \delta _\tau \tau _d^{p_i1}& =& 0,\hfill & \\ \hfill \delta _\tau \eta _{dr_i+q}^{q1}& =& i_\tau \eta _{dr_i+q+1}^{q2},\hfill & q=0,\mathrm{},r_i1,\hfill \\ \hfill \delta _\tau \eta _d^{r_i1}& =& 0.\hfill & \end{array}$$ (20) The algebra (12) of the VSUSY-transformations of the classical gauge fields closes on-shell $`[s,\delta _\tau ]X_{p_i}=_\tau X_{p_i}(1)^{d(p_i+1)}i_\tau {\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta Y_{r_i}}},[s,\delta _\tau ]Y_{r_i}=_\tau Y_{r_i}i_\tau {\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta X_{p_i}}}.`$ (21) In general we describe the symmetry content of a model with a Ward-operator $`𝒲^I={\displaystyle __d}{\displaystyle \underset{\phi }{}}\delta ^I\phi {\displaystyle \frac{\delta }{\delta \phi }},`$ (22) where $`\delta ^I\phi `$ denotes the field-transformations under the symmetry $`I`$ and $`\phi `$ stands for all fields characterizing the model in question. In this sense we define a Ward-operator $`𝒲_\tau `$ according to the above $`\delta _\tau `$-transformations (16)–(20). By choosing the remaining gauge-parameters $`{}_{}{}^{2k+1}\alpha _{(q)}^{}={}_{}{}^{2k+1}\beta _{(q)}^{}=0`$, the application of $`𝒲_\tau `$ onto $`\mathrm{\Gamma }^{(0)}`$ leads to a linear breaking term in the quantum fields $`𝒲_\tau \mathrm{\Gamma }^{(0)}=\mathrm{\Delta }_\tau ,`$ (23) where $`\mathrm{\Delta }_\tau `$ $`=`$ $`{\displaystyle __d}{\displaystyle \underset{i}{}}(1)^{p_i}\{(1)^{d+1}({\displaystyle \underset{q=0}{\overset{p_i}{}}}\tau _{dp_i+q}^{q1}_\tau X_{p_iq}^q+\kappa _{p_i1}di_\tau \eta _{p_i+1}^1)`$ (24) $`+`$ $`({\displaystyle \underset{q=0}{\overset{r_i}{}}}\eta _{dr_i+q}^{q1}_\tau Y_{r_iq}^q+\lambda _{r_i1}di_\tau \tau _{r_i+1}^1)\}.`$ ## 5 Ł-symmetry ### 5.1 General setup For the following section we assume that the action is complete in the sense, that it contains all possible types of $`p`$-forms that are allowed, but where no particular $`p`$-form shall occur more than once. The definitions $`I_1^{\stackrel{~}{Y}_0}=D^{\stackrel{~}{A}}\stackrel{~}{Y}_0`$, $`I_2^{\stackrel{~}{Y}_1}F_2^{\stackrel{~}{A}}`$, $`I_3^{\stackrel{~}{Y}_2}=D^{\stackrel{~}{A}}\stackrel{~}{Y}_2`$, …, $`I_r^{\stackrel{~}{Y}_{r1}}=D^{\stackrel{~}{A}}\stackrel{~}{Y}_{r1}`$ admit to write (5) as $`S_{min}={\displaystyle __d}\left\{{\displaystyle \underset{i=1}{\overset{r}{}}}\stackrel{~}{X}_{di}I_i^{\stackrel{~}{Y}_{i1}}\right\}|_d^0.`$ (25) The upper limit is given by $`r=d/2`$ for even or $`r=(d+1)/2`$ for odd dimensions. We define a scalar transformation Ł with $`\mathrm{\Phi }\mathrm{\Pi }`$-charge -1 with the algebra $`[\text{Ł},s]=0`$. The Ł-transformations act on the generalized fields as follows $$\begin{array}{cccc}\hfill \text{Ł}\stackrel{~}{X}_{di}& =& \stackrel{~}{X}_{di1},\hfill & i=1,\mathrm{},r1,\hfill \\ \hfill \text{Ł}\stackrel{~}{X}_{dr}& =& 0,\hfill & \\ \hfill \text{Ł}\stackrel{~}{Y}_0& =& 0,\hfill & \\ \hfill \text{Ł}\stackrel{~}{Y}_{i1}& =& (1)^{d+i}\stackrel{~}{Y}_{i2},\hfill & i=2,\mathrm{},r.\hfill \end{array}$$ (26) ### 5.2 Explicit results The elimination of the antifields via (10) yields the explicit transformation properties. The classical fields transform as $$\begin{array}{cccc}\hfill \text{Ł}X_{di}& =& \eta _{di}^1+(1)^{(d+1)i+1}d\overline{w}_{i1}^1,\hfill & i=1,\mathrm{},r1,\hfill \\ \hfill \text{Ł}X_{dr}& =& 0,\hfill & \\ \hfill \text{Ł}Y_0& =& 0,\hfill & \\ \hfill \text{Ł}Y_{i1}& =& (1)^{(d+1)i}\left(\tau _{i1}^1+(1)^{(d+1)i+d}d\overline{v}_{di}^1\right),\hfill & i=2,\mathrm{},r.\hfill \end{array}$$ (27) The ghosts vary under the Ł-symmetry like $$\begin{array}{ccccc}\hfill \text{Ł}X_{diq}^q& =& X_{diq}^{q1},\hfill & i=1,\mathrm{},r1,\hfill & q=1,\mathrm{},di,\hfill \\ \hfill \text{Ł}X_{drq}^q& =& 0,\hfill & & q=1,\mathrm{},dr,\hfill \\ \hfill \text{Ł}Y_{i1q}^q& =& (1)^{d+i}Y_{i1q}^{q1},\hfill & i=2,\mathrm{},r,\hfill & q=1,\mathrm{},i1.\hfill \end{array}$$ (28) The antighost fields transform as $$\begin{array}{ccccc}\hfill \text{Ł}\overline{v}_{din}& =& 0,\hfill & i=1,\mathrm{},r,\hfill & \\ \hfill \text{Ł}\overline{v}_{dinq}^q& =& \overline{v}_{dinq}^{q1},\hfill & i=1,\mathrm{},r1,\hfill & q=1,\mathrm{},din,\hfill \\ \hfill \text{Ł}\overline{v}_{drnq}^q& =& 0,\hfill & & q=1,\mathrm{},drn,\hfill \\ \hfill \text{Ł}\overline{v}_{d1nq}^{q1}& =& 0,\hfill & & q=0,\mathrm{},d1n\hfill \\ \hfill \text{Ł}\overline{v}_{dinq}^{q1}& =& (1)^{i+1}\overline{v}_{dinq}^{q2},\hfill & i=2,\mathrm{},r,\hfill & q=0,\mathrm{}din,\hfill \end{array}$$ (29) and $$\begin{array}{ccccc}\hfill \text{Ł}\overline{w}_{i1n}& =& 0,\hfill & i=1,\mathrm{},r,\hfill & \\ \hfill \text{Ł}\overline{w}_{i1nq}^q& =& (1)^{d+i}\overline{w}_{i1nq}^{q1},\hfill & i=2,\mathrm{},r,\hfill & q=1,\mathrm{}i1n,\hfill \\ \hfill \text{Ł}\overline{w}_{i1nq}^{q1}& =& \overline{w}_{i1nq}^{q2},\hfill & i=1,\mathrm{},r1,\hfill & q=0,\mathrm{}.i1n,\hfill \\ \hfill \text{Ł}\overline{w}_{r1nq}^{q1}& =& 0,\hfill & & q=0,\mathrm{},r1n.\hfill \end{array}$$ (30) Due to $`[\text{Ł},s]=0`$ the corresponding multiplier fields transform as $`\text{Ł}\mathrm{\Pi }^\alpha =s\text{Ł}\overline{C}^\alpha ,`$ (31) for some element of the BRST-doublets $`(\overline{C}^\alpha ,\mathrm{\Pi }^\alpha )`$. Finally, the Ł-variations of the external sources $`\rho _a^{}`$ are $$\begin{array}{ccccc}\hfill \text{Ł}\tau _{1+q}^{q1}& =& 0,\hfill & & q=0,\mathrm{},d2,\hfill \\ \hfill \text{Ł}\tau _{i+q}^{q1}& =& (1)^i\tau _{i+q}^{q2},\hfill & i=2,\mathrm{},r,\hfill & q=0,\mathrm{},di1,\hfill \\ \hfill \text{Ł}\eta _{di+1+q}^{q1}& =& \eta _{di+1+q}^{q2},\hfill & i=1,\mathrm{},r1,\hfill & q=0,\mathrm{},i2,\hfill \\ \hfill \text{Ł}\eta _{dr+1+q}^{q1}& =& 0,\hfill & & q=0,\mathrm{},r2.\hfill \end{array}$$ (32) The algebra $`[\text{Ł},s]=0`$ on the classical fields is valid on-shell $`[\text{Ł},s]X_{di}=(1)^{d(i+1)}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta Y_i}},[\text{Ł},s]Y_{i1}=(1)^{d+i}{\displaystyle \frac{\delta \mathrm{\Gamma }^{(0)}}{\delta X_{di+1}}}.`$ (33) The inclusion of the external sources also yields a linear breaking of the Ward-identity (c. f. (22)) associated to the latter presented Ł-transformations (27)–(32) $`𝒲^\text{Ł}\mathrm{\Gamma }^{(0)}=\mathrm{\Delta }^\text{Ł},`$ (34) where $`\mathrm{\Delta }^\text{Ł}`$ $`=`$ $`{\displaystyle __d}\left\{{\displaystyle \underset{i=1}{\overset{r1}{}}}(1)^{i+1}\kappa _{di1}d\eta _{di}^1+{\displaystyle \underset{i=2}{\overset{r}{}}}(1)^{d(i+1)}\lambda _{i2}d\tau _{i1}^1\right\}.`$ (35) The symmetry property of the above transformations can also be understood in another context. The action (25) can be seen as the dimensional reduction of a similar model in $`d+1`$ space-time dimensions. In this sense the Ł-symmetry is nothing else but the leftover of the surplus VSUSY-transformation in the reduced direction. Hence, it is obvious that Ł is indeed a symmetry transformation. ## 6 Conclusion In this short note we applied the procedure of to the case of an arbitrary topological $`p`$-form model of Schwarz-type in $`d`$ space-time dimensions . We presented the BRST-transformations in terms of generic form fields, and we extended the formalism to the derivation of the VSUSY-transformations and a scalar supersymmetry called Ł-symmetry. Obviously, the model under consideration incorporates an ordinary BF theory in arbitrary dimensions by setting all fields except $`B_pX_{d2}`$ and $`AY_1`$ to zero. As a three-dimensional application of the generic $`p`$-form model one can reconstruct the results for the so-called BFK model . With a similar method the authors of analyzed the BFK model enlarged by a Chern-Simons term. ## 7 Acknowledgements It is a great pleasure to thank Jesper Grimstrup for discussions and comments.
warning/0004/cond-mat0004338.html
ar5iv
text
# Statistical coarse-graining as an approach to multiscale problems in magnetism ## Abstract Multiscale phenomena which include several processes occuring simultaneously at different length scales and exchanging energy with each other, are widespread in magnetism. These phenomena often govern the magnetization reversal dynamics, and their correct modeling is important. In the present paper, we propose an approach to multiscale modeling of magnets, applying the ideas of coarse graining. We have analyzed the choice of the weighting function used in coarse graining, and propose an optimal form for this function. Simple tests provide evidence that this approach may be useful for modeling of realistic magnetic systems. A large number of phenomena taking place in magnets include processes occuring simultaneously at different length scales. A good example is the magnetization reversal in a macroscopic piece of magnetic material possessing different kinds of defects, voids, surfaces etc. The reversal starts by a nucleation of a domain with the magnetization opposite to the initial direction. As a rule, the nucleation happens near defects, where spins can be frustrated. Here, the different length scales involved can be clearly identified. First, there is the microscopic scale with a characteristic length of the order of several interatomic distances (several tens of angstroms), which corresponds to the region of spin frustration and contains the microstructure in the vicinity of the defect. Next, there is a “micromagnetic” length scale (of order of the domain wall width, several thousands of angstroms) where the formation of the general structure of the nucleus takes place. And, finally, the truly macroscopic length scale (of order of several microns or even millimeters), where the magnons created in the course of the reversal propagate. These magnons play an important role in energy transfer and sometimes can initiate the magnetization reversal in other areas of the sample . A similar picture of several interacting length scales appears in many situations, such as the breakthrough of a domain wall pinned by a defect , influence of the surface on the magnetic structure of the core of a magnetic particle etc. Processes of this type, called multiscale processes, are receiving considerable attention nowadays. Along with very interesting and rich physics, these are the very processes which govern the switching behavior of magnetic systems (coercive field, switching time, etc.), so that an adequate understanding of multiscale phenomena is of paramount importance for development and creation of new magnetic storage media. Micromagnetic simulations can provide a realistic description of the processes taking place at both micromagnetic and macroscopic length scales. On the other hand, microscopic inhomogeneities require atomistic simulations (e.g., spin dynamics modeling is an adequate tool for materials where spins are well localized at the sites of the crystalline lattice). However, for the description of real systems, all the length scales should be coupled, i.e. modeled simultaneously and seamlessly, with the possibility of energy transfer between them. A simple scheme, where the region of micromagnetic simulations is just attached to the region of atomic spin simulations, does not give a satisfactory description of the energy transfer between the lengthscales. The problem is that micromagnetics does not take into account atomistic degrees of freedom; even reduction of the computational mesh in micromagnetic modeling down to atomic scale does not describe the short-wavelength excitations properly, contrary to spin dynamics simulations. The artificial sharp boundary appearing between the micromagnetic and atomistic regions leads to the unphysical scattering of magnons transferring the energy from one region to the other. To couple length scales correctly, some transition region between the micromagnetic and atomic regions is necessary, which allows for gradual exclusion of short-wavelength modes until they die out in the regions far from the defect. Similar difficulties arise in modeling dynamics of structural defects in crystals (dislocation motion, crack propagation, etc.). The coupling of lengthscales for this class of problems has been extensively discussed in the literature . One of the most promising approaches to the solution of this problem is coarse-grained molecular dynamics (CGMD), developed in Ref. . In the present paper, we apply the ideas of CGMD to magnetic materials modeling. We propose a computational scheme which employs basic concepts of nonequilibrium statistical mechanics to couple micromagnetism and the dynamic modeling of classical spins. We identify the key problems arising in the course of implementation of this scheme and their possible solutions. To explain the idea of the method, let us consider a ferromagnet where some magnetic inhomogeneity (defect, interface, etc.) is present. We describe it as a system of classical magnetic moments $`𝐌_\mu `$ of fixed length $`M=|𝐌_\mu |`$, located at the $`\mu `$-th site of the crystalline lattice (greek indices enumerate the lattice sites). We assume that the system is described by the Hamilton function $``$ $`=`$ $`^0+𝒱,`$ (1) $`^0`$ $`=`$ $`{\displaystyle \underset{\mu }{}}_\mu ^0={\displaystyle \underset{\mu ,\nu }{}}J_{\mu \nu }𝐌_\mu 𝐌_\nu ,`$ (2) where $`^0`$ describes the isotropic exchange interaction, while $`𝒱^0`$ represents all the other, much weaker interactions present in the system, and $`J_{\mu \nu }`$ is the exchange integral between the sites $`\mu `$ and $`\nu `$. Let us focus on the region far from the defect, where the amplitudes of short-wavelength excitations are small (the region where their amplitudes are not small should be treated completely atomistically). Suppose, a computational grid is defined in this region, so that the magnetic state is described with required precision by defining the magnetization direction at the grid nodes: $`𝐌_j=(M_j^x,M_j^y,M_j^z)`$ at the $`j`$-th node (below, latin indices enumerate the nodes of the mesh). These data represent the necessary number of large-scale degrees of freedom, and exact modeling of all the atomic-scale modes would be too excessive (and too expensive resoursewise). However, the atomic-scale modes can not be excluded completely, since their cumulative effect (dissipation, energy transfer, etc.) is not negligible, and should be taken into account. This problem can be solved by invoking statistical physics, i.e. under the assumption of ergodicity the exact dynamics of short-scale modes can be replaced by their statistical distribution. For many relevant situations it has been shown that the short-scale modes relax almost immediately to the state of local quasiequilibrium determined uniquely by large-scale modes, so that knowledge of only the large-scale parameters (“gross variables”) determines the dynamics of the system with necessary rigor. This description of the system, where only large-scale modes are essential (while short-scale modes are “slaved” due to the requirement of the local equilibrium), is often referred to as the coarse-grained description. The theory of local quasiequilibrium states has been developed in the 1950s-1960s, and a number of approaches exist . In the following, we will use the convenient modification of the nonequilibrium statistical operator (NSO) method . For simplicity, we do not consider the dissipation processes and quantum spin effects, but in principle they can also be included following Ref. . It is convenient to introduce new dynamic variables $`\alpha _{1,\mu }`$ and $`\alpha _{2,\mu }`$ for the magnetic moments $`𝐌_\mu `$ according to the following relations: $`M_\mu ^x`$ $`=`$ $`\alpha _{1,\mu }\sqrt{2M\alpha _{1,\mu }^2\alpha _{2,\mu }^2},`$ (3) $`M_\mu ^y`$ $`=`$ $`\alpha _{2,\mu }\sqrt{2M\alpha _{1,\mu }^2\alpha _{2,\mu }^2},`$ (4) $`M_\mu ^z`$ $`=`$ $`M\alpha _{1,\mu }^2\alpha _{2,\mu }^2.`$ (5) These variables can be expressed via conventional polar $`\theta _\mu `$ and asimuthal $`\varphi _\mu `$ angles as $`\alpha _{1,\mu }`$ $`=`$ $`\sqrt{2M}\mathrm{sin}(\theta _\mu /2)\mathrm{cos}\varphi _\mu ,`$ (6) $`\alpha _{2,\mu }`$ $`=`$ $`\sqrt{2M}\mathrm{sin}(\theta _\mu /2)\mathrm{sin}\varphi _\mu .`$ (7) It can be checked that the variables $`\alpha _{1,\mu }`$ and $`\alpha _{2,\mu }`$ are canonically conjugated, so that Hamiltonian formalism can be used, and a computational scheme preserving the simplectic structure can be employed. The conventional way to develop a coarse-grained description of the system is to introduce the large-scale variables $`\alpha _j`$ as averages of $`\alpha _\mu `$: $$\alpha _{1,j}=\underset{\mu }{}f_{\mu ,j}\alpha _{1,\mu },\alpha _{2,j}=\underset{\mu }{}f_{\mu ,j}\alpha _{2,\mu },$$ (8) where $`f_{\mu ,j}`$ is an appropriate weighting function, satisfying the normalization condition $`_\mu f_{\mu ,j}=1`$, and localized near the node $`j`$; proper choice of the weighting function $`f_{\mu ,j}`$ will be discussed below in more detail. Averaging is a natural way to introduce gross variables for a ferromagnet, where the local equilibrium is governed by the Hamiltonian $`^0`$, so that all the magnetic moments $`𝐌_\mu `$ near the node $`j`$ are almost parallel to $`𝐌_j`$. Introduced in this way, $`\alpha _{(1,2),j}`$ constitute collective variables of the system changing slowly with respect to quickly relaxing short-scale modes, so they can be considered as quasi-integrals of motion (for more detailed discussion see Refs. ). Following standard procedures of statistical physics, the integrals of motion are included into the distribution function via Legendre transformations, i.e. we introduce Legendre multipliers $`F_j`$ and $`G_j`$ corresponding to the node variables $`\alpha _{1,j}`$ and $`\alpha _{2,j}`$, so that the distribution function can be written as $`\rho `$ $`=`$ $`Q^1\mathrm{exp}({\displaystyle \underset{\mu }{}}\beta _\mu _\mu ^0{\displaystyle \underset{j}{}}F_j{\displaystyle \underset{\mu }{}}\beta _\mu f_{\mu ,j}\alpha _{1,\mu }`$ (10) $`{\displaystyle \underset{j}{}}G_j{\displaystyle \underset{\mu }{}}\beta _\mu f_{\mu ,j}\alpha _{2,\mu }),`$ $`Q`$ $`=`$ $`{\displaystyle }\mathrm{exp}({\displaystyle \underset{\mu }{}}\beta _\mu _\mu ^0{\displaystyle \underset{j}{}}F_j{\displaystyle \underset{\mu }{}}\beta _\mu f_{\mu ,j}\alpha _{1,\mu }`$ (12) $`{\displaystyle \underset{j}{}}G_j{\displaystyle \underset{\mu }{}}\beta _\mu f_{\mu ,j}\alpha _{2,\mu }){\displaystyle }_\mu {\displaystyle \frac{d\alpha _{1,\mu }d\alpha _{2,\mu }}{2\pi M}},`$ where $`\beta _\mu =1/(k_BT_\mu )`$, $`k_B`$ is Boltzmann’s constant, $`T_\mu `$ is the spin temperature at the $`\mu `$-th site, and $`Q`$, as can be seen, is the normalization factor, analogous to the statistical sum of a canonical ensemble in the equilibrium case. Note that smooth variations of the temperature across the sample can be taken into account (temperature is a Legendre multiplier for the integral of energy), but to make the consideration simpler, we neglect it, assuming constant temperature, small enough to satisfy the condition $`\beta J_{\mu \nu }1`$. The variables $`F_j`$ and $`G_j`$ can be considered as parameters of a local fictitious field acting on the moments $`𝐌_\mu `$. Values of the parameters $`F_j`$ and $`G_j`$ should be chosen in such a way that the resulting equilibrium values $`\alpha _{(1,2),\mu }`$ averaged with the weighting function $`f_{\mu ,j}`$ give the required values of $`\alpha _{(1,2),j}`$. It is easy to see from Eq. 10 that the values $`F_j^0`$, $`G_j^0`$ producing the required values $`\alpha _{(1,2),j}`$ are determined from the equations $`{\displaystyle \frac{}{F_j}}|_{F_j^0,G_j^0}`$ $`=`$ $`\alpha _{1,j},{\displaystyle \frac{}{G_j}}|_{F_j^0,G_j^0}=\alpha _{2,j},`$ (13) $`(F_j,G_j)`$ $`=`$ $`\beta ^1\mathrm{ln}Q`$ (14) where $``$ is an analog of the Gibbs’ free energy function. Now, having all the information at hand, we can use averaging to get the equations of motion for the coarse-grained variables $`\alpha _{(1,2),j}`$. The underlying dynamics of the microscopic variables $`\alpha _{(1,2),\mu }`$ is Hamiltonian, i.e. $$\dot{\alpha }_{1,\mu }=\frac{}{\alpha _{2,\mu }},\dot{\alpha }_{2,\mu }=\frac{}{\alpha _{1,\mu }},$$ (15) where $``$ is the Hamilton function (1). Using the distribution function (10) with the values $`F_j^0`$ and $`G_j^0`$ determined from Eq. 13, we obtain: $`\dot{\alpha }_{1,j}`$ $`=`$ $`{\displaystyle \underset{\mu }{}}f_{\mu ,j}{\displaystyle \frac{𝒱}{\alpha _{2,\mu }}}{\displaystyle \underset{\mu }{}}f_{\mu ,j}{\displaystyle \underset{k}{}}f_{\mu ,k}G_k^0,`$ (16) $`\dot{\alpha }_{2,j}`$ $`=`$ $`{\displaystyle \underset{\mu }{}}f_{\mu ,j}{\displaystyle \frac{𝒱}{\alpha _{1,\mu }}}+{\displaystyle \underset{\mu }{}}f_{\mu ,j}{\displaystyle \underset{k}{}}f_{\mu ,k}F_k^0,`$ (17) where $`\mathrm{}`$ means averaging with the distribution (10). Note that the equations of motion are nonlocal over the node indices even if only local exchange interactions are present in the system; this important feature is totally missing in the micromagnetic description. Direct implementation of the scheme presented above can be rather expensive computationally. To make the problem easier, we can take into account that these computations are to be performed only inside the relatively narrow “belt” between micromagnetic and atomic-scale regions (see Fig. 1), where departures of magnetization from equilibrium are already small (otherwise, atomic-scale simulations should be used). If the $`z`$-axis of the co-ordinate frame is aligned with the equilibrium direction of magnetization the values of $`\alpha _{(1,2),\mu }`$ are small, and the Hamiltonian $`^0`$ can be expanded in terms of $`\alpha _{(1,2),\mu }`$ up to second order: $$^0=\frac{1}{2}\underset{u,v=1,2}{}\underset{\mu ,\nu }{}D_{u\mu ,v\nu }\alpha _{u,\mu }\alpha _{v,\nu },$$ (18) where the indices $`u,v=1,2`$ denote the variables $`\alpha _1`$ and $`\alpha _2`$. In this case, the distribution (10) is Gaussian, so that the integral $`Q`$ and the function $``$ in (13) can be calculated exactly . As a result, Eq. 13 determining the values of generalized torques $`F_j^0`$ and $`G_j^0`$ can be written as a set of linear equations: $$\alpha _{u,j}=\underset{k}{}\underset{v=1,2}{}T_{v,k}^0\underset{\mu ,\nu }{}f_{\mu ,j}D_{u\mu ,v\nu }^1f_{\nu ,k},$$ (19) where $`D_{u\mu ,v\nu }^1`$ is the inverse of the dynamic matrix $`D_{u\mu ,v\nu }`$ in the linearized exchange Hamiltonian (18), and we used the vector $`T_{v,k}^0`$ ($`v=1,2`$) to denote both $`F_k^0`$ and $`G_k^0`$ as follows: $`T_{1,k}^0F_k^0`$, and $`T_{2,k}^0G_k^0`$. Hence, the problem of dynamic coupling of length scales is reduced to two linear problems, Eqs. 19 and 16. Note that the rotation of the co-ordinate frame which brings the $`z`$-axis into coincidence with the equilibrium direction of magnetization makes the dynamic matrix independent of $`\alpha _{(1,2),j}`$, so that its inverse can be calculated once and stored for further references. However, there is a subtlety in inverting $`D_{u\mu ,v\nu }`$: this matrix, being determined only by the isotropic exchange interactions, has a zero eigenvalue corresponding to a shift of all $`\alpha _{(1,2),\mu }`$ by the same value, or, in other words, the dynamic matrix has an eigenvector $`d^0=(1,1,\mathrm{},1)`$ corresponding to the zero eigenvalue. It reflects the fact that the exchange Hamiltonian (18) is invariant with respect to rotation of the system as a whole. Thus, when inverting numerically the dynamic matrix, a component corresponding to the zero eigenvector $`d^0`$ should be excluded. In the present form, the essense of coarse-graining becomes especially clear. Imagine that, applying some fictitious torques $`F_j`$ and $`G_j`$ we brought the system into such a state that the equilibrium values of the large-scale variables are $`\alpha _{(1,2),j}`$. Then, the atomic magnetic moments $`𝐌_\mu `$ (i.e., the microscopic variables $`\alpha _{(1,2),\mu }`$) move in such a way that, after the stage of quick relaxation is finished, their positions minimize the total energy of the system with respect to the constraints imposed by the torques $`F_j`$, $`G_j`$. The last but not least problem is an appropriate choice of the weighting function $`f_{\mu ,j}`$: it can be verified that an arbitrary function does not automatically give meaningful results. To study this question, let us inspect closely the basic idea of the coarse-grained description. For a general system consisting of $`N`$ microscopic moments, an exact description of the system’s dynamics requires knowledge of all $`2N`$ microscopic canonical variables $`\alpha _{(1,2),\mu }`$. However, we expect that under certain conditions (which are yet to be formulated), the system can be described with reasonable precision using the much smaller set of $`2L`$ gross variables $`\alpha _{(1,2),j}`$. In other words, we expect that under certain approximations, we can define such a set of variables $`\alpha _{(1,2),j}`$ that allows specification of all microscopic variables $`\alpha _{(1,2),\mu }`$ with sufficient precision. In particular, it can be shown that, for a given weighting function $`f_{\mu ,j}`$, an optimal (in the least-square sense) restoration of microvariables is achieved via linear transformation $$\alpha _{u,\mu }=\underset{j}{}N_{j,\mu }\alpha _{u,j},$$ (20) where $`N_{j,\mu }=_kf_{\mu ,k}(_\nu f_{\nu ,j}f_{\nu ,k})^1`$. Thus, the problem is to find such a function $`f_{\mu ,j}`$ which would make the restoration (20) as accurate as possible. In the coarse-grained region, where the linearized Hamiltonian (18) can be used, the system’s dynamics can be represented as an independent motion of different normal modes (eigenvectors of the Hamiltonian (18)), and the choice of the $`2L`$ gross variables becomes obvious: they should be amplitudes of the normal modes corresponding to lowest $`L`$ eigenfrequencies. This choice gives an almost complete description of the system provided that the amplitudes of all the other modes, which are excluded from consideration, are much smaller. The frequencies of the excluded modes are much larger, so that their dynamics is much faster, and they can relax to local equilibrium quickly in comparison with the adiabatically slowly varying gross variables. This choice of the gross variables is in correspondence with the dynamical approach to nonequilibrium statistical mechanics. When considering motion of the system subjected to some small rapidly varying perturbation (provided, in our case, by the fast excluded modes), the well-known basic problem is to exclude from the solution so-called secular terms, which appear due to entanglement of slow and fast motions in the system. After slow and fast modes are properly separated, the standard procedure of averaging can be performed. The use of the lowest-frequency normal modes as gross variables allows elimination of the secular terms in the coarse-grained equations of motion. An analogous procedure can be identifyed also in the approaches of Zwanzig and Mori . Thus, one possible recipe is to use the weighting function $`f_{\mu ,j}=d_\mu ^j`$, where $`d^j`$ is one of the lowest-frequency eigenvectors of the dynamic matrix. In real calculations it could be inconvenient to work with delocalized collective modes. It can be shown, that an equally accurate coarse-graining can be achieved with any orthonormal combination of the eigenvectors $`d^j`$, i.e. $`f_{\mu ,j}=_kC_k^jd_\mu ^k`$ is an equally good weighting function if $$\underset{k}{}C_k^jC_k^j^{}=\delta _{j,j^{}}\underset{j}{}C_k^jC_k^{}^j=\delta _{k,k^{}}.$$ (21) The function $`f_{\mu ,j}`$ can be made well-localized using an appropriate set of $`C_k^j`$. As a specific example, we performed coarse-grained modeling of magnons in a 1-D ferromagnetic chain consisting of $`N`$ classical magnetic moments with nearest-neighbor and next-nearest-neighbor exchange interactions. The corresponding Hamilton function is $``$ $`=`$ $`{\displaystyle \underset{\mu }{}}J_\mu ^0𝐌_\mu (𝐌_{\mu +1}+𝐌_{\mu 1})`$ (23) $`+\gamma J_\mu ^0𝐌_\mu (𝐌_{\mu +2}+𝐌_{\mu 2}),`$ and periodic boundary conditions are used. A computational mesh is imposed, consisting of $`L`$ nodes, and two gross variables $`\alpha _{1,j}`$ and $`\alpha _{2,j}`$ are attributed to each of $`L`$ nodes. These two gross variables provide a coarse-grained description of a block containing $`n=N/L`$ individual moments. Several weighting functions have been used. One possible choice, giving an exact magnon spectrum, is $$f_{\mu ,j}^{(0)}=\frac{1}{N}\frac{\mathrm{sin}\pi (\mu jn)/n}{\mathrm{tan}\pi (\mu jn)/N},$$ (24) which corresponds to an orthonormal combination of exact normal modes with coefficients $`C_k^j=(1/\sqrt{N})\mathrm{exp}(\mathrm{i}jkn)`$ where $`\mathrm{i}=\sqrt{1}`$ (so it is a discrete analog of the Nyquist-Shannon uniform sampling). However, this function, due to its long tails, is not convenient for computations, and we tested its rescaled cutoff modification: $`f_{\mu ,j}^{(1)}=`$ $`Af_{\mu ,j}^0,`$ $`\mu jnn,`$ (25) $`f_{\mu ,j}^{(1)}=`$ $`0,`$ $`\mu jn>n,`$ (26) where $`A`$ is the normalization factor necessary to satisfy the condition $`_\mu f_{\mu ,j}=1`$. A second form of the weighting function $`f_{\mu ,j}^{(2)}=`$ $`1/n,`$ $`\mu jnn,`$ (27) $`f_{\mu ,j}^{(2)}=`$ $`0,`$ $`\mu jn>n,`$ (28) although very far from optimal, can be used for crude semi-qualitative computations, so we also tested its performance. The results of our tests, the dispersion curves $`\omega (k)`$ for magnons with different wave vectors $`k`$, and their group velocities $`c(k)=d\omega /dk`$, are calculated with different weighting functions, as shown in Fig. 2. Dispersion curves $`\omega (k)`$ describe the magnon dynamics in the chain, while the group velocity curve $`c(k)`$ characterizes the propagation of magnons (the latter should be tested separately since a good approximation for $`\omega (k)`$ does not necessarily imply a good approximation of its derivative). For comparison, exact curves are presented, along with the results of micromagnetic calculations. The data points for $`k=0`$ and $`k=2\pi /n`$ are not shown because the dynamic matrix formally has a singularity at these values of wave vector. For very long-wavelength magnons all types of modeling work rather well, but for shorter wave lengths the differences are large. The coarse-grained description is better than the micromagnetic even for the worst function $`f_{\mu ,j}^{(2)}`$. For the appropriately chosen weighting function $`f_{\mu ,j}^{(1)}`$, in spite of the cutoff, the difference with the exact solution is very small, even at maximal allowed wave vector values. Summarizing, we propose an approach to modeling of multiscale processes in magnets, applying the ideas of coarse grained molecular dynamics to magnetic modeling. The scheme proposed employs basic concepts of nonequilibrium statistical mechanics to couple lengthscales. We discussed possible implementation of this approach, paying particular attention to the problem of the correct choice of the weighting function used in coarse graining. Simple tests verify our conclusions and evidence that this approach can be applicable to larger and more complicated systems. Authors would like to thank R. Rudd, J. Morris, O. N. Mryasov, R. Sabiryanov, and T. Schulthess for helpful discussions. This work was partially carried out at the Ames Laboratory, which is operated for the U. S. Department of Energy by Iowa State University under Contract No. W-7405-82 and was supported by the Director of the Office of Science, Office of Basic Energy Research of the U. S. Department of Energy. This work was partially supported by Russian Foundation for Basic Research, grant 98-02-16219.
warning/0004/cond-mat0004022.html
ar5iv
text
# Quantum-Mechanical Non-Perturbative Response of Driven Chaotic Mesoscopic Systems ## Abstract Consider a time-dependent Hamiltonian $`(Q,P;x(t))`$ with periodic driving $`x(t)=A\mathrm{sin}(\mathrm{\Omega }t)`$. It is assumed that the classical dynamics is chaotic, and that its power-spectrum extends over some frequency range $`|\omega |<\omega _{\text{cl}}`$. Both classical and quantum-mechanical (QM) linear response theory (LRT) predict a relatively large response for $`\mathrm{\Omega }<\omega _{\text{cl}}`$, and a relatively small response otherwise, independently of the driving amplitude $`A`$. We define a non-perturbative regime in the $`(\mathrm{\Omega },A)`$ space, where LRT fails, and demonstrate this failure numerically. For $`A>A_{\text{prt}}`$, where $`A_{\text{prt}}\mathrm{}`$, the system may have a relatively strong response for $`\mathrm{\Omega }>\omega _{\text{cl}}`$ due to QM non-perturbative effect. The shape of the response function becomes $`A`$ dependent. The wall formula for the calculation of friction in nuclear physics , and the Drude formula for the calculation of conductivity in mesoscopic physics, are just two special results of a much more general formulation of ‘dissipation theory’ . The general formulation of the ‘dissipation’ problem is as follows: Assume a time-dependent chaotic Hamiltonian $`(Q,P;x(t))`$. For $`x=\text{const}`$ the energy is constant of the motion. For non-zero $`V\dot{x}`$ the energy distribution evolves, and the average energy increases with time. This effect is known as dissipation. Ohmic dissipation means that the rate of energy absorption (’heating’) is $`d/dt=\mu V^2`$, where $`\mu `$ is defined as the dissipation coefficient. In case of periodic driving $`x(t)=A\mathrm{sin}(\mathrm{\Omega }t)`$, one should replace $`V^2`$ by the mean square value $`\frac{1}{2}(A\mathrm{\Omega })^2`$, and the dissipation coefficient $`\mu (\mathrm{\Omega })`$ becomes frequency dependent. For simplicity we assume that conservative work is not involved in changing $`x`$. In case of the wall formula, $`(Q,P)`$ is a particle moving inside a chaotic ‘cavity’, and $`x`$ controls the deformation of the boundary. Ohmic dissipation (in the sense defined above) implies a friction force which is proportional to the velocity, where $`\mu `$ is the ‘friction coefficient’, and $`\mu V^2`$ is the ‘heating’ rate. A mesoscopic realization of such system would be a quantum dot whose shape is controlled by electric gates. In case of the mesoscopic Drude formula, $`(Q,P)`$ is a charged particle moving inside a chaotic ‘ring’, and $`x`$ is the magnetic flux through the hole in the ring. Ohmic dissipation implies Ohm law, where $`V\dot{x}`$ is the electro-motive-force, $`\mu `$ is the conductance, and $`\mu V^2`$ is the ‘heating’ rate. For a mesoscopic realization of such system note that ring geometry is not important. One may consider a simple two dimensional quantum dot driven by a time-dependent homogeneous perpendicular magnetic field . (For the latter geometry it is better not to use the term conductance while referring to the dissipation coefficient $`\mu `$). In the general analysis of the ‘dissipation’ problem one argues that due to the driving there is diffusion in energy space. This diffusion process is biased because of its $`E`$ dependence, leading to systematic increase of the average energy. This is the reason for having dissipation. Therefore we find convenient from now on to consider the diffusion coefficient $`D_\text{E}`$ as the object of our study. The relation between $`d/dt`$ and $`D_\text{E}`$ constitutes a generalization of the so called fluctuation-dissipation relation. Fig.1. Upper diagram: The various $`V`$ regimes in the theory of quantum dissipation for linear driving $`x(t)=Vt`$. Lower diagram: The various $`(\mathrm{\Omega },A)`$ regimes for periodic driving $`x(t)=A\mathrm{sin}(\mathrm{\Omega }t)`$. Note the analogy with Fig.5 of Ref. with $`xA`$ and $`VA\mathrm{\Omega }`$. The QM-adiabatic regime (including the regime $`A<A_c`$, but excluding the narrow stripes of QM-resonances) is defined by having vanishing first-order probability to go to other levels. See the text for further explanations and definitions of $`A_c`$ and $`A_{\text{prt}}`$. Ohmic dissipation is implied if $`D_\text{E}V^2`$, or $`D_\text{E}A^2`$ in case of periodic driving. Such behavior can be established within the framework of classical mechanics using general classical considerations . The classical formulation of the dissipation problem can be regarded as a systematic scheme that justifies the use of classical ‘linear response theory’ (LRT). The precise conditions for the applicability of the classical LRT result are further discussed in , and we are going to mention later on what we call ‘the trivial slowness condition’. We are interested in the quantum mechanical (QM) theory of dissipation. The traditional derivation of QM LRT leads formally to the same result as in the classical analysis . Therefore from now on we no longer distinguish between the ‘classical’ LRT result and the ‘quantal’ LRT result and use just the term ‘LRT result’. As a matter of terminology, it should be noted that the QM formulation of LRT, also known as Kubo-Greenwood formalism, is completely equivalent to the well known Fermi golden rule (FGR) picture. So let us assume that the obvious classical conditions for the validity of the LRT result are satisfied. Now the question is whether, upon quantization, there are additional $`\mathrm{}`$-dependent conditions for the applicability of the LRT result. In the traditional quantum mechanical literature, as well as in the recent mesoscopic literature, the focus is on the consequences of having finite mean level spacing $`\mathrm{\Delta }`$. This leads to the identification of the QM-adiabatic regime (extremely slow driving), and to the discussion of either the Landau-Zener mechanism or else the Debye relaxation absorption mechanism for dissipation, as well as to the discussion of QM-resonances. The main observation of is that there is another regime, the non-perturbative regime (see Fig.1), where QM LRT is not valid. As strange as it sounds, this does not imply a failure of the LRT result. On the contrary, another observation of is that the regime where the classical approximation applies, is well-contained in the non-perturbative regime, hence the LRT result becomes valid again because of quantal-classical correspondence (QCC) considerations. However, if the system does not have a good classical limit (as in RMT models) this ‘recovery’ of the LRT result is not guaranteed. Moreover, as we are going to discuss later, QCC consideration cannot exclude the possibility of having a relatively large quantal non-perturbative response whenever the LRT result is small in comparison. The outline of this letter is as follows: (1) We extend the theoretical considerations of to the case of periodic driving. (2) We give a specific example where the LRT result fails because of a quantal non-perturbative effect. (3) We comment on the issue of localization. (4) We discuss the role of QCC considerations in the theory. Based on the theoretical considerations, the reader should realize that the existence of the non-perturbative regime is not related to having finite mean level spacing $`\mathrm{\Delta }`$, but rather to having finite bandwidth $`\mathrm{\Delta }_b=\mathrm{}\omega _{\text{cl}}`$, where $`\omega _{\text{cl}}`$ is the dropoff frequency of the LRT response. In the context of mesoscopic physics this bandwidth is known as the Thouless energy. Given $`(Q,P;x)`$ with $`x=\text{const}`$, we can define a fluctuating quantity $`(t)=/x`$. The autocorrelation function of $`(t)`$ will be denoted by $`C(\tau )`$. The power spectrum $`\stackrel{~}{C}(\omega )`$ is defined as its Fourier transform. The intensity of fluctuations is defined as $`\nu =\stackrel{~}{C}(0)`$, and it is convenient to define the correlation time as $`\tau _{\text{cl}}=\stackrel{~}{C}(0)/C(0)`$. We assume for simplicity of presentation that the single time scale $`\tau _{\text{cl}}`$ completely characterizes the chaotic dynamics of the system: The power spectrum of the chaotic motion is assumed to be continuous, and it is non-vanishing up to the cutoff frequency $`\omega _{\text{cl}}=2\pi /\tau _{\text{cl}}`$. We assume that $`\stackrel{~}{C}(\omega )`$ is vanishingly small for $`\omega >\omega _{\text{cl}}`$. Fig.2. The response of a quantum mechanical system is displayed as a function of $`A`$ and $`\mathrm{\Omega }`$. The evolution is determined by the WBRM model Eq.(9). The units of energy and time and amplitude are chosen such that $`\mathrm{\Delta }=0.5`$ and $`\mathrm{}=1`$ and $`\sigma =1`$ respectively. Left: Plots of $`D_\text{E}/A^2`$ versus $`\mathrm{\Omega }/\omega _{\text{cl}}`$ for few values of $`A`$. For small $`\omega `$ the plots coincide as expected from Eq.(1). As $`A`$ becomes larger the deviations from Eq.(1) become more pronounced, and we get response also for $`\mathrm{\Omega }>\omega _{\text{cl}}`$. Right: Plots of $`D_\text{E}/D_0`$ versus $`A/\sqrt{b}`$ for few values of $`\mathrm{\Omega }/\omega _{\text{cl}}`$. The LRT result Eq.(1) implies $`D_\text{E}/D_0=1`$ for $`\mathrm{\Omega }/\omega _{\text{cl}}<1`$ and $`D_\text{E}/D_0=0`$ for $`\mathrm{\Omega }/\omega _{\text{cl}}>1`$. The purpose of the horizontal scaling is to demonstrate that $`A_{\text{prt}}`$ rather than $`A_c`$ is responsible for the deviation from this LRT expectation. Each ‘point’ in the above plots is determined by a simulation that involves typically $`35`$ realizations of the evolution until we start to see saturation due to dynamical localization effect. The typical time step is $`dt=10^4`$. In each step we verify that the normalization is preserved to an accuracy of $`0.01\%`$. In order to eliminate finite size effects we have used a self-expanding algorithm. Namely, additional $`10b`$ sites are added to each edge whenever the probability in the edge sites exceeds $`10^{15}`$. The diffusion coefficient is determined from the fitting $`\delta E(t)^2=\text{const}\times t^\beta `$. For sub-diffusive behavior ($`\beta <0.86`$) we set $`D_\text{E}=0`$. The preparation of each ’point’ in the above plots requires $`4`$ CPU days on Alpha XP1000 machine. Consider the time dependent case $`x(t)=A\mathrm{sin}(\mathrm{\Omega }t)`$. The LRT result for the diffusion in energy is $`D_\text{E}={\displaystyle \frac{1}{2}}\stackrel{~}{C}(\mathrm{\Omega })\times \frac{1}{2}(A\mathrm{\Omega })^2`$ (1) and we shall use the notation $`D_0=\frac{1}{4}\nu (A\mathrm{\Omega })^2`$. The most transparent QM derivation of this result is based on the FGR picture. The energy levels of the systems are $`E_n`$, and the mean level spacing is $`\mathrm{\Delta }`$. The Heisenberg time is $`t_\text{H}=2\pi \mathrm{}/\mathrm{\Delta }`$. The transitions between levels are determined by the coupling matrix elements $`(/x)_{nm}`$. It is well known that for reasonably small $`\mathrm{}`$ this matrix is a banded matrix. The bandwidth is $`\mathrm{\Delta }_b=2\pi \mathrm{}/\tau _{\text{cl}}`$, and the variance of the in-band elements is $`\sigma ^2=\nu /t_\text{H}`$. It is common to define the QM system using the four parameters $`(\mathrm{\Delta },b,\sigma ,\mathrm{})`$ where $`b=\mathrm{\Delta }_b/\mathrm{\Delta }`$. It is also useful to regard the semiclassical relations $`\tau _{\text{cl}}=2\pi \mathrm{}/(b\mathrm{\Delta })`$ and $`\nu =(2\pi \mathrm{}/\mathrm{\Delta })\sigma ^2`$ as definitions, whenever the classical limit is not explicitly specified \[as in Random Matrix Theory (RMT) models\]. The FGR picture implies strong response if and only if $`\mathrm{}\mathrm{\Omega }<\mathrm{\Delta }_b`$, leading to $`\mathrm{\Omega }<\omega _{\text{cl}}`$ as in the classical case. Using the FGR picture it is straightforward to recover $`D_0=(\pi /2)(\mathrm{}/\mathrm{\Delta })(\sigma A\mathrm{\Omega })^2`$ in agreement with Eq.(1). The trivial slowness condition for the applicability of classical LRT is $`V\tau _{\text{cl}}\delta x_c^{\text{cl}}`$. Here $`\delta x_c^{\text{cl}}`$ is the parametric change that leads to the breakdown of the linearization of $`(Q,P;x+\delta x)`$ with respect to $`\delta x`$. Upon quantization there are two other parametric scales that become important , namely $`\delta x_c^{\text{qm}}`$ and $`\delta x_{\text{prt}}`$. The former is the parametric change which is required in order to mix neighboring levels, while the latter is the parametric change required in order to mix all the levels within the band. Hence we define $`A_c\delta x_c^{\text{qm}}=`$ $`{\displaystyle \frac{\mathrm{\Delta }}{\sigma }}`$ $`\mathrm{}^{(1+d)/2}`$ (2) $`A_{\text{prt}}\delta x_{\text{prt}}=`$ $`\sqrt{b}{\displaystyle \frac{\mathrm{\Delta }}{\sigma }}`$ $`={\displaystyle \frac{2\pi \mathrm{}}{\sqrt{\nu \tau _{\text{cl}}}}}`$ (3) The above parametric scales, of time-independent first-order perturbation theory (FOPT), manifest themselves also in the time-dependent analysis. FOPT gives the following result for the probability to make a transition from an initial level $`m`$ to some other level $`n`$, $`P_t(n|m)=\left|{\displaystyle \frac{1}{\mathrm{}}}\left({\displaystyle \frac{}{x}}\right)_{nm}{\displaystyle _0^t}x(t^{})\mathrm{exp}\left(i{\displaystyle \frac{(E_nE_m)t^{}}{\mathrm{}}}\right)𝑑t^{}\right|^2`$ (4) The total transition probability is $`p(t)=_n^{}P_t(n|m)`$ where the prime imply omission of the $`n=m`$ term. In the regime $`\mathrm{\Omega }<\omega _{\text{cl}}`$ one obtains $`p(t)={\displaystyle \frac{1}{\mathrm{}^2}}\nu A^2\times \{\begin{array}{ccc}(\mathrm{\Omega }^2/\tau _{\text{cl}})\frac{1}{4}t^4& \text{for}& 0<t\tau _{\text{cl}}\\ \mathrm{\Omega }^2\frac{1}{3}t^3& \text{for}& \tau _{\text{cl}}t1/\mathrm{\Omega }\\ \frac{1}{2}t& \text{for}& 1/\mathrm{\Omega }tt_\text{H}\end{array}`$ (5) while in the regime $`\mathrm{\Omega }>\omega _{\text{cl}}`$ one obtains $`p(t)={\displaystyle \frac{1}{\mathrm{}^2}}{\displaystyle \frac{\nu }{\tau _{\text{cl}}}}{\displaystyle \frac{A^2}{\mathrm{\Omega }^2}}\times \{\begin{array}{ccc}(1\mathrm{cos}(\mathrm{\Omega }t))^2& \text{for}& 0<t\tau _{\text{cl}}\\ 3/2& \text{for}& \tau _{\text{cl}}tt_\text{H}\end{array}`$ (6) In both cases for $`t>t_\text{H}`$ we have recurrences, and therefore $`p(t)p(t_\text{H})`$. \[One should be more careful near resonances: There $`t_\text{H}`$ should be replaced by $`2\pi \mathrm{}/\delta `$, where $`\delta `$ is the detuning\]. The necessary condition for applicability of FOPT at time $`t`$ is that $`p(t^{})1`$ for any $`t^{}<t`$, which can be written as $`p([0,t])1`$. The necessary condition for the applicability of the FGR picture is $`p([0,\tau _{\text{cl}}])1`$. This is the FGR condition which guarantees the separation of time scales $`\tau _{\text{cl}}\tau _{\text{prt}}`$. The FOPT breaktime $`\tau _{\text{prt}}`$ is defined as the maximal $`t`$ for which $`p([0,t])<1`$. Now we can define a non-perturbative regime by the requirement $`p([0,\tau _{\text{cl}}])>1`$. It is straightforward to observe that the non-perturbative regime is contained in the region $`A>A_{\text{prt}}`$ where $`\left({\displaystyle \frac{A_{\text{prt}}}{A}}\right)\omega _{\text{cl}}<\mathrm{\Omega }<\left({\displaystyle \frac{A}{A_{\text{prt}}}}\right)\omega _{\text{cl}}`$ (7) The location of the non-perturbative regime is illustrated in Fig.1. For completeness of presentation we have also indicated the subregion in $`\mathrm{\Omega }<\omega _{\text{cl}}`$ where we have first-order response equal to zero. The condition is $`p([0,\mathrm{}])1`$ or equivalently $`p(t_\text{H})1`$. This region contains the QM adiabatic regime, including the region $`A<A_c`$, but excluding the narrow stripes of resonances. We have defined the location of the non-perturbative regime, but we did not yet give a suggestion how Eq.(1) should be modified. Using RMT assumptions with regard to $`(/x)_{nm}`$, and inspired by related studies of wavepacket dynamics , we expect the result $`D_\text{E}=(C/\sqrt{v_{\text{PR}}})\times D_0`$ (8) where $`v_{\text{PR}}=V/(\delta x_{\text{prt}}/\tau _{\text{cl}})`$, and $`C`$ is a numerical constant. A detailed derivation of (8) will be presented in the future. (The crucial step is to argue that at $`\tau _{\text{prt}}`$ there is a crossover from ballistic behavior to diffusion in the sense of ). For periodic driving this result should be averaged over a period leading to $`D_\text{E}A^{2\alpha }`$ with $`\alpha =1/2`$. We wanted to give a numerical example that demonstrate the non-perturbative response effect. Evidently, the simplest is to consider a time-dependent version of Wigner’s banded random matrix (WBRM) model, $`=𝐄_0+x(t)𝐁`$ (9) where $`𝐄_0`$ is an ordered diagonal matrix, and $`𝐁`$ is a banded matrix. This model is characterized by the parameters $`(\mathrm{\Delta },b,\sigma ,\mathrm{})`$ which we have defined previously. For the numerical experiment we have assumed rectangular band profile such that all the elements $`0<|nm|b`$ are taken from the same distribution, and outside the band all the elements are identically zero. The results of the simulations are summarized in Fig.2. It should be realized that WBRM model Eq.(9) has a big disadvantage. Namely, unlike the physical examples of the introduction, the statistical properties of the model are not invariant for $`x(t)x(t)+\text{const}`$. One may wonder why we do not use one of the two other popular variations of Wigner model , eg $`=𝐄+(\mathrm{cos}x)𝐁_1+(\mathrm{sin}x)𝐁_2`$. The problem is that for these models one obtains $`\delta x_{\text{prt}}\delta x_c^{\text{cl}}2\pi `$. Therefore there is no regime there where LRT fails because of quantal non-perturbative effect. (Such failure requires the generic separation of scales $`\delta x_{\text{prt}}\delta x_c^{\text{cl}}`$). Thus it seems that the only way to make a RMT model $`x`$-invariant, is to keep it ‘perturbative’ in nature. The lack of $`x`$-invariance in the standard WBRM model complicates the calculation of the period-averaged $`D_\text{E}`$ and leads to $`D_\text{E}A^{2\alpha }`$ with $`\alpha `$ changing gradually from $`1/2`$ to $`1`$. The numerical analysis of Fig.2b fits well to $`\alpha 3/4`$. There are two types of localization effects that we had to consider in our numerical experiments. The “WBRM model localization” can be avoided by using amplitudes $`A<b^{3/2}(\mathrm{\Delta }/\sigma )`$ in order to guarantee that the instantaneous eigenstates of Eq.(9) are not localized at any time. The “dynamical localization effect” on the other hand cannot be avoided. It is associated with the periodic nature of the driving. Extending standard argumentation one observes that the eigenstates of the (one period) Floquet operator have localization length $`\xi \times \mathrm{\Delta }`$, and that the associated breaktime is $`t^{}=\xi \times (2\pi /\mathrm{\Omega })`$. The two must be related by $`2D_\text{E}t^{}=(\xi \mathrm{\Delta })^2`$ leading (in the LRT regime) to the result $`t^{}=2\pi ^2(A/A_c)^2\times t_\text{H}`$. In all our numerical experiments the diffusion has been determined for times where dynamical localization is not yet apparent. Another possibility, which we have not used, is to add a small noisy component to the driving, such as to mimic the typical experimental situation of having dephasing time much shorter than $`t^{}`$. It is not obvious that the non-perturbative behavior that is implied by RMT assumptions, and applies to RMT models, should apply also to Hamiltonians that possess a well defined classical limit. On the contrary, the same considerations as in can be applied in order to argue that RMT considerations are not compatible with the QCC principle. Here we are going to explain the main idea, and to define our expectations. Taking $`\mathrm{}`$ to be very small, it is obvious that eventually we shall find ourselves in the regime where $`AA_{\text{prt}}`$. Let us consider the dynamics during a specified time interval $`0<t^{}<t`$. The time $`t`$ is chosen to be much larger than $`\tau _{\text{cl}}`$. On the basis of QCC considerations, we should be able to make $`\mathrm{}`$ sufficiently small such that the quantum evolution becomes similar to the classical evolution up to the time $`t`$. The classical analysis implies that during this time the stochastic behavior is established. Therefore having detailed QCC during the time $`t`$ implies that the quantal $`D_\text{E}`$ can be approximated by the classical result. This leads to a contradiction with the RMT prediction Eq.(8) in the domain $`\mathrm{\Omega }<\omega _{\text{cl}}`$, but not in the domain $`\mathrm{\Omega }>\omega _{\text{cl}}`$. We are going to further explain this last point. Denote the energy dispersion by $`\delta E_{\text{qm}}(t)`$, and the corresponding classical result by $`\delta E_{\text{cl}}(t)`$. For sufficiently small $`\mathrm{}`$ it should be possible to make a leading order approximation $`\delta E_{\text{qm}}(t)\delta E_{\text{cl}}(t)+\mathrm{}^\gamma g(t)`$, with $`\gamma >0`$. In the $`\mathrm{\Omega }<\omega _{\text{cl}}`$ regime the first term in this approximation is dominant. On the other hand for $`\mathrm{\Omega }>\omega _{\text{cl}}`$ the first term gives a vanishingly small result for $`D_\text{E}`$. Therefore, without any contradiction with QCC considerations, the second term becomes important. Therefore we may have in principle an enhanced quantal response for $`\mathrm{\Omega }>\omega _{\text{cl}}`$. In conclusion, we have defined a non-perturbative regime in the $`(\mathrm{\Omega },A)`$ plane, where LRT cannot be trusted. We have demonstrated an actual failure of LRT for a particular (RMT) Hamiltonian. We believe that for generic chaotic systems the RMT mechanism for diffusion competes with the classical mechanism. The actual response of the system is expected to be determined by the predominant mechanism. The study of this conjecture is the theme of our future studies. We thank the Centro Internacional de Ciencias (Cuernavaca Mexico) for their hospitality during the Quantum Chaos workshop. DC thanks ITAMP for support.
warning/0004/astro-ph0004082.html
ar5iv
text
# Network and internetwork: a compared multiwavelength analysis ## 1 Introduction The chromospheric bright network has long been observed in narrowband spectroheliograms taken in the H and K cores of the Ca II resonance lines. The Ca II network typically shows H and K profiles with high double peaks and enhanced line wings that persist for extended periods of time (longer than 10 minutes, see, e.g., Rutten & Uitenbroek 1991 ). The chromospheric network emission pattern is cospatial with small-scale magnetic field concentrations, and defines the supergranular network boundaries. It is this atmospheric component that produces the correlation between H and K excess line-core flux and magnetic activity of cool stars (Schrijver et al. (1989)). The dynamics of the network elements, compared with the internetwork or quiet chromosphere, has been extensively studied (especially from the observational point of view) since these small-scale structures can be important in channeling the energy from photospheric layers to the transition region and corona (Kneer & von Uexküll (1986), 1993, Deubner & Fleck (1990), Kulaczewski (1992), Al et al. 1998). An assessment of the spectral characteristic properties of Network Bright Points (NBPs) at different layers in the atmosphere has been provided by Lites et al. (1993) using spectral observations in the range of the Ca II H line. In their work, these authors analyzed spectrographic observations of a single network bright patch and of several internetwork points. The wavelength shifts of photospheric and chromospheric lines allowed them to perform a compared analysis between the dynamics of the two atmospheric components. One of the relevant characteristics they describe is that at chromospheric levels (Ca II H<sub>3</sub>) the NBPs show long period oscillations ($`\nu <`$ 3 mHz) not correlated with oscillations in the lower atmosphere, while they lack power at higher temporal frequencies. The internetwork regions display instead enhanced power at higher frequencies, well correlated with photospheric oscillations. The presence of these low frequency oscillations in the network has been confirmed by Lites (1994) also for the chromospheric He I 10830 line, in contrast to Bocchialini et al. (1994) which observe, for the same line, oscillations only in the 5 minutes range. An enhanced power in the low frequency range for network points with respect to the internetwork has also been observed by Kneer & von Uexküll (1986) in the center of the chromospheric H$`\alpha `$ line. These authors however interpret this feature as not due to oscillations, but of mainly stochastic origin, and attribute it to erratic motions of the corresponding photospheric footpoints. The problem is still open, and further observations to better address this issue are required (Lites (1994)). In particular one would need observations: 1- on a larger number of NBPs, to improve on the statistics; 2- at different heights in the atmosphere, since the analysis of the coherence between fluctuations at different levels can help exploring the nature of oscillations. To this end, a reliable method for the identification of the same physical structure at different atmospheric levels is mandatory, since the inclination of the magnetic field could displace the chromospheric network points with respect to the corresponding photospheric ones. In this paper, we address some of these issues, and present observational results on the NBPs and internetwork characteristics as derived from a multiwavelength analysis. The observations were obtained in August 1996, during a coordinated observing program between ground-based observatories and the Solar and Heliospheric Observatory (SOHO). For the ground-based observations we used the cluster of instruments at the NSO/Sacramento Peak R.B. Dunn Solar Telescope (NSO/SP-DST), that could provide a complete coverage at lower atmospheric levels. The dataset used is described in Sect. 2. General properties of a sample of NBPs, followed from the photosphere up to the chromosphere and including their relationship with the magnetic structures, are given in Sect. 3. The temporal development of the NBPs is described in Sect. 4. Sect. 5 and 6 provide an analysis of the power, phase difference and coherence spectra for the fluctuations observed separately within the NBPs and the surrounding internetwork. Finally, discussion and conclusions are given in Sect. 7. ## 2 Observations and data reduction A description of the general data acquisition has been presented in Cauzzi et al. (1997, 1999). Table 1 gives a summary of the observing setup, and we recall here only some short information on the data used in this paper. Monochromatic intensity images were obtained with the tunable Universal Birefringent Filter (UBF) and the Zeiss filter, at high spatial and temporal resolution. Several spectra were obtained around the chromospheric CaII K line with the Horizontal Spectrograph (HSG). The spectra have been acquired setting the spectrograph slit on different bright points at different times; the field of view in the HSG row of Table 1 hence refers to a single slit exposure. Onboard SOHO, the Michelson Doppler Imager (MDI, Scherrer et al. (1995)) acquired data in high resolution mode, i.e. with an image scale of 0.605″/pixel. Maps of pseudo-continuum intensity, line-of-sight velocity, and longitudinal magnetic flux were obtained in the NiI 6768 Å line at a rate of one per minute for several hours. The line-of-sight velocity images were available in a binned 2x2 format, i.e. with an effective spatial scale of 1.2″/pixel. We observed the small Active Region (AR) NOAA 7984 over 5 consecutive days (Aug. 15 - 19, 1996). Its activity was very low and although some stronger magnetic structures (a small spot and some pores) were present in the field of view (FOV), no major eruptions of magnetic flux were recorded. The situation was then appropriate to study and characterize the properties of NBPs visible within and around the AR. We remark that our set of data allows us to directly compare the NBPs as visible at different wavelengths with the corresponding magnetic field structure, and to follow them from the deep photosphere to the higher chromosphere. In this paper we analyze the data obtained on August 15th, 1996, since for that day we had the best uniformity in time coverage for NSO and MDI data. Fig. 1 shows the FOV at several wavelengths. The period of best seeing for ground-based observations ran from 15:15 to 16:05 UT. This interval is adequate for the study of network points, since it allows an analysis of their (possible) periodical properties, while they still maintain their identity (Lites et al. (1993); von Uexküll & Kneer (1995)). MDI data were available for many hours around this interval; we consider in this work the period 14:00 - 17:00 UT. The data were re-scaled to the 0.605″/pixel of the MDI maps, and the co-alignment of the whole dataset was obtained comparing the position of the prominent solar features, i.e. the little spot and pores (see Fig. 1). At each given time the alignment among the images acquired with different instruments was better than about $`1^{\prime \prime }`$, the mean spatial resolution limit of the ground-based frames. ## 3 Characteristics of NBPs In order to identify suitable network points, we first selected on the spectra bright features showing strong K<sub>2</sub> peaks and enhanced wings emission in CaII K. Since the NBPs lifetime is typically longer than 10 minutes (Rutten & Uitenbroek 1991), we further required that the points be visible for the entire observing period in the NaD<sub>2</sub> images as bright structures with intensity above the average. A total of 11 NBPs with the required characteristics were selected. They are distributed over the FOV both near the center of the AR and away from it, as can be seen in Fig. 1-a. Several more network points (or structures) are visible on the longitudinal magnetic flux maps, but in this work we limit our analysis to only these 11 points for which corresponding CaII K spectra are available. ### 3.1 Intensity and velocity Within the areas enclosing the NBPs defined above, we tried to identify the bright points at each wavelength by choosing an intensity (or other signature) threshold that could clearly separate them from their surroundings. On white light and NiI pseudo-continuum images, an intensity threshold cannot clearly discriminate between NBPs and other areas. The contrast averaged over the spatial locations corresponding to the NBPs is of the order of 1%, i.e. smaller than the rms noise calculated in quiet areas (about 3.5% and 2% for the white light and pseudo-continuum images, respectively). This is consistent with the observations of Topka et al. (1997), that find low continuum intensity contrast in network points with magnetic flux density smaller than $``$ 300 G (as typical of our points, see Sect. 3.2). The line of sight velocities averaged over the NBP areas are small, about 80 m/s downward with respect to the average values over the quiet areas of the FOV. Since the standard deviation of the measurements is about 200 m/s, for both NPBs and quiet areas, it is not possible to define a velocity threshold that allows to discriminate the NBPs within the FOV. The small average red-shift is however in agreement with recent observations by Solanki (1993) and Martínez-Pillet et al. (1994). The NBPs are instead well visible in the images acquired in the H$`\alpha `$ far wings and NaD<sub>2</sub> center, as sharp and isolated bright structures of comparable size (3″$``$4″wide). The bright points, as seen at these wavelengths, spatially coincide within the overlapping error of 1″. In the H$`\alpha `$ center images the morphology is less clear. The presence of contiguous features, at times brighter than the selected NBPs, made more uncertain their identification. In total, however, 10 of the 11 considered NBPs were unambiguosly identified in the H$`\alpha `$ center images. The characteristics of the NBPs for different signatures, including their typical contrast with respect to quiet areas, are summarized in Table 2. It must be remarked that the formation height of these signatures has been computed in a mean quiet atmosphere, i.e. that it represents only a generic indication for magnetic structures such as the NBPs. In particular, due to the Wilson depression, the radiation coming from photospheric magnetic structures is believed to be formed in deeper geometric layers with respect to non-magnetic ones. Since the spatial resolution for our observations is not sufficient to resolve the (supposedly) elementary magnetic fluxtubes (with dimensions smaller than 0.3″, as seen for example in G-band images), the signals we analyze are a non-linear combination of magnetic and non-magnetic ones. ### 3.2 Magnetic structures The selected NBPs are clearly recognizable on the MDI magnetic maps as sharp and isolated structures, 3″$``$4″ wide. Each one corresponds to a patch of definite polarity, with magnetic flux densities ranging from a minimum of 30 G, to a maximum of about 250 G (for comparison, the spot in the FOV has a maximum flux density of 1100 G). The error on a single pixel in each image is given at about 15 G (Schrijver et al. (1997)). Since the network fields are mostly vertical (see, e.g., Lites et al. (1999)), the flux measure is only slightly affected by the position of the FOV on the solar disk ($`\mathrm{cos}\theta 0.9`$). The magnetic evolution of the NBPs is quite varied. Some of the points maintain stable positions and flux values, while others experience a steady increase during the observing period. In one single case we see a weak magnetic structure appearing during the observing sequence, simultaneously with the appearance of a network bright point. The spatial correspondence between the NaD<sub>2</sub> network points and the magnetic structures is very good, and will be analyzed in more detail in the next section. The same correspondence is noticeable also for the network points as seen in the H$`\alpha `$ wings, although this is less evident than for NaD<sub>2</sub> due to their lower contrast. ### 3.3 Magnetic structure and NaD<sub>2</sub> emission We checked the spatial correspondence between the network points as visible in NaD<sub>2</sub> and in magnetic maps. To this end we first removed from both signatures, by means of appropriate smoothing techniques, short term variations due to noise and oscillations (especially in the 5-minutes range). We then subtracted a threshold, chosen as the average quiet area value for the NaD<sub>2</sub> images, and as the nominal data noise (15 G) for the magnetic ones. Finally, within the 11 selected areas, a NBP was identified in both signatures as the locus of the pixels exceeding 50% of the local maximum. This allowed its clear separation from the surroundings. We find that, at any given time, the NaD<sub>2</sub> network points are coincident in position, size and shape with the corresponding magnetic patches, within 1″ (the overlapping error). Any change in the characteristics of the NaD<sub>2</sub> NBPs reflect almost perfectly those of the magnetic features, within the temporal resolution of the MDI data. This is well exemplified in Fig. 2, where we show NaD<sub>2</sub> contours overlaying magnetic flux maps of several NBPs at two different times. To our knowledge, this is the first time that such a correspondence is reported at high spatial and temporal resolution. A good agreement between the chromospheric network emission pattern and the locations of enhanced magnetic flux had been noted in earlier works (Skumanich et al (1975); Schrijver et al. (1989); Nindos & Zirin 1998 ) but mostly for the CaII K emission, with lower spatial and temporal resolution. No temporal resolution was available in the observations of Beckers (1976) or in those of Daras-Papamargaritis & Koutchmy (1983), that established a correlation between magnetic structures and facular structures in the wings of the Mgb lines. In analogy with the CaII H and K case, we could establish a quantitative relationship between the NaD<sub>2</sub> excess and the magnetic flux density for the network points. This property implies that also the emission in the center of the NaD<sub>2</sub> line can be used as a proxy for the magnetic field structures. For sake of simplicity, the details of the determination of this relationship are given in A. ## 4 Temporal development: light curves To study the temporal development of the NBPs, we computed the light curves for the bright points at each wavelength or signature. The curves for each NBP were obtained by selecting an area that contained the bright point throughout the whole observing period (even if it moved spatially), and then averaging, for each time, over all the pixels whose intensity exceeded the threshold value described in Sect. 3.1. A threshold equal to the nominal data noise was used for the magnetic curves. As said in Sect. 3.1, the NBPs are not directly visible in some photospheric signatures, such as white light, NiI pseudo-continuum and velocity images, hence we couldn’t use an intensity (velocity) threshold to obtain the corresponding light curves. To guarantee the comparability with the other light curves, in these cases we computed, at each time, the average value over the spatially corresponding magnetic areas. We also computed the light curves of 11 areas randomly selected in the quiet regions of our FOV, and of size comparable to that of the NBPs (about 3″$`\times `$3″). No threshold was applied for their computation. These quiet areas should represent the so-called internetwork regions, which appear field-free at MDI sensitivity. Using the same number of internetwork areas as of network bright points, and performing the analysis in the same way, will give us confidence in the comparison in a statistical sense. We remark that this is not always the case, especially for spectrographical observations where the slit samples a number of internetwork points that is usually much larger than that of network structures. In Fig. 3, as an example, we show the light curves of different signatures for a NBP. Fluctuations are evident at each wavelength. The red and blue wings of H$`\alpha `$ show a very similar temporal evolution, with simultaneous intensity variations of the same magnitude for most NBPs. For other signatures a direct comparison does not show any simple relation between the variations at different atmospheric heights. We must comment here on the fact that most authors, when studying the temporal properties of network points, use a different method to determine the “light curves”. Basically, the network points are spatially identified using the temporal average of some suitable signature, and then the temporal development of each image pixel belonging to the average structures is considered and analyzed. We believe that the method we adopted has several advantages: \- Since the network points might move over the course of time, they do not always correspond spatially to the structures identified on the average maps. Our method guarantees that the structure is properly followed in time, avoiding the loss of relevant pixels, or the inclusion of spurious ones; - If the magnetic structure giving rise to the network point inclines with height, a set of pixels identifying the NBP at a given wavelength might not well represent the same structure at a different wavelength. This is especially relevant when performing comparisons between different atmospheric layers, for example in the phase difference analysis of Sect. 6. We overcame this problem by selecting areas large enough to include the NBPs at each wavelength and each time, as explained earlier. Computing the light curves as an average over a given area implies the assumption of a spatial coherence over the whole area (of about 3″$`\times `$3″, equivalent to a spatial frequency of about 3 Mm<sup>-1</sup>) competing to each NBP. This seems a reasonable assumption because we do not see any significant inhomogeneities within the single structures. ## 5 Power spectra A search for possible periodicities in the fluctuations of the NBPs light curves was performed using temporal power spectra. Before computing the power spectra, the light curves were detrended using a smoothing window of 600 s. A check on this procedure showed that changing the smoothing window between 360 s and 840 s affected the power at frequencies lower than 1.2 mHz, but without changing the frequency of the peaks. The power at higher frequencies remained unaffected. A power spectrum was computed for each light curve of the 11 NBPs and of the 11 internetwork areas. To analyze the differences between these two atmospheric components, we averaged separately the power spectra over all of the NBPs and over all of the quiet regions. In Fig. 4 we show some of these averaged curves. It must be remembered that the NSO and MDI observations have different temporal coverage (50 and 180 minutes, respectively) and temporal resolution (12 s and 60 s), so that lower temporal frequencies are better represented in the NiI series, while frequency coverage extends to higher frequencies for ground-based data. However, the power at $`\nu 8`$ mHz in all the ground-based signatures is due essentially to noise (Fig. 4), hence the analysis can be limited to the frequency coverage of the NiI observations, 0 - 8 mHz. We describe here the power spectra characteristics from lower photospheric signatures to higher chromospheric ones. Intensity fluctuations may be plausibly interpreted as temperature fluctuations for photospheric LTE signatures such as Ni I or the H$`\alpha `$ wings, formed over depths where the velocity gradient is small. In these cases, the intensity fluctuations directly reflect fluctuations of the source function (the Planck function) and hence of temperature. In the center of chromospheric lines as Na$`D_2`$ or H$`\alpha `$ the NLTE effects are important and the intensity fluctuations are the response to temperature, density and even velocity changes (Cram 1978) and then are a sort of average of the variation of the state variables of the chromosphere. ### 5.1 Photospheric signatures In Fig. 4 a, b, d we show the power spectra of Ni I pseudo-continuum intensity, Ni I velocity and H$`\alpha `$ red wing intensity. The power spectra computed for white light intensity (not shown in figure) and for the NiI pseudo-continuum are very similar, and we are confident that we can compare the two series of observations, even if their frequency resolution is different. The H$`\alpha `$ blue wing spectrum (not shown in figure) has a similar behaviour but a lower amplitude with respect to the red wing (see Table 3). A value smaller than 1 for the ratio of the power in the blue and red wings of H$`\alpha `$ is consistent with the observations of Bertello (1987), that found the same trend for the power of velocity oscillations in the wings of photospheric lines formed at heights lower than 150 km. As is well known, the distribution of power for the photospheric velocity fluctuations is rather different from the one of pseudo-continuum intensity fluctuations (Fig. 4 a-b). The velocity power is concentrated around the range of frequencies corresponding to the 5-minutes oscillations. The pseudo-continuum power spectrum peaks at low frequencies, around 1.5 mHz and then show a decay that might indicate the stochastic character of the granulation intensity variation, as already reported for the first time in Noyes (1967). As a global characteristic, power spectra computed in photospheric signatures do not show any significant difference between network and internetwork structures within the 50% confidence limit (Fig. 4 a, b, d). This result is consistent with previous spectral observations by several authors (Deubner & Fleck (1990); Kulaczewski (1992); Lites et al. 1993) that analyzed both intensity and velocity oscillations in the photosphere for network and internetwork features. The power spectrum of the magnetic flux variations averaged over the NBPs is shown in Fig. 4 c. Internetwork areas are not considered because the noise in the magnetic flux measure is too high for a reliable determination of fluctuations. Significant peaks are visible at low frequency (around 1.5 mHz), indicating long term evolution of the magnetic field, and around 3.5 mHz corresponding to the 5 minutes oscillations. A signal at the latter timescale might represent the magnetic response to oscillations already present in the photosphere and be of importance in the context of generation and dissipation of MHD waves in the solar atmosphere (Ulrich (1996)). Observations of flux variations in small magnetic structures are scarce in the literature, but we can compare this result with those presented by Norton et al. (1999), that used a similar set of MDI data obtained in the area of a big sunspot. They found a significant peak near 5 min only for structures whose magnetic flux density exceeded 600 G, while the points we analyzed had a maximum value of about 300 G. ### 5.2 Chromospheric signatures The intensity power spectra computed in chromospheric signatures display strong differences between NBPs and internetwork as shown in Fig. 4 e - f. We will analyze in detail these differences, keeping in mind that the regime of oscillations changes with height in the chromosphere. First of all we notice that the power spectrum of internetwork intensity fluctuations in NaD<sub>2</sub> shares some characteristics with the photospheric NiI velocity power spectrum rather than with the one of Ni I intensity. In particular the strongest peak appears around 3.5 mHz, while the enhanced low frequency component, typical of photospheric intensity power spectra, is lacking. This characteristic could be explained if the intensity fluctuations in the NaD<sub>2</sub> line center were related more to velocity than to temperature perturbations. This might be indirectly confirmed by the results of Pallé et al. (1999) in their study of the current performances of the GOLF experiment on SOHO. In determining the relative contributions of velocity and intensity signals to the intensity variations measured in the blue wings ($``$100 mÅ) of the sodium doublet, they conclude that the effect due to “pure” intensity changes is only 14% that of velocity changes for the p \- mode frequency range. Since the width of the filter used for our observations includes that same portion of the line wing, this conclusion might apply, at least partially, also to our case. Comparing network and internetwork power spectra for NaD<sub>2</sub>, we see that the power of the NBPs is smaller than the corresponding power for internetwork at each temporal frequency. In particular, even if both NBPs and internetwork points show a maximum around 3.5 mHz, the power at this frequency is almost an order of magnitude smaller in the NBPs. (Fig. 4 d suggests that this effect might be already present in the wings of H$`\alpha `$, although its amplitude is not large enough to give an unambiguous result). This suppression of power in the NBPs indicates that the presence of the magnetic field in some way perturbs and reduces the oscillations at low chromospheric levels, especially in the $`p`$-mode range. A compression of power in magnetic structures at frequencies below 7-8 mHz, for lines formed at similar heights, has not been reported by other authors. Only Al et al. (1998) observe a similar effect, but much smaller, for the power spectrum of velocity fluctuations measured in the center of NaD<sub>2</sub> with a narrowband filter (30 mÅ). We think that the stronger effect seen in our observations is real because our analysis selects the horizontal scale typical of chromospheric network to determine the light curves and the power spectra (see Sect. 4), and is therefore more suitable to outline characteristics and differences on the same horizontal scale for NBPs and internetwork areas. The power spectrum computed for the intensity of H$`\alpha `$ center is shown in Fig. 4 f. For both NBPs and internetwork the power distribution peaks at low frequencies ($`\nu 3`$ mHz), without any relevant peak at the $`p`$-mode frequencies. No enhanced power for “3-minutes” oscillations ($`\nu >5`$ mHz), is detectable in the internetwork. This is consistent with observations in the H$`\alpha `$ center by Cram (1978) and in the Ca II - H3 by Lites et al. (1993) showing a power peak in the “3 minutes” range only for velocity power spectrum. The spectrum of NBPs in H$`\alpha `$ line has a power higher than that of the internetwork areas at each frequency, reversing the effect present in the NaD<sub>2</sub> line, and shows three well separated peaks reminiscent of the peaks observed by Lites et al. (1993). The more relevant peak is around 2.2 mHz (“7-minutes” oscillations) and is lowered about a factor 6 in the internetwork. We cannot judge on the relevance of the peak around 1.3 mHz, since its amplitude is heavily affected by the smoothing window, as described in Sect.5, and we cannot consider real the peak at 0.6 mHz, because it is related to the time interval of our observations. However, the increasing power at low frequencies in the spectrum of NBPs suggests that the rôle of the magnetic field in the oscillations, detectable at high chromospheric levels, is certainly different from the one at lower levels and strenghtens the hypothesis of magnetic-hydrodynamic waves present at these high levels. Characteristics of power spectra for network and internetwork are summarized in Table 3 for the two relevant frequency windows 1.5 - 2.5 mHz and 3 - 4 mHz. ## 6 Phase difference and coherence spectra In order to look for propagation characteristics of waves at different heights in the atmosphere, we computed phase difference ($`\mathrm{\Delta }\mathrm{\Phi }`$) and phase coherence ($`C`$) spectra for many signature pairs (Straus (1995)). We exclude in this analysis the NaD<sub>2</sub> signature because, if the measured intensity oscillations are due essentially to velocity oscillations, we cannot establish the direction of the motion and hence assign the correct value to the phase difference. Following Edmonds & Webb (1972), we adopted a smoothing width of about 1.5 mHz in the Fourier domain for the computation of phase and coherence spectra. Frequencies smaller than 0.7 mHz hence do not convey any significant information. Taking into account our smoothing width, a coherence smaller than $``$0.5 implies that phase differences at those frequencies are completely unreliable. NSO and MDI data were analyzed independently, using their own frequency resolution and coverage. Finally, the study was performed separately for internetwork and network areas, to distinguish between features with different magnetic characteristics. We describe the characteristics of the spectra from lower photospheric layers through higher chromospheric ones. In Fig. 5 left and central column, we show intensity (I$``$I) phase difference and coherence spectra for the pairs H$`\alpha `$ $`+`$ 1.5 Å /H$`\alpha `$ $``$ 1.5 Å, and H$`\alpha `$ $`+`$ 1.5 Å / white light, originating at photospheric levels. We use here the white light signal (and not the NiI pseudo-continuum) in order to keep the maximum possible frequency resolution. Phase difference and coherence spectra between magnetic flux density and velocity (B$``$V) for the NBPs are shown in Fig. 5 right column. To search for the relationship between the oscillations present at chromospheric and photospheric levels we computed the I$``$I phase difference and coherence spectra separately for the pairs H$`\alpha `$ center / H$`\alpha `$ $`1.5`$ Å and H$`\alpha `$ center / H$`\alpha `$ $`+1.5`$ Å, shown in Fig. 6. At each considered atmospheric level, a general characteristic is that the coherence for NBPs is smaller than for internetwork, hence the phase values for NBPs are more uncertain. We examine different signature pairs separately in the two frequeny intervals 1.5 - 2.5 mHz and 3 - 4 mHz, disregarding the features with negligible power, and we summarize in Table 4 the phase and coherence values for each pair. ### 6.1 Low frequency (1.5 - 2.5 mHz) For both internetwork and network points the I$``$I phase difference between the H$`\alpha `$ red and blue wings is 0°. The two signals should be in phase if the observed oscillations are due only to temperature and in antiphase if due to velocity. We can then state that the observed oscillations at low frequencies are essentially due to temperature oscillations. It follows from the previous considerations that the analysis of the phase difference spectra between the H$`\alpha `$ red wing and white light is essentially a study of the correlation between temperature oscillations at slightly different levels ($`\mathrm{\Delta }h100`$ km in the quiet average photosphere). For internetwork areas the extremely high coherence makes the $`\mathrm{\Delta }\mathrm{\Phi }`$ value (5 - 10°) highly significant (see Table 4) and strongly suggests that the observed power is due to oscillations. A positive value of $`\mathrm{\Delta }\mathrm{\Phi }`$ between two layers with decreasing heights, indicates the presence of waves directed radially inward. For the internetwork areas, free from magnetic fields, at the spatial frequency of about 3Mm<sup>-1</sup> used in our analysis and within the considered frequency range, these waves might be interpreted as gravity waves (Straus (1995)). This same interpretation has been adopted for internetwork by Rutten (2000), who reported a similar value of $`\mathrm{\Delta }\mathrm{\Phi }`$ between the intensity of two continuum levels observed by TRACE. Downward directed gravity waves had been proposed earlier by Staiger et al. (1984) to explain similar phase differences for velocity signatures at photospheric levels. At chromospheric levels, the most significant peak in the power spectrum of the H$`\alpha `$ center intensity appears at 2.2 mHz for NBPs and is not related to the oscillations present at photospheric levels ( C $`=0.45`$ between H$`\alpha `$ center /H$`\alpha `$ $`+1.5`$ Å, see Table 4). This means that at chromospheric levels the NBPs experience a new regime of oscillations that seem to be independent from what happens in photosphere. ### 6.2 $`p`$-mode frequency (3.0 - 4.0 mHz) In the $`p`$-mode range of frequencies, for internetwork there is a 10°phase lag between the two H$`\alpha `$ wings (see Table 4) that can be due to different coupling of velocity and intensity fluctuations in the two wings. This effect has been extensively studied for photospheric lines by Cavallini et al. (1987) and Alamanni et al. (1990). As described in Sect. 5.1, the presence of a peak at these frequencies in both the magnetic flux and velocity power spectra could betray the presence of MHD waves in the network structures. In the limit of ideal MHD, a definite phase relation between velocity and magnetic field variations is expected as signature of Alfvèn waves ($`\mathrm{\Delta }\mathrm{\Phi }=0\mathrm{°}`$) or magnetoacoustic waves ($`\mathrm{\Delta }\mathrm{\Phi }=90\mathrm{°}`$, with v leading B. See Ulrich (1996)). However, we find a very low coherence ($`<`$ 0.4) at all frequencies in the NBPs, i.e. the magnetic fluctuations are not related to the velocity ones, at least with the present sensitivity and resolution. ## 7 Discussion and conclusions The observations presented in this paper allowed us to define the characteristics of network bright points at different atmospheric heights, and to compare them with those of the surrounding internetwork areas. We improved on the existing statistics using a good-sized sample of NBPs, and the same number of “test” internetwork areas, defined in a comparable way. The method we adopted to study the temporal evolution of NBPs insures that each bright structure is properly followed in time and position at each height. In fact, the evaluation of the light curves and their properties after a spatial averaging over a well defined area guarantees that we are studying the same NBP at all heights, and avoids the problem (first pointed out by Lites (1994)) of a possible structure displacement due to the magnetic field inclination. Given the characteristic horizontal size of the NBPs, the analysis and the comparison of power spectra and phase differences concern the propagation of waves pertaining to a horizontal wavenumber of about 3 Mm<sup>-1</sup>. The quasi-simultaneous series of NaD<sub>2</sub> images and of MDI maps allowed us to establish for the first time a correspondence between NaD<sub>2</sub> bright network and magnetic network at high spatial and temporal resolution. A correspondence between bright chromospheric structures (Ca II, Ly$`\alpha `$, Mg I and UV continuum) and magnetic structures had been observed before, but not at this high temporal resolution. We also established for the NBPs a quantitative relationship between the Na excess and the corresponding absolute value of magnetic flux density. This relationship is best expressed by a power law with an exponent very close to the one found by Schrijver et al. (1989, 1996) for the Ca II - K excess, and indicates that the emission in NaD<sub>2</sub> may be used as a proxy for the magnetic flux density. The NBPs considered in this work have the following properties: - are bright in the Ca II wings and in the Ca II K<sub>2</sub> peaks; - are visible in the NaD<sub>2</sub> images for about 1 hr; - coincide spatially with the magnetic structures; - are nearby or within a lower activity region. The general characteristics found for these NBPs do not differ from the ones derived in absolutely quiet regions (Deubner & Fleck (1990), Lites et al. (1993)). Our results referring to photospheric and chromospheric properties are so summarized: At photospheric levels: No difference is detected between network and internetwork power spectra, either in intensity or in velocity, within the limits of sensitivity and accuracy of the instruments used for this work. The phase difference spectra between photospheric signatures in general do not show different characteristics for network or internetwork. However, when analyzing the phase difference between H$`\alpha `$ red wing and white light images($`\mathrm{\Delta }h100`$ km), we find $`5\mathrm{°}\mathrm{\Delta }\mathrm{\Phi }`$ $`10\mathrm{°}`$ in the frequency window 1.5 - 2.5 mHz and in the internetwork. (The $`\mathrm{\Delta }\mathrm{\Phi }`$ value is more uncertain in the network, due to a lower coherence value). A phase lag of this amplitude and sign is usually considered a signature of gravity waves directed radially inward. A possible explanation for their origin might be sought in recent models of convection, described as a non-local process driven by cooling at the solar surface rather than by heating from the lower layers (Spruit, 1997). One can imagine that the downward flowing cooled plasma can trigger some inward directed waves, and hence justify the fact that the external layers “lead” the deeper layers. The general inhibition of convection in magnetic structures might be the reason for the lack of this signature in network points. The power spectrum of the magnetic flux variations in NBPs shows a small but significant peak around 3 mHz, that could be related to a “transformation” of acoustic waves into MHD waves. However, the phase difference and coherence spectra between magnetic flux and velocity (B$``$V) for the NBPs indicate a very low correlation between the two signals so we cannot conclude anything on the presence of MHD waves within the network points. At chromospheric levels: Network and internetwork areas have a rather different behaviour in the power spectra. We do not see any evidence for the typical chromospheric period of 3 minutes (but it must be reminded that they are best seen in velocity variations rather than intensity). In the low chromospheric levels, where NaD<sub>2</sub> originates, the NBPs power spectrum is compressed at all frequencies if compared to the internetwork, while in the high chromosphere, where H$`\alpha `$ originates, the power of NBPs is higher than the one of internetwork. This opposite effect may be an indication that the magnetic field disturbs and reduces the amplitude of oscillations already present in the low chromosphere while it assumes a leading rôle in the high chromosphere. In the layers contributing to the NaD<sub>2</sub> emission it seems that the oscillations present in network points change regime with respect to both the photosphere and the high chromosphere and we think that it would be important to perform observations of NBPs in tha Na line, with high spectral resolution. Unfortunately we cannot analyze the phase difference spectrum for NaD<sub>2</sub> intensity fluctuations with respect to others formed at different layers, since the NaD<sub>2</sub> intensity fluctuations, measured with the UBF filter (FWHM$`=0.2`$Å), are more related to velocity than to temperature fluctuations (see Sect 5.2). The power spectrum of H$`\alpha `$ intensity in NBPs has the more relevant peak at 2.2 mHz, but this signal is not correlated with the photospheric fluctuations, as indicated by the very low coherence measured at all frequencies between the H$`\alpha `$ core and the blue and red wings. We can then confirm, using a larger sample of NBPs, the presence of the peak found by Lites et al. (1993) around 2 mHz in the power spectrum of K3 velocity fluctuations for one network point. Kalkofen (1997) and Hasan & Kalkofen (1999) proposed an explanation for this peak in terms of transverse magneto-acoustic waves in magnetic flux tubes, excited by granular buffeting in the solar photosphere. In their model the low coherence between photospheric and chromospheric signatures could be explained by a partial conversion of the transverse waves to longitudinal modes in the higher chromosphere. A general result of our analysis, valid from the low photosphere to the high chromosphere, is that the NBPs always show a coherence lower than the internetwork, pointing out that the presence of the magnetic field changes the propagation regime of waves with respect to the non-magnetic regions. ###### Acknowledgements. The authors are indebted to the NSO/SP-DST staff for the generous telescope time allocation and the unvaluable help during the observations. The authors express their thanks to the MDI team (P.I. P.H. Scherrer) for the efficient support during the observing run and their dedication in operating this instrument. MDI is part of SOHO, the Solar and Heliospheric Observatory, mission of international cooperation between ESA and NASA. We woud like to thank Thomas Straus for his comments during the preparation of this paper and for fruitful discussion on the phase difference problem. ## Appendix A A relationship between Na excess and magnetic field A quantitative relationship between the CaII K line core intensity and the absolute value of the magnetic flux density has been clearly established by several studies (Skumanich et al (1975); Schrijver et al. (1989); Nindos & Zirin (1998)). The best agreement with the data is given by a power law relation with an exponent of 0.6 (Schrijver et al. (1989)). Given the excellent spatial coincidence of the sodium network points and the magnetic structures present in our data (see Sect. 3.3, we searched for such a quantitative relationship also for this case. Fig. 7 shows a scatterplot of the sodium “excess” for the 11 NBPs, vs. the corresponding absolute values of magnetic flux density. As done by Schrijver et al. (1989, 1996) for the CaII K emission, we subtracted to the NaD<sub>2</sub> intensities a threshold equal to the average intensity in the quiet areas ($`10^4`$ in arbitrary units). The graph was obtained plotting the temporal average of both the sodium intensity and magnetic flux. It does not display, hence, any temporal variation of either quantity, but only a general trend among the persistent structures. Five-minutes oscillations do not play a role in this relationship, nor do they contribute to the scatter of the points. Saturation, as reported by Schrijver et al. (1989) is not apparent in our data, but it must be remarked that the maximum value of the magnetic flux density for the average quantities is below 400 G, the “critical” value indicated by those authors. The sodium excess is best fitted by a power law of the type $`\mathrm{\Delta }`$I $`\mathrm{\Phi }^\beta `$, with $`\beta `$=0.58 $`\pm `$0.1. The exponent is very close to $`\beta `$=0.6$`\pm `$0.1 found by Schrijver et al. (1989, 1996) for the CaII K excess. These results indicate that the emission in the center of the NaD<sub>2</sub> line is also a good proxy for the magnetic flux density and, at least for values of magnetic flux density smaller than a few hundred G, its use is equivalent to that of the CaII K emission.
warning/0004/quant-ph0004042.html
ar5iv
text
# Two-mode Nonlinear Coherent States ## I Introduction Coherent states of a simple harmonic oscillator have considerable applications in the study of quantum optics. Recently another type of coherent states, nonlinear coherent states (NLCSs) as well as their superpositions, have been introduced and studied. The so-called nonlinear coherent states are defined as the right-hand eigenstates of the product of the boson annihilation operator $`a`$ and a non-constant function of number operator $`\widehat{N}=a^{}a,`$ $$f(\widehat{N})a|\alpha ,f=\alpha |\alpha ,f,$$ (1) where $`f(\widehat{N})`$ is an operator-valued function of the number operator and $`\alpha `$ is a complex eigenvalue. The ordinary coherent states $`|\alpha `$ are recovered for the special choice of $`f(\widehat{N})=1.`$ A class of NLCSs can be realized physically as the stationary states of the center-of-mass motion of a trapped ion. These NLCSs exhibit various non-classical features like squeezing and self-splitting. The notion of the NLCS was generalized to the two-photon NLCS. One two-photon NLCS is the squeezed vacuum state and another is squeezed first Fock state. We want to generalize the notion of the NLCS to two-mode case in this paper. In a previous paper, we have given an exponential form of one-mode NLCS $`|\alpha ,f`$ and proved that photon-added one-mode NLCSs are still NLCSs. In analogy to the defination of the one-mode NLCS, the two-mode nonlinear coherent state(TMNLCS) is defined as $$f(\widehat{N}_a,\widehat{N}_b)ab|\alpha ,f,q=\alpha |\alpha ,f,q,$$ (2) where $`a`$ and $`b`$ are boson annihilation operators; $`f(\widehat{N}_a,\widehat{N}_b)`$ is the function of the number operator $`\widehat{N}_a=a^{}a`$ and $`\widehat{N}_b=b^{}b;`$ $`q`$ is the photon number difference between two modes of the field. In section II, we give the expansion and exponential form of the TMNLCS. In section III and IV, we discuss the photon-added TMNLCS and photon-subtracted TMNLCS. Section V introduces the parity pair coherent states and parity two-mode Perelomov cohernet states as interesting examples of the TMNLCSs. A conclusion is given in section VI. ## II Expansion and exponential form of the TMNLCS. We next determine the solution to the eigenvalue equation (Eq.(2)). The TMNLCS has the form $$|\alpha ,f,q=\underset{n=0}{\overset{\mathrm{}}{}}C_n|n+q,n.$$ (3) Substituting Eq.(3) into Eq.(2), we find the recursion relation among the coefficients $`C_n`$’s $`C_{n+1}`$ $`=`$ $`{\displaystyle \frac{\alpha }{f(n+q,n)\sqrt{(n+1)(n+q+1)}}}C_n,`$ (4) $`C_n`$ $`=`$ $`\alpha ^n\sqrt{{\displaystyle \frac{q!}{n!(n+q)!}}}\left[{\displaystyle \underset{m=1}{\overset{n}{}}}{\displaystyle \frac{1}{f(m+q1,m1)}}\right]C_0.`$ (5) Thus the expansion of the TMNLCS is obtained as $`|\alpha ,f,q`$ $`=`$ $`C_0{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}\alpha ^n\sqrt{{\displaystyle \frac{q!}{n!(n+q)!}}}\left[{\displaystyle \underset{m=1}{\overset{n}{}}}{\displaystyle \frac{1}{f(m+q1,m1)}}\right]|n+q,n`$ (6) $`=`$ $`C_0{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{\alpha ^nq!a^nb^n}{n!(n+q)!}}\left[{\displaystyle \underset{m=1}{\overset{n}{}}}{\displaystyle \frac{1}{f(m+q1,m1)}}\right]|q,0.`$ (7) One can show that $$[g(\widehat{N}_a,\widehat{N}_b)a^{}b^{}]^n=a^nb^{}{}_{}{}^{n}\underset{m=1}{\overset{n}{}}g(\widehat{N}_a+m,\widehat{N}_b+m).$$ (8) Here $`g(\widehat{N}_a,\widehat{N}_b)`$ is an arbitrary function of $`\widehat{N}_a`$ and $`\widehat{N}_b.`$ Then using Eq.(7) with $$g(\widehat{N}_a,\widehat{N}_b)=\frac{\alpha }{f(\widehat{N}_a1,\widehat{N}_b1)\widehat{N}_a},$$ (9) the state $`|\alpha ,f,q`$ is finally written in the exponential form $`|\alpha ,f,q`$ $`=`$ $`C_0{\displaystyle \underset{n=0}{\overset{\mathrm{}}{}}}{\displaystyle \frac{[g(\widehat{N}_a,\widehat{N}_b)a^{}b^{}]^n}{n!}}|q,0`$ (10) $`=`$ $`C_0\mathrm{exp}[g(\widehat{N})a^{}b^{}]|q,0`$ (11) $`=`$ $`C_0\mathrm{exp}[{\displaystyle \frac{\alpha }{f(\widehat{N}_a1,\widehat{N}_b1)\widehat{N}_a}}a^{}b^{}]|q,0.`$ (12) Here, $`C_0`$ can be determined as $$C_0=\left\{\underset{n=0}{\overset{\mathrm{}}{}}\frac{q!|\alpha |^{2n}}{n!(n+q)!}\left[\underset{m=1}{\overset{n}{}}\frac{1}{f(m+q1,m1)}\right]^2\right\}^{1/2}.$$ (13) In fact, by direct verification, we have $$[f(\widehat{N}_a,\widehat{N}_b)ab,\frac{1}{f(\widehat{N}_a1,\widehat{N}_b1)\widehat{N}_a}a^{}b^{}]=1.$$ (14) From the above commutaion relations and the definition of the TMNLCS, we can also obtain the exponential form of the TMNLCS. Next we want to give some examples of the TMNLCS. The pair coherent state is an important correlated two-mode field. It is defined as $$ab|\zeta ,q=\zeta |\zeta ,q,$$ (15) which is the state $`|\alpha ,f,q`$ with $`f(\widehat{N}_a,\widehat{N}_b)1.`$ Then from Eq.(2), (9) and (12), the exponential form of the pair coherent state is $$|\zeta ,q=\mathrm{exp}(\frac{\zeta }{\widehat{N}_a}a^{}b^{})|q,0$$ (16) up to a normalization constant. Another important correlated two-mode field is the two-mode Perelomov coherent state, which is defined as $$|\xi ,q=\mathrm{exp}(\xi a^{}b^{}\xi ^{}ab)|q,0$$ (17) We find that the state $`|\xi ,q`$ satisfies the equation $$\frac{2}{\widehat{N}_a+\widehat{N}_b+q+2}ab|\xi ,q=\frac{\xi \mathrm{tanh}(|\xi |)}{|\xi |}|\xi ,q,$$ (18) which shows that two-mode Perelomov coherent states are TMNLCSs with the nonlinear function $`2/(\widehat{N}_a+\widehat{N}_b+q+2).`$ In the subspace $`F_q`$span$`\{|n+q,n|n=0,1,2\mathrm{}\}`$ of the two-mode Fock space, Eq.(15) can be simplified as $$\frac{1}{\widehat{N}_a+1}ab|\xi ,q=\frac{\xi \mathrm{tanh}(|\xi |)}{|\xi |}|\xi ,q,$$ (19) Then From Eq.(9) , the exponential form of the state $`|\xi ,q`$ is obtained as $$|\xi ,q=\mathrm{exp}[\frac{\xi \mathrm{tanh}(|\xi |)}{|\xi |}a^{}b^{}]|q,0$$ (20) up to a normalization constant. Actually using the well-known identity $`\mathrm{exp}(\xi a^{}b^{}\xi ^{}ab)`$ (21) $`=`$ $`\mathrm{exp}[{\displaystyle \frac{\xi \mathrm{tanh}(|\xi |)}{|\xi |}}a^{}b^{}][{\displaystyle \frac{1}{\mathrm{cosh}(|\xi |)}}]^{\widehat{N}_a+\widehat{N}_b+1}\mathrm{exp}[{\displaystyle \frac{\xi ^{}\mathrm{tanh}(|\xi |)}{|\xi |}}ab],`$ (22) one can directly verifiy Eq.(17). One special type of two-mode Perelomov coherent state, $`|\xi ,0,`$ is the eigenstate of the two-mode phase operator $`1/\sqrt{(1+\widehat{N}_a)(1+\widehat{N}_b)}ab`$ (also called two-mode squeezed vacuum state). The eigenstate of the two-mode phase operator seems different from the state $`|\xi ,0.`$ However, since $`q=0,`$ one can show that the operator $`1/(1+\widehat{N}_a)ab`$ is identical to the two-mode phase operator in the subspace $`F_0`$span$`\{|n,n|n=0,1,2\mathrm{}\}`$ of the two-mode Fock space. ## III Photon-added TMNLCS The photon-added quantum states were first introduced by Agarwal and Tara as photon-added coherent states. Sivakumar found that the photon-added coherent states are NLCSs. As a generalization of his work, we have given a more general result that the photon-added NLCSs are still NLCSs. In this section, we consider the photon-added TMNLCS which is defined as $$|m,n,\alpha ,f,q=\frac{a^mb^n|\alpha ,f,q}{\alpha ,f,q|b^na^ma^mb^n|\alpha ,f,q}.$$ (23) Multiplying both sides of Eq.(2) by $`a^mb^n`$ from the left yields $$a^mb^nf(\widehat{N}_a,\widehat{N}_b)ab|\alpha ,f,q=\alpha a^mb^n|\alpha ,f,q.$$ (24) By the fact $`a^mf(\widehat{N}_a,\widehat{N}_b)`$ $`=`$ $`f(\widehat{N}_am,\widehat{N}_b)a^m,`$ (25) $`b^nf(\widehat{N}_a,\widehat{N}_b)`$ $`=`$ $`f(\widehat{N}_a,\widehat{N}_bn)b^n,`$ (26) Eq.(20) can be written as $`f(\widehat{N}_am,\widehat{N}_bn)(\widehat{N}_am+1)(\widehat{N}_bn+1)a^{(m1)}b^{(n1)}|\alpha ,f,q`$ (27) $`=`$ $`\alpha a^mb^n|\alpha ,f,q`$ (28) Multiplying the both sides of the above equation by $`ab`$ from the left, we get $`f(\widehat{N}_am+1,\widehat{N}_bn+1)(\widehat{N}_am+2)(\widehat{N}_bn+2)`$ (30) $`\times aba^{(m1)}b^{(n1)}|\alpha ,f,q`$ $`=`$ $`\alpha (\widehat{N}_a+1)(\widehat{N}_b+1)a^{(m1)}b^{(n1)}|\alpha ,f,q`$ (31) Now we replace $`m1(n1)`$ by $`m(n)`$ and multiply the both sides of Eq.(24) from the left by the operator $`[(\widehat{N}_a+1)(\widehat{N}_b+1)]^1`$, the following equation is obtained as $`f(\widehat{N}_am,\widehat{N}_bn)(1{\displaystyle \frac{m}{\widehat{N}_a+1}})(1{\displaystyle \frac{n}{\widehat{N}_b+1}})ab|m,n,\alpha ,f,q`$ (32) $`=`$ $`\alpha |m,n,\alpha ,f,q`$ (33) From Eq.(25), we can see that that the photon-added TMNLCSs are still TMNLCSs with the corresponding nonlinear function $`f(\widehat{N}_am,\widehat{N}_bn)[1m/(\widehat{N}_a+1)][1n/(\widehat{N}_b+1)]`$. The above discussion can be directly generalized to the multi-mode case, but here we restrict us to the two-mode case. Lu and Guo studied the nonclassical properties of the photon-added pair coherent states. As seen from the defination of pair coherent state and Eq.(25), we conclude that the photon-added pair coherent states are TMNLCSs with the nonlinear function $`[1m/(\widehat{N}_a+1)][1n/(\widehat{N}_b+1)].`$ The photon-added two-mode Perelomov’s coherent states are introduced and studied. From Eq.(16) and (25) we see that the photon-added two-mode Perelomov coherent states are TMNLCSs with the nonlinear function $`[1m/(\widehat{N}_a+1)][1n/(\widehat{N}_b+1)]/(\widehat{N}_am+1).`$ ## IV Photon-subtracted TMNLCS By analogy to the defination of photon-added TMNLCS, the photon-subtracted TMNLCS is defined as $$|m,n,\alpha ,f,q=\frac{a^mb^n|\alpha ,f,q}{\alpha ,f,q|b^{}_{}{}^{}na^{}_{}{}^{}ma^mb^n|\alpha ,f,q}.$$ (34) Multiplying both sides of Eq.(2) by $`a^mb^n`$ from the left yields $$a^mb^nf(\widehat{N}_a,\widehat{N}_b)ab|\alpha ,f,q=\alpha a^mb^n|\alpha ,f,q.$$ (35) Using the identity $$a^mb^nf(\widehat{N}_a,\widehat{N}_b)=f(\widehat{N}_a+m,\widehat{N}_b+m)a^mb^n,$$ (36) we obtain $`f(\widehat{N}_a+m,\widehat{N}_b+n)ab|m,n,\alpha ,f,q`$ (37) $`=`$ $`\alpha |m,n,\alpha ,f,q.`$ (38) From the above equation, we see that the photon-subtracted state is a TMNLCS with the nonlinear function $`f(\widehat{N}_a+m,\widehat{N}_b+n).`$ Then photon-subtracted pair coherent states are still pair coherent states, and photon-subtracted two-mode Perelomov coherent states are TMNLCSs with the nonlinear function $`1/(\widehat{N}_a+1+m).`$ ## V Parity Pair coherent states By analogy to the defination in the references, we define the parity operator $`\mathrm{\Pi }`$ in the subspace $`F_q`$ as $$\mathrm{\Pi }=(1)^{\widehat{N}_b},\text{ }\mathrm{\Pi }^2=1,\text{ }\mathrm{\Pi }^{}=\mathrm{\Pi }.$$ (39) Now we solve the eigenvalue equation $$\mathrm{\Pi }ab|\zeta ,q_\mathrm{\Pi }=\zeta |\zeta ,q_\mathrm{\Pi }.$$ (40) and call the state $`|\zeta ,q_\mathrm{\Pi }`$ the parity pair coherent state following the term of the parity harmonic oscillator coherent state. Comparing Eq.(2) and Eq.(31), we know that the parity pair coherent state is a TMNLCS with the nonlinear function $`(1)^{\widehat{N}_b}.`$ From Eq.(6) and (31), the expansion of the parity pair coherent state is easily obtained as $$|\zeta ,q_\mathrm{\Pi }=\underset{n=0}{\overset{\mathrm{}}{}}\sqrt{\frac{q!}{n!(n+q)!}}\zeta ^n(1)^{n(n1)/2}|n+q,n.$$ (41) The state $`|\zeta ,q_\mathrm{\Pi }`$ can be rewritten as $$|\zeta ,q_\mathrm{\Pi }=\frac{1}{\sqrt{2}}\left(e^{i\pi /4}|i\zeta ,q+e^{i\pi /4}|i\zeta ,q\right),$$ (42) which is a superposition of two pair coherent states with phase difference $`\pi .`$ The form of this parity coherent states is similar to that of the parity harmonic oscillator coherent states. The parity pair coherent state can be genearated in the nonlinear Kerr medium. The Hamiltonian describing the Kerr medium is $$H=\omega \widehat{N}_b+\gamma \widehat{N}_b(\widehat{N}_b1),$$ (43) where $`\omega `$ is the frequency of mode $`b,`$ and the parameter $`\gamma `$ is proportional to the third-order nonlinear susceptibility $`\chi ^3.`$ The evolution operator in the interaction picture is $$U=e^{i\gamma t\widehat{N}_b(\widehat{N}_b1)}$$ (44) We assume the intial state is pair coherent state and let $`\gamma t=\pi ,`$ then the final state is just the parity pair coherent state (Eq.(32)) Analogously we can define the parity two-mode Perelomov coherent states as the TMNLCSs with the nonlinear function $`(1)^{\widehat{N}_b}/(\widehat{N}_a+1)`$ and the state can be written as $$|\xi ,q_\mathrm{\Pi }=\frac{1}{\sqrt{2}}\left(e^{i\pi /4}|i\xi ,q+e^{i\pi /4}|i\xi ,q\right),$$ (45) which is the superposition of two two-photon Perelomov cohererent states. Similarly the parity two-mode Perelomov coherent states can be generated in the Kerr medium. ## VI Conclusion In conclusion, we have introduced the two-mode nonlinear coherent states. The Perelomov coherent states, parity pair coherent states and parity Perelomov coherent states are examples of the two-mode nonlinear coherent states. An interesting class of TMNLCSs are the photon-added or photon-subtracted two-mode nonlinear coherent states. The corresponding nonlinear functions are obtained for these states. Special examples, photon-added (subtracted) pair coherent states and photon-added (subtracted) two-mode Perelomov coherent states are discussed and the correpoding nonlinear functions are obtained. The parity coherent states can be generated in the Kerr medium when the correponding coherent states are the input states. They are the superpositions of two corresponding coherent states. The notation of the TMNLCS can be directly generalized to the multi-mode case. The work on the multi-mode nonlinear coherent states are in progress. Acknowledgement: The author is grateful for many valuble discussions with Dr. Barry C. Sanders during his visit to Insitute of Physics, Chinese Academy of Sciences. This work is supported in part by the National Science Foundation of China with grant number:19875008.
warning/0004/hep-th0004097.html
ar5iv
text
# References NDA-FP-73 April 2000 BRANE WORLD INFLATION INDUCED BY QUANTUM EFFECTS Shin’ichi NOJIRI<sup>1</sup><sup>1</sup>1email: nojiri@cc.nda.ac.jp, Sergei D. ODINTSOV<sup>2</sup><sup>2</sup>2 On leave from Tomsk State Pedagogical University, RUSSIA. email: odintsov@ifug5.ugto.mx, Department of Applied Physics National Defence Academy, Hashirimizu Yokosuka 239, JAPAN $`\mathrm{}`$ Instituto de Fisica de la Universidad de Guanajuato Apdo.Postal E-143, 37150 Leon, Gto., MEXICO ABSTRACT We consider brane-world universe where an arbitrary large $`N`$ quantum CFT is living on the domain wall. This corresponds to implementing of Randall-Sundrum compactification within the context of AdS/CFT correspondence. Using anomaly induced effective action for domain wall CFT the possibility of self-consistent quantum creation of 4d de Sitter wall Universe (inflation) is demonstrated. In case of maximally SUSY Yang-Mills theory the exact correspondence with radius and effective tension found by Hawking-Hertog-Reall is obtained. The hyperbolic wall Universe may be induced by quantum effects only for exotic matter (higher derivatives conformal scalar) which has unusual sign of central charge. The idea that we live on the brane (where four-dimensional gravity is recovered) embedded in higher-dimensional spacetime initiated enermous activity in the study of brane worlds. The attempts to realize the inflationary brane-worlds have been made in refs. (and refs.therein). However, there is arbitrariness here. It is caused by the fact that details of scenario depend on the matter content (equation of state), etc. The inflationary brane-world scenario realized due to quantum effects of brane matter looks more attractive and universal. Such idea has been recently suggested in refs. (see also) where addition of conformal quantum brane matter to complete effective action has been made. That corresponds to implementing of RS compactification within the context of renormalization group flow in AdS/CFT set-up. In particular, using conformal anomaly induced effective action the indication to a possibility of quantum creation of de Sitter or Anti-de Sitter wall in 5d AdS Universe has been demonstrated in ref.. As brane quantum matter the maximally SUSY Yang-Mills theory on the wall has been considered. In ref. the role of conformal anomaly in inducing of effective tension which is responsible for a de Sitter geometry of domain wall has been stressed and the corresponding effective tension has been calculated. The extensive calculation of (lorentzian) graviton correlator has been also made there. It was shown that domain wall CFT may significally suppress the metric perturbations. In the present Letter we give the details of such proposal (addition of brane quantum matter to RS scenario) with applications to 4d cosmology in the elegant form, using anomaly induced effective action. As a result one can consider the arbitrary content of CFT living on the wall. Moreover, the formalism is applied not only to 4d de Sitter wall but also to 4d hyperbolic wall in 5d Anti-de Sitter Universe or 4d conformally flat Universe. The equivalence of two approaches is explicitly proved: in case of spherical geometry the same effective tension and same equation for radius of de Sitter wall Universe is obtained. We also show that hyperbolic wall Universe may be realized due to quantum effects only for exotic matter (higher derivatives scalar). Note that consideration of large $`N`$ conformal field theory living on the domain wall justifies such approach to brane-world quantum cosmology as then quantum contribution is essential. We consider the spacetime whose boundary is 4 dimensional sphere $`\mathrm{S}_4`$, which can be identified with a D3-brane. The bulk part is given by 5 dimensional Euclidean anti de Sitter space $`\mathrm{AdS}_5`$ $$ds_{\mathrm{AdS}_5}^2=dy^2+\mathrm{sinh}^2\frac{y}{l}d\mathrm{\Omega }_4^2.$$ (1) Here $`d\mathrm{\Omega }_4^2`$ is given by the metric of $`\mathrm{S}_4`$ with unit radius. One also assumes the boundary (brane) lies at $`y=y_0`$ and the bulk space is given by gluing two regions given by $`0y<y_0`$. We start with the action $`S`$ which is the sum of the Einstein-Hilbert action $`S_{\mathrm{EH}}`$, the Gibbons-Hawking surface term $`S_{\mathrm{GH}}`$, the surface counter term $`S_1`$ and the trace anomaly induced action $`W`$<sup>3</sup><sup>3</sup>3For the introduction to anomaly induced effective action in curved space-time (with torsion), see section 5.5 in .: $`S`$ $`=`$ $`S_{\mathrm{EH}}+S_{\mathrm{GH}}+2S_1+W`$ (2) $`S_{\mathrm{EH}}`$ $`=`$ $`{\displaystyle \frac{1}{16\pi G}}{\displaystyle d^5x\sqrt{g_{(5)}}\left(R_{(5)}+\frac{12}{l^2}\right)}`$ (3) $`S_{\mathrm{GH}}`$ $`=`$ $`{\displaystyle \frac{1}{8\pi G}}{\displaystyle d^4x\sqrt{g_{(4)}}_\mu n^\mu }`$ (4) $`S_1`$ $`=`$ $`{\displaystyle \frac{3}{8\pi G}}{\displaystyle d^4x\sqrt{g_{(4)}}}`$ (5) $`W`$ $`=`$ $`b{\displaystyle d^4x\sqrt{\stackrel{~}{g}}\stackrel{~}{F}A}`$ (6) $`+b^{}{\displaystyle }d^4x\{A[2\stackrel{~}{\mathrm{}}^2+\stackrel{~}{R}_{\mu \nu }\stackrel{~}{}_\mu \stackrel{~}{}_\nu {\displaystyle \frac{4}{3}}\stackrel{~}{R}\stackrel{~}{\mathrm{}}^2+{\displaystyle \frac{2}{3}}(\stackrel{~}{}^\mu \stackrel{~}{R})\stackrel{~}{}_\mu ]A`$ $`+(\stackrel{~}{G}{\displaystyle \frac{2}{3}}\stackrel{~}{\mathrm{}}\stackrel{~}{R})A\}`$ $`{\displaystyle \frac{1}{12}}\left\{b^{\prime \prime }+{\displaystyle \frac{2}{3}}(b+b^{})\right\}{\displaystyle d^4x\left[\stackrel{~}{R}6\stackrel{~}{\mathrm{}}A6(\stackrel{~}{}_\mu A)(\stackrel{~}{}^\mu A)\right]^2}.`$ Here the quantities in the 5 dimensional bulk spacetime are specified by the suffices <sub>(5)</sub> and those in the boundary 4 dimensional spacetime are by <sub>(4)</sub>. The factor $`2`$ in front of $`S_1`$ in (2) is coming from that we have two bulk regions which are connected with each other by the brane. In (4), $`n^\mu `$ is the unit vector normal to the boundary. In (6), one chooses the 4 dimensional boundary metric as $$g_{(4)}^{}{}_{\mu \nu }{}^{}=\mathrm{e}^{2A}\stackrel{~}{g}_{\mu \nu }$$ (7) and we specify the quantities with $`\stackrel{~}{g}_{\mu \nu }`$ by using $`\stackrel{~}{}`$. $`G`$ ($`\stackrel{~}{G}`$) and $`F`$ ($`\stackrel{~}{F}`$) are the Gauss-Bonnet invariant and the square of the Weyl tensor, which are given as <sup>4</sup><sup>4</sup>4We use the following curvature conventions: $`R`$ $`=`$ $`g^{\mu \nu }R_{\mu \nu }`$ $`R_{\mu \nu }`$ $`=`$ $`R_{\mu \lambda \nu }^\lambda `$ $`R_{\mu \rho \nu }^\lambda `$ $`=`$ $`\mathrm{\Gamma }_{\mu \rho ,\nu }^\lambda +\mathrm{\Gamma }_{\mu \nu ,\rho }^\lambda \mathrm{\Gamma }_{\mu \rho }^\eta \mathrm{\Gamma }_{\nu \eta }^\lambda +\mathrm{\Gamma }_{\mu \nu }^\eta \mathrm{\Gamma }_{\rho \eta }^\lambda `$ $`\mathrm{\Gamma }_{\mu \lambda }^\eta `$ $`=`$ $`{\displaystyle \frac{1}{2}}g^{\eta \nu }\left(g_{\mu \nu ,\lambda }+g_{\lambda \nu ,\mu }g_{\mu \lambda ,\nu }\right).`$ $`G`$ $`=`$ $`R^24R_{ij}R^{ij}+R_{ijkl}R^{ijkl}`$ $`F`$ $`=`$ $`{\displaystyle \frac{1}{3}}R^22R_{ij}R^{ij}+R_{ijkl}R^{ijkl},`$ (8) In the effective action (6), with $`N`$ scalar, $`N_{1/2}`$ spinor and $`N_1`$ vector fields, $`b`$, $`b^{}`$ and $`b^{\prime \prime }`$ are $$b=\frac{(N+6N_{1/2}+12N_1)}{120(4\pi )^2},b^{}=\frac{(N+11N_{1/2}+62N_1)}{360(4\pi )^2},b^{\prime \prime }=0$$ (9) but in principle, $`b^{\prime \prime }`$ may be changed by the finite renormalization of local counterterm in gravitational effective action. As we shall see later, the term proportional to $`\left\{b^{\prime \prime }+\frac{2}{3}(b+b^{})\right\}`$ in (6), and therefore $`b^{\prime \prime }`$, does not contribute to the equations of motion (see (29)). For $`𝒩=4`$ $`SU(N)`$ SYM theory $`b=b^{}=\frac{N^21}{4(4\pi )^2}`$. We should also note that $`W`$ in (6) is defined up to conformally invariant functional, which cannot be determined from only the conformal anomaly. The conformally flat space is a pleasant exclusion where anomaly induced effective action is defined uniquely. However, one can argue that such conformally invariant functional gives next to leading contribution as mass parameter of regularization may be adjusted to be arbitrary small (or large).. The metric of $`\mathrm{S}_4`$ with the unit radius is given by $$d\mathrm{\Omega }_4^2=d\chi ^2+\mathrm{sin}^2\chi d\mathrm{\Omega }_3^2.$$ (10) Here $`d\mathrm{\Omega }_3^2`$ is described by the metric of 3 dimensional unit sphere. If we change the coordinate $`\chi `$ to $`\sigma `$ by $$\mathrm{sin}\chi =\pm \frac{1}{\mathrm{cosh}\sigma },$$ (11) one obtains $$d\mathrm{\Omega }_4^2=\frac{1}{\mathrm{cosh}^2\sigma }\left(d\sigma ^2+d\mathrm{\Omega }_3^2\right).$$ (12) On the other hand, the metric of the 4 dimensional flat Euclidean space is given by $$ds_{4\mathrm{E}}^2=d\rho ^2+\rho ^2d\mathrm{\Omega }_3^2.$$ (13) Then by changing the coordinate as $$\rho =\mathrm{e}^\sigma ,$$ (14) one gets $$ds_{4\mathrm{E}}^2=\mathrm{e}^{2\sigma }\left(d\sigma ^2+d\mathrm{\Omega }_3^2\right).$$ (15) For the 4 dimensional hyperboloid with the unit radius, the metric is given by $$ds_{\mathrm{H4}}^2=d\chi ^2+\mathrm{sinh}^2\chi d\mathrm{\Omega }_3^2.$$ (16) Changing the coordinate $`\chi `$ to $`\sigma `$ $$\mathrm{sinh}\chi =\frac{1}{\mathrm{sinh}\sigma },$$ (17) one finds $$ds_{\mathrm{H4}}^2=\frac{1}{\mathrm{sinh}^2\sigma }\left(d\sigma ^2+d\mathrm{\Omega }_3^2\right).$$ (18) We now discuss the 4 dimensional hyperboloid whose boundary is the 3 dimensional sphere $`S_3`$ but we can consider the cases that the boundary is a 3 dimensional flat Euclidean space $`R_3`$ or a 3 dimensional hyperboloid $`H_3`$. We will, however, only consider the case that the boundary is $`S_3`$ since the results for other cases are almost equivalent. Motivated by (1), (12), (15) and (18), one assumes the metric of 5 dimensional space time as follows: $$ds^2=dy^2+\mathrm{e}^{2A(y,\sigma )}\stackrel{~}{g}_{\mu \nu }dx^\mu dx^\nu ,\stackrel{~}{g}_{\mu \nu }dx^\mu dx^\nu l^2\left(d\sigma ^2+d\mathrm{\Omega }_3^2\right)$$ (19) and we identify $`A`$ and $`\stackrel{~}{g}`$ in (19) with those in (7). Then we find $`\stackrel{~}{F}=\stackrel{~}{G}=0`$, $`\stackrel{~}{R}=\frac{6}{l^2}`$ etc. Due to the assumption (19), the actions in (3), (4), (5), and (6) have the following forms: $`S_{\mathrm{EH}}`$ $`=`$ $`{\displaystyle \frac{l^4V_3}{16\pi G}}{\displaystyle }dyd\sigma \{(8_y^2A20(_yA)^2)\mathrm{e}^{4A}`$ (20) $`+(6_\sigma ^2A6(_\sigma A)^2+6)\mathrm{e}^{2A}+{\displaystyle \frac{12}{l^2}}\mathrm{e}^{4A}\}`$ $`S_{\mathrm{GH}}`$ $`=`$ $`{\displaystyle \frac{3l^4V_3}{8\pi G}}{\displaystyle 𝑑\sigma \mathrm{e}^{4A}_yA}`$ (21) $`S_1`$ $`=`$ $`{\displaystyle \frac{3l^3V_3}{8\pi G}}{\displaystyle 𝑑\sigma \mathrm{e}^{4A}}`$ (22) $`W`$ $`=`$ $`V_3{\displaystyle }d\sigma [b^{}A(2_\sigma ^4A8_\sigma ^2A)`$ (23) $`2(b+b^{})(1_\sigma ^2A(_\sigma A)^2)^2].`$ Here $`V_3`$ is the volume or area of the unit 3 sphere. In the bulk, one obtains the following equation of motion from $`S_{\mathrm{EH}}`$ by the variation over $`A`$: $$0=\left(24_y^2A48(_yA)^2+\frac{48}{l^2}\right)\mathrm{e}^{4A}+\frac{1}{l^2}\left(12_\sigma ^2A12(_\sigma A)^2+12\right)\mathrm{e}^{2A}.$$ (24) Then one finds a solution $$A=\mathrm{ln}\mathrm{sinh}\frac{y}{l}\mathrm{ln}\mathrm{cosh}\sigma ,$$ (25) which corresponds to the metric of $`\mathrm{AdS}_5`$ in (1) with (12). There exists also the solution $$A=\frac{y}{l}+\sigma ,$$ (26) which corresponds to (15), and another solution $$A=\mathrm{ln}\mathrm{cosh}\frac{y}{l}\mathrm{ln}\mathrm{sinh}\sigma ,$$ (27) corresponds to (18). One should note that all the metrics in (25), (26) and (27) locally describe the same spacetime, that is the local region of $`\mathrm{AdS}_5`$, in the bulk. As we assume, however, that there is a brane at $`y=y_0`$, the shapes of the branes are different from each other due to the choice of the metric. On the brane at the boundary, one gets the following equation: $`0`$ $`=`$ $`{\displaystyle \frac{48l^4}{16\pi G}}\left(_yA{\displaystyle \frac{1}{l}}\right)\mathrm{e}^{4A}+b^{}\left(4_\sigma ^4A16_\sigma ^2A\right)`$ (28) $`4(b+b^{})\left(_\sigma ^4A+2_\sigma ^2A6(_\sigma A)^2_\sigma ^2A\right).`$ We should note that the contributions from $`S_{\mathrm{EH}}`$ and $`S_{\mathrm{GH}}`$ are twice from the naive values since we have two bulk regions which are connected with each other by the brane. Substituting the bulk solution (25) into (28), one obtains $$0=\frac{48l^3}{16\pi G}\left(\mathrm{coth}\frac{y_0}{l}1\right)\mathrm{sinh}^4\frac{y_0}{l}+24b^{}.$$ (29) Note that eq.(29) does not depend on $`b`$. The effective tension of the domain wall is given by $$\sigma _{\mathrm{eff}}=\frac{3}{4\pi Gl}\mathrm{coth}\frac{y_0}{l}.$$ (30) As in , defining the radius $`R`$ of the brane in the following way $$Rl\mathrm{sinh}\frac{y_0}{l},$$ (31) one can rewrite (29) as $$0=\frac{1}{\pi G}\left(\frac{1}{R}\sqrt{1+\frac{R^2}{l^2}}\frac{1}{l}\right)R^4+8b^{}.$$ (32) This equation generalizes the corresponding result of ref. for the case when the arbitrary amount of quantum conformal matter sits on de Sitter wall. Adopting AdS/CFT correspondence one can argue that in symmetric phase the quantum brane matter comes due to maximally SUSY Yang-Mills theory. As we have $`b^{}\frac{N^2}{4(4\pi )^2}`$ in case of the large $`N`$ limit of $`𝒩=4`$ $`SU(N)`$ SYM theory, we find $$\frac{R^3}{l^3}\sqrt{1+\frac{R^2}{l^2}}=\frac{R^4}{l^4}+\frac{GN^2}{8\pi l^3},$$ (33) which exactly coincides with the result in . This equation has the unique solution for positive radius which defines brane-world de Sitter Universe (inflation) induced by quantum effects. On the other hand, if we substitute the solution (26), corresponding to flat Euclidean brane, into (28), we find that (28) is always (independent of $`y_0`$) satisfied since $`_yA=\frac{1}{l}`$ and $`_\sigma ^2A=0`$. If one substitutes (27), which corresponds to the brane with the shape of the hyperboloid, then $$0=\frac{48l^3}{16\pi G}\left(\mathrm{tanh}\frac{y_0}{l}1\right)\mathrm{cosh}^4\frac{y_0}{l}+24b^{}.$$ (34) We should note that eq.(34) does not depend on $`b`$ again. In order that Eq.(34) has a solution, $`b^{}`$ must be positive, which conflicts with the case of $`𝒩=4`$ $`SU(N)`$ SYM theory or usual conformal matter. In general, however, for some exotic theories, like higher derivative conformal scalar<sup>5</sup><sup>5</sup>5Such higher derivative conformal scalar naturally appears in infra-red sector of quantum gravity., $`b^{}`$ can be positive and one can assume for the moment that $`b^{}>0`$ here<sup>6</sup><sup>6</sup>6For higher derivative conformal scalar $`b=8/120(4\pi )^2`$, $`b^{}=28/360(4\pi )^2`$.. Defining the radius $`R_\mathrm{H}`$ of the brane in the following way $$R_\mathrm{H}l\mathrm{cosh}\frac{y_0}{l},$$ (35) one can rewrite (29) as $$0=\frac{1}{\pi G}\left(\pm \frac{1}{R_\mathrm{H}}\sqrt{1+\frac{R_{\mathrm{H}}^{}{}_{}{}^{2}}{l^2}}\frac{1}{l}\right)R_{\mathrm{H}}^{}{}_{}{}^{4}+8b^{}.$$ (36) Hence, we showed that quantum, conformally invariant matter on the wall, leads to the inducing of inflationary 4d de Sitter-brane Universe realized within 5d Anti-de Sitter space (a la Randall-Sundrum). Of course, analytical continuation of our 4d sphere to Lorentzian signature is supposed what leads to ever expanding inflationary brane-world Universe. In 4d QFT (no higher dimensions) such idea of anomaly induced inflation has been suggested long ago in refs.. On the same time the inducing of 4d hyperbolic wall in brane Universe is highly suppressed and may be realized only for exotic conformal matter. The analysis of the role of domain wall CFT to metric fluctuations may be taken from results of ref.. It is interesting to note that our approach is quite general. In particulary, it is not difficult to take into account the quantum gravity effects (at least, on the domain wall). That can be done by using the corresponding analogs of central charge for various QGs which may be taken from beta-functions listed in book . In other words, there will be only QG contributions to coefficients $`b,b^{}`$ but no more changes in subsequent equations<sup>7</sup><sup>7</sup>7For example, for Einstein gravity $`b=611/120(4\pi )^2`$ and $`b^{}=1411/360(4\pi )^2`$.. The next question which deserves careful investigation is about the (in)stability of such anomaly driven inflation when it evolves to matter dominated Universe. This will be discussed elsewhere. Acknowledgements. The work by SDO has been supported in part by CONACyT (CP, ref.990356) and in part by RFBR. SDO would like to thank members of Instituto de Fisica de la UG for kind hospitality and Kim Milton for very helpful discussions.
warning/0004/hep-ex0004035.html
ar5iv
text
# A Introduction ## A Introduction The existence of three families of quarks and leptons suggests the possibility of a substructure for these objects . The hypothetical constituents known generically as preons, interact via a new strong interaction called Metacolor. The characteristic energy scale, $`\mathrm{\Lambda }`$, for the new interactions is, of course, unknown. The strength of the interactions through a contact term can be written as $`\widehat{s}/(\alpha _S\mathrm{\Lambda }^2)`$, where $`\widehat{s}`$ is the square of the energy in the center of mass frame of the (normal) interacting partons, and $`\alpha _S`$ is the QCD coupling. The first limit on the size of the atomic nucleus was obtained by Geiger and Mardsen in the Rutherford scattering of $`\alpha `$ particles from nuclei. In an analogous way, we can set a limit on the size of quarks and leptons by observing the scattering of the highest energy quarks and antiquarks at the Fermilab Tevatron at $`\overline{p}p`$ center-of-mass energies of 1.8 TeV for collider experiments, and 0.8 TeV for fixed-target experiments. The collider detectors at Fermilab, CDF and DØ, have performed searches for compositeness, and this paper gives a summary of those searches. Those detectors are general-purpose, have nearly $`4\pi `$ acceptance, and measure lepton and jet energies to high precision. In addition, the neutrino detector, CCFR, which utlilized the 800 GeV proton line at Fermilab has performed a compositeness search. ## B Contact Interactions: Indirect Searches for Compositeness Several previous searches for compositeness relied on direct searches for excited quark states . The searches that we discuss assume that $`\sqrt{\widehat{s}}`$ is small compared to the characteristic mass scale, $`\mathrm{\Lambda }`$. These are generically called “contact interaction” searches, and each interaction may be regarded as a contact term in the Lagrangian, which has the form $$\begin{array}{cc}\hfill L=\frac{2\pi }{\mathrm{\Lambda }^2}\{& \eta _{LL}(\overline{f}_L\gamma ^\mu f_L)(\overline{f}_L\gamma _\mu f_L)+\hfill \\ & \eta _{LR}(\overline{f}_L\gamma ^\mu f_L)(\overline{f}_R\gamma _\mu f_R)+\hfill \\ & 2\eta _{RL}(\overline{f}_R\gamma ^\mu f_R)(\overline{f}_L\gamma _\mu f_L)\},\hfill \end{array}$$ (1) where $`f_{L,R}`$ are the left and right-handed chiral components of the the quark or lepton, and $`\eta _{ij}`$ is the sign of each term, where $`\eta _{ij}=+1`$ corresponds to destructive, and $`\eta _{ij}=1`$ to constructive interference . Only leading-order calculations are available for compositeness. Hence, the following approximate relation is used to calculate the correction for higher-order contributions: $$\frac{\sigma (\mathrm{\Lambda }=X)_{LO}}{\sigma (\mathrm{\Lambda }=\mathrm{})_{LO}}\times \sigma (\mathrm{\Lambda }=\mathrm{})_{NLOorNLL},$$ (2) where $`\mathrm{\Lambda }=\mathrm{}`$, corresponds to the standard model. The next-to-leading order (NLO) or next-to-leading log (NLL) cross-sections are calculated using the event generators jetrad or the Monte Carlo written by Hamburg et al, respectively. The presense of any compositness is expected to contribute to an increase in cross section at large jet and lepton transverse momenta. Discussed in this paper are the most up-to-date quark-quark and quark-lepton contact interaction searches. Of course, excellent lepton-lepton searches have been accomplished as well . ## C Quark Compositenss ### 1 Limits from Dijet Mass Measurements DØ has measured the ratio of the dijet mass spectrum for jets at pseudorapidities $`|\eta ^{jet}|<0.5`$ relative to jets at $`0.5<|\eta ^{jet}|<1.0`$. The resulting cancellation of systematic uncertainties results in an absolute systematic uncertainty of $`<8\%`$ of the ratio. Predictions for quark compositeness are obtained using the LO pythia event generator . The ratios of these LO predictions with, and without compositeness, are used to scale the jetrad NLO predictions (see Eq. 2), and are compared to data in Fig.1. Limits on models with color singlet (octet), vector, or axial contact interactions were set using an analytic LO calculation rather than pythia. The resulting limits are given in Table I. Models of quark compositeness with a contact interaction scale 2.4 TeV for quark-quark interactions are excluded at the 95% CL. The limits on the scale of $`\mathrm{\Lambda }_{V_8}^{}`$ can be converted into limits on a flavor-universal coloron resulting in a 95% CL of $`M_C/cot\theta >837\mathrm{G}\mathrm{e}\mathrm{V}/c^2`$, where $`M_C`$ is the mass of the coloron and $`cot\theta `$ a parameter of the theory. The comparison is given in Fig. 2. ### 2 Limits from The Dijet Angular Distribution At small center of mass scattering angles, $`\theta ^{}`$, the dijet angular distribution predicted by leading order QCD is proportional to the Rutherford cross section: $`d\widehat{\sigma }_{\mathrm{ij}}/d\mathrm{cos}\theta ^{}1/\mathrm{sin}^4(\theta ^{}/2)`$. It is useful to measure the angular distribution in the variable $`\chi `$, rather than $`cos\theta ^{}`$, where $`\chi =(1+\mathrm{cos}\theta ^{})/(1\mathrm{cos}\theta ^{})=e^{|\eta _1\eta _2|}`$. The variable $`\chi `$ facilitates comparison of theory with data. The angular distribution from composite quarks in quark-quark interactions is isotropic. Hence, the measurement of the dijet angular distribution is sensitive to the presence of such new physics. The quantity measured in the dijet angular analysis is $`1/N(dN/d\chi )`$, which is measured in bins of dijet mass $`M_{\mathrm{JJ}}`$. Figure 3 depicts the angular distribution as measured by DØ for $`M_{\mathrm{JJ}}>625\mathrm{G}\mathrm{e}\mathrm{V}/c^2`$. To remove the point to point correlated errors, the distribution can be described by a single parameter, $`R_\chi `$=N($`\chi <X)/N(X<\chi <\chi _{\mathrm{max}}`$), which is the ratio of the number of events with $`\chi <X`$ to the number between $`X<\chi <\chi _{max}`$. The CDF analysis uses $`X=2.5`$, and DØ uses $`X=4`$. Figure 4 exhibits $`R_\chi `$ measured by CDF as a function of $`M`$ for two different renormalization scales, along with the predictions for different compositeness scales. The dijet angular distribution from QCD was calculated using the jetrad event generator. The predictions for quark compositeness scale are obtained using a LO analytic calculation. The ratio of these LO predictions with compositeness, to the LO with no compositeness, is used again to scale the jetrad NLO predictions. Analysis of the CDF data exludes models with $`\mathrm{\Lambda }_{LL}^+<`$ 2.1 TeV and $`\mathrm{\Lambda }_{LL}^{}<`$ 2.2 TeV. In addition, DØ uses their measurements to place limits on the production of colorons, requiring $`M_c/\mathrm{cot}\theta >759\mathrm{GeV}/c^2`$ (see Fig. 2). ### 3 Limits from Large $`H_T`$ DØ has also used the $`H_T=E_T^{jets}`$ variable to set limits on quark compositeness based on . By treating the event as a whole, this analysis complements the other searches for compositeness. It studies the compositeness of left-handed quarks in the left-left isoscalar contact term of the Lagrangian given in . The scale parameter for this term is $`\mathrm{\Lambda }_{\mathrm{𝐿𝐿}}`$. The events are chosen to have $`H_T>500\mathrm{GeV}`$, well above the contribution expected from top quarks that peaks near $`2m_t350\mathrm{GeV}`$. Jets with $`E_T>20\mathrm{GeV}`$ and $`|\eta ^{jet}|<3.0`$ are included in the calculation of $`H_T`$ for each event. In addition, the events are required to have at least one jet with $`E_T>115\mathrm{GeV}`$. Other cuts are applied to reduce the instrumental backgrounds, e.g. from mis-vertexing. The only important backgrounds considered are from such instrumental sources. Figure 5 displays the fractional deviation between the data and the Monte Carlo for the CTEQ4M PDF with renormalization scale $`\mu =E_T^{\mathrm{max}}/2`$. For $`\mathrm{\Lambda }_{\mathrm{𝐿𝐿}}`$ values between $`1.4`$ and $`7.0\mathrm{TeV}`$, pythia is used to simulate the effects of quark compositeness to leading order. This leading order calculation is scaled using Eq. 2, where jetrad is used to compute the NLO component. The results for composite quarks relative to expectations from the standard model are shown in Fig. 5 for $`\mathrm{\Lambda }_{\mathrm{𝐿𝐿}}=1.7,2.0`$ and $`2.5\mathrm{TeV}`$. The ratios are found to be independent of the pythia renormalization scale. As seen in Fig. 5, quark compositeness would produce a rise in the cross section as a function of $`H_T`$. Changes in renormalization scale affect the absolute cross section, but not the shape of the $`H_T`$ distribution. Cross sections calculated using CTEQ4M or MRST PDFs differ in normalization but only slightly in shape. The measured $`H_T`$ distribution above $`500\mathrm{GeV}`$ is well modeled by the jetrad (NLO QCD) event generator. No evidence is found for compositeness in quarks, and the data are used to set limits on compositeness. As seen in Table II, the limits are not greatly affected by choice of PDF. The depencence of the 95% CL limit is also studied as a function of renormalization scale, and shows little effect on the limits. The limits for $`\mathrm{\Lambda }_{\mathrm{𝐿𝐿}}^{}`$ are not given, but are nearly identical (slightly higher than on $`\mathrm{\Lambda }_{\mathrm{𝐿𝐿}}^+`$). ## D Quark-Lepton Compositeness ### 1 Neutrino-Nucleon Scattering The CCFR experiment used the Fermilab 800 GeVproton beam directed at a Beryllium-Oxide target to produce a beam of neutrinos (86% $`\nu _\mu `$, 12% $`\overline{\nu }_\mu `$, and 2% $`\nu _e`$($`\overline{\nu }_e`$)) with a mean energy of 125 GeV. The $`\nu `$’s are produced from decays of secondaries downstream from the BeO target. The CCFR detector was a $`3\times 3\times 18m`$ neutrino sampling calorimeter with the ability to resolve minimum-ionizing tracks (mostly from muons) and determine shower energy. The CCFR experiment ran from 1984 to 1988 and $`1\times 10^6\nu `$ scatterings pass typical analysis cuts. This detector has evolved to be what is now known as NuTeV, which used an experimental program of switching between $`\nu `$ and $`\overline{\nu }`$ beams to drastically reduce systematic errors and improve measurements of quantities such as $`sin^2\theta _W`$, and $`M_W`$ . The primary goal of these detectors was to do high-statistics studies of charged and neutral current interactions via $`\nu N`$ scattering. Most of these studies rely on measuring the ratio of the Neutral Current (NC) to Charged Current (CC) interactions seen in the detector. The NC interactions produce a compact shower of energy, while the CC interactions produce a shower, with a long ($`>2m`$) minimum-ionizing trail left by the $`\mu `$ or $`\overline{\mu }`$. An important correction to the ratio comes from the $`\nu _e`$($`\overline{\nu }_e`$) flux component, where the CC interaction produces $`e`$’s that, unlike $`\mu `$’s, cannot as easily be distinguished from the hadronic shower in a full range of energies. Such events can fake the NC signature. The $`R_{meas}=NC/CC`$ ratio is used to measure the coupling, $`\kappa `$, of the $`Z^0`$ boson to quarks, where the error in the comparison of $`R_{meas}`$ to theory is about 0.6%. With this measurement, CCFR has set limits on quark-neutrino contact interactions . Table III shows these limits. The NuTeV collaboration will soon provide complimentary and more precise information which can further constrain these contact interactions . ### 2 Drell-Yan Production Both CDF and DØ have placed limits on the combined quark-electron compositeness scale by analyzing Drell-Yan dilepton production. CDF has measured the Drell-Yan cross section for both electrons and muons, satisfying $`|\eta |<1.0`$ for each lepton (see Fig. 6). The backgrounds to the signal from QCD jets mimicking electrons, from $`W^+W^{}`$, $`\tau ^+\tau ^{}`$, and heavy quark production, are estimated using data, and subsequently subtracted. The Drell-Yan production cross section is based on an NLL calculation, with and the data normalized to the prediction in the mass range between 50 and 150 GeV/c<sup>2</sup>. The data are then used to place limits on the combined quark-electron compositeness (see Table IV). DØ has also measured the Drell-Yan cross section for electrons satisfying either $`|\eta |<1.1`$ or $`1.5<|\eta |<2.5`$ (Fig. 7). DØ measures the contribution to the cross section from misidentified QCD events, and other processes, using a combination of data and pythia Monte Carlo. The production cross section for Drell-Yan is calculated using an NNLO prediction. Limits are placed on various models of quark-electron compositeness (Table IV). DØ and CDF rule out models of quark-electron compositeness with interaction scales below 2.5 to 6.3 TeV, depending on the details of the model. ## E Conclusion The Fermilab experiments, CDF, DØ, and CCFR have shown that predictions from the Standard Model are in excellent agreement with the data, and there is no evidence for compositeness in quarks below a scales from 2-8 TeV. The next run of the Tevatron, beginning in early 2001, will reach far greater luminosities, and provide more opportunity for finding new physics at higher mass scales. ## F Acknowledgements The author would like to acknowledge assistance from Iain Bertram in the preparation of the talk and this proceeding.
warning/0004/hep-th0004057.html
ar5iv
text
# Cosmological Constant and Fermi-Bose Degeneracy ## A Introduction. Broken supersymmetry is hard to reconcile with vanishing of the cosmological constant. In the presence of gravity even unbroken supersymmetry alone does not suffice to guarantee the zero vacuum energy, but under reasonable assumptions, e.g. such as $`R`$ symmetry, this can be done. In this view perhaps the way for understanding the vanishing of the vacuum energy, is to look for the theories in which unbroken supersymmetry in vacuum cannot guarantee Fermi-Bose degeneracy among excited states. One may think of two possible strategies in this direction. One is to search for non-degenerate states directly in four-dimensional theories, generalizing earlier observation by Witten in $`2+1`$ dimensions. Alternatively , Fermi-Bose non-degeneracy may be compatible with unbroken bulk supersymmetry, if matter is localized on a brane embedded in infinite-volume extra dimensions. In this way cosmological constant may be controlled by bulk supersymmetry, while supersymmetry may be completely broken on the brane. Example of infinite-volume extra dimensions was recently suggested in . In this framework four dimensional Newtonian gravity is reproduced by a meta-stable graviton localized on the brane. In this letter we consider both possibilities. First we shall study some candidate solitonic states that make impossible the definition of unbroken supercharges in four-dimensions, although supersymmetry and $`R`$-symmetry are unbroken in vacuum. These objects are global monopoles that are known to create a solid angle deficit at infinity. Due to this solid angle deficit, the covariantly constant spinors do not exist on such a background and there are no conserved supercharges. Thus there is no reason for Fermi-Bose degeneracy in the monopole spectrum. In many respect the dynamics of the global monopoles is reminiscent of the one of the point-like particles with a finite mass localized in the tiny core. We show that this particles can carry gauge quantum numbers and in particular both Abelian or non-Abelian magnetic charges. In particular, with an appropriate choice of a “magnetic” gauge group $`H_{magnetic}`$, the magnetic charge assignment of the global monopoles can be made identical to the electric Yang-Mills charges of the standard model particles under $`H_{electric}=SU(3)SU(2)U(1)`$, much in the same way as for the gauge monopoles in $`SU(5)`$, as observed by Vachaspati . However, we shall not restrict $`H_{magnetic}`$ by such a choice. In this way we end up with the objects that in certain sense are “dual” to ordinary particles. Instead of global Noether charge (baryon or lepton charge), they carry global topological charge (winding number), instead of electric charges (color and ordinary electric charge), they carry $`H_{magnetic}`$-magnetic charge, etc. An interesting fact about these states is that despite the unbroken supersymmetry in the vacuum, they are not Fermi-Bose degenerate. So the theory describes a toy Universe with zero cosmological term, but non-equal masses of fermions and bosons. The hope for more realistic model-building is that in some dual description solitons may be replaced by particles (topological charges with baryon or lepton numbers, etc..) which are not Fermi-Bose degenerated, but cosmological constant is still zero due to supersymmetry. Secondly we consider infinite volume theories and suggest some new ways of quasi-localization of gravity without invoking negative tension branes. One possibility is that our brane originates from spontaneous breaking of global $`O(3)`$ symmetry in theories with three extra dimensions and has the structure of global monopole in transverse space. Such an object produces a deficit angle at infinity and in a certain limit may support a meta-stable graviton state. Since the brane appears as a solution of an underlying sigma model on a background with zero bulk cosmological constant, it may avoid problems with violation of the weak energy positivity conditions pointed in (see also ). ## B Supersymmetry in the Global Monopole Background in Four Dimensions. In the inspiring paper Witten made an observation that unbroken supersymmetry (and thus zero cosmological constant) is not incompatible with Fermi Bose non-degeneracy in $`2+1`$ dimensions. The idea is roughly as follows. In supergravity theories the supercurrent in general is not conserved in the usual sense, but rather is covariantly conserved $$D_\mu J^\mu =0$$ (1) However, in the presence of a covariantly constant spinor $$D_\mu ϵ=0$$ (2) the conserved current can be constructed $$\overline{ϵ}J^\mu $$ (3) and thus one can define a globally conserved supercharge $$Q=𝑑x^3\overline{ϵ}J^0$$ (4) Now in three dimensions any localized mass is known to produce a conical geometry at infinity. Such a geometry makes in general impossible the existence of covariantly constant spinors and thus of unbroken supercharges.In the presence of gauge fields, Killing spinors may still exist, since the deficit angle can be compensated by an Aharonov-Bohm phase.. Thus although supersymmetry is unbroken in the vacuum and vacuum energy vanishes, there is no Fermi-Bose degeneracy among the excited states. Explicit examples along Witten’s idea were considered in some papers ,. Needless to say it would be very important to find some sort of generalization of this effect to four-dimensions. In this letter we will consider some possible candidates that may generalize the three-dimensional behavior of point masses to four dimensions. We shall look for $`N=1`$ $`D=4`$ supergravity theories, in which both supersymmetry and $`R`$ symmetry are unbroken in the vacuum, but yet there is no Fermi-Bose degeneracy among certain excited solitonic states. The model consists of four chiral superfields. Three of them, $`\mathrm{\Phi }_a,a=1,2,3`$ compose a triplet under an internal symmetry group $`O(3)`$, while the fourth superfield $`X`$ is a singlet. The superpotential is given by $$W=X(\mathrm{\Phi }_a^2v^2)$$ (5) where $`v`$ is a real mass parameter and for simplicity the coupling constant is set to one. This theory has a global $`O(3)`$ invariance, plus an $`U(1)_R`$ R-symmetry, under which $`X`$ transforms in the same way as $`W`$, whereas $`\mathrm{\Phi }_a`$-s are invariant. The vacuum of the theory is given by (below we shall denote chiral superfields and their scalar components by the same symbols) $$X=0,\mathrm{\Phi }_a^2=v^2$$ (6) Thus $`O(3)`$ is broken spontaneously to $`O(2)`$, whereas both supersymmetry and $`R`$ symmetry are unbroken. This ensures that vacuum energy is zero to all orders in perturbation theory. Due to a nontrivial topological structure, however, this theory admits topological knots. These knots are global monopoles, which in spherical coordinates can be given by the solution $$X=0,\mathrm{\Phi }_a=f(r)\frac{r_a}{r}$$ (7) where $`r_a`$ are the components of the radius-vector $`𝐫`$ and $`f(r)`$ is a smooth function such that $$f(0)=0,f(r)|_r\mathrm{}v$$ (8) As shown by Barriola and Vilenkin , with this ansatz, the metric takes the following form $$ds^2=a^2dt^2+b^2dr^2+r^2d\mathrm{\Omega }^2$$ (9) where asymptotically $$a^2=b^2=\left(1\frac{v^2}{M_P^2}\frac{2M_{core}}{M_P^2r}\right)$$ (10) where $`M_P`$ is the reduced Planck mass and $`M_{core}v`$ is an integration constant, which is a negative quantity. Due to this fact the Newtonian potential of the monopole core is repulsive. This metric describes a space with a solid angle deficit $`4\pi \frac{v^2}{M_P^2}`$. The $`\theta =\pi /2`$ surface is a cone with a deficit angle $`\delta =2\pi v^2/M_p^2`$. Of our interests are the supersymmetric transformations in the monopole background. Choosing the vierbein as $$e_a^\mu =diag(\frac{1}{a},a,\frac{1}{r},\frac{1}{r\mathrm{sin}\theta }),a^2=\left(1\frac{v^2}{M_P^2}\right)$$ (11) (where $`\mu =t,r,\theta ,\varphi `$) the only non-vanishing components of the spin connection are $`\mathrm{\Gamma }_\theta =\frac{a}{2}\gamma _1\gamma _2`$ and $`\mathrm{\Gamma }_\varphi =\frac{a}{2}\mathrm{sin}\theta \gamma _1\gamma _3\frac{1}{2}\mathrm{cos}\theta \gamma _2\gamma _3`$. On such a background, the conditions for Killing spinors $`\delta \psi _\theta =(_\theta +{\displaystyle \frac{a}{2}}\gamma _1\gamma _2)ϵ=0`$ (12) $`\delta \psi _\varphi =(_\varphi +{\displaystyle \frac{a}{2}}\mathrm{sin}\theta \gamma _1\gamma _3+{\displaystyle \frac{1}{2}}\mathrm{cos}\theta \gamma _2\gamma _3)ϵ=0`$ (13) can not be satisfied and no conserved supercharges can be defined to control Fermi-Bose degeneracy in the monopole spectrum. Alternatively, this can be seen in the language of zero modes. In the global supersymmetry limit the zero mode in the monopole background is $`\delta \psi _{X+}=\sqrt{2}(f^2(r)v^2)ϵ_+`$ (14) $`\delta \psi _+^a=i\sqrt{2}_\mu \mathrm{\Phi }^a\gamma ^\mu ϵ_{},`$ (15) where $`ϵ_\pm `$ is an eigenspinor of $`\gamma _5`$. Note that $`\delta \psi _{X+}`$ and $`\frac{r_a}{|r|}\delta \psi _{a+}`$ are normalizable, whereas $`\delta \psi _\theta `$ and $`\delta \psi _\varphi `$ are not. The reason behind this is the following. We can say that global supersymmetry is spontaneously broken in the monopole background. The order parameters for this breaking are the auxiliary ($`F`$) component of the chiral superfield $`X`$ $$F_X=(f(r)^2v^2)$$ (16) and also an auxiliary ($`D`$) component of a composite real superfield $`\mathrm{\Phi }_a^{}\mathrm{\Phi }_a`$ $$(\mathrm{\Phi }_a^{}\mathrm{\Phi }_a)_D=|_j\mathrm{\Phi }_a|^2$$ (17) Since both $`F_X`$ as well as radial derivatives vanish away from the core at least as $`1/r^2`$, the corresponding fermion variations $`\delta \psi _X`$ and $`\frac{r_a}{|r|}\delta \psi _a`$ are normalizable. Away from the core breaking is dominated by angular gradients, which only drop-off as inverse square of distance and goldstino is not normalizable. In fact this is also to be expected from the fact that monopole configuration spontaneously breaks both internal symmetry as well as coordinate rotations and leaves unbroken an $`O(3)`$ symmetry of combined space and internal rotations, which leaves invariant a product $`r_a\mathrm{\Phi }_a`$. Non-normalizable fermionic zero modes are simply the supersymmetric partners of the Nambu-Goldstone fields corresponding to this breaking. Monopole configuration interpolates between the core where internal $`O(3)`$ is restored, but coordinate symmetry and supersymmetry are broken maximally, to the outer region where the strength of supersymmetry and coordinate symmetry breakings weakens but internal symmetry is maximally broken. Since both coordinate and supersymmetries are broken by angular gradients which vanish away from the core, all energy densities must vanish as $`1/r^2`$. Although locally energy density can be arbitrarily small, in the background of the global monopole the globally conserved supercharge linearly diverges together with the total energy of the configuration $$Q(ϵ)4\pi \overline{ϵ}\gamma _0ϵv^2R$$ (18) where $`ϵ`$ is an arbitrary constant spinor and $`R`$ is the distance from the core of the monopole. In the case of supergravity, the zero mode acquires an additional non-normalizable gravitino component $`\delta \psi _\mu `$ which renders it unphysical. ## C Monopoles as Particles. The global monopoles due to the divergent energy, are different from the ordinary localized sources. However, in many respect they can behave like objects with a finite mass localized in the core. We can treat the cores of the global monopoles as point-like objects of finite mass moving in the “clouds” of the Goldstone gradient energy produced by their Goldstone field. In this way divergence in the total energy of the configuration is not important for the core dynamics, in a same way as the total infinite energy of the matter in the Universe is not important for the particle interactions at short distances. Of course, the precise numbers do not work as well as in the case of the ordinary matter. For instance, roughly $`n10^{19}`$ global monopoles, with the core mass equal to the proton mass, presented within the observable part of the Universe would be enough to produce an energy density equal to the critical This estimate assumes that all the monopoles within the observable part have the same charge (like baryons). If one assumes the equal number of monopoles and anti-monopoles within the Hubble horizon, the allowed number would be much higher.. However, we shall ignore this complications, since our goal is to understand whether in principle $`4D`$ theories with unbroken vacuum supersymmetry and Fermi-Bose non-degeneracy are possible. We see no reason to think that the “dual” Universe in which particles are described by monopoles must obey similar cosmological constraints. To understand how far the analogy between particles and the global monopoles can be extended, we can estimate their interaction potential due to various forces. We shall discuss gravitational effects first. The leading contribution comes from a tree-level one-graviton exchange $$G𝑑x^4d^4x^{}T_{\mu \nu }(x)G^{\mu \nu \alpha \beta }(xx^{})T_{\alpha \beta }^{}(x^{}),$$ (19) where $`G`$ is a Newtons constant and the graviton propagator is given by $$G^{\mu \nu \alpha \beta }(xx^{})=\frac{dp^4}{(2\pi )^4}\frac{\frac{1}{2}(\eta ^{\mu \alpha }\eta ^{\nu \beta }+\eta ^{\mu \beta }\eta ^{\mu \alpha })\frac{1}{2}\eta ^{\mu \nu }\eta ^{\alpha \beta }}{p^2iϵ}e^{ip(xx^{})}$$ (20) For the rough estimate we shall ignore the effect of the curvature (which dies-off as $`1/r^2`$ away from the core) and use the flat space propagator. In the first approximation, we can assume that the structure of gravitating sources are not affected by simultaneous presence of two monopoles, and ignoring the effect of the core we can set: $$T_t^t=T_r^r=\frac{v^2}{|𝐫|},T_t^{}_{}{}^{}t=T_r^{}_{}{}^{}r=\frac{v^2}{|𝐫^{}𝐑|}$$ (21) and all other components zero. Here $`R`$ is a distance between the monopoles. In this case (19) is zero. So the gravitational interaction of monopoles will be governed by the Newtonian interactions between the two cores<sup>§</sup><sup>§</sup>§We shell ignore the tidal acceleration $`1/r^2`$, which may be very important for non-static sources moving with a nonzero impact parameter. $$V(r)G\frac{M_{core}^2}{r}$$ (22) This is sub-dominant with respect to the force mediated by the Goldstone field. For a monopole-anti-monopole pair the Goldstone-mediated interaction gives an attractive constant force $$V(r)v^2r$$ (23) However, we are more interested in monopoles of the same charge, since a monopole-anti-monopole pair produces no deficit angle at infinity. Below we shall assume that vacuum topology is such that monopoles with the higher winding numbers are possible. In this case the long range interaction is repulsive (there still however can be an attractive short-range interaction between the cores). Due to this the $`n`$-monopole system is unstable. This fact can be understood qualitatively as follows. Consider $`n`$ monopoles uniformly distributed inside the sphere of radius $`R`$. For distances $`r>>R`$ the configuration of the Higgs field is similar to the one of the monopole with winding number $`n`$, and gradient energy diverges as $$E_{out}n^2v^2(r_{max}R)$$ (24) while inside the sphere it scales as $$E_{in}nv^2R$$ (25) Therefore it is energetically favorable for $`R`$ to grow. However, the linear potential between monopoles can be modified in many cases. In particular, it is easy to imagine situation when this force vanishes because of the complex structure of the monopole. For instance, imagine that there are two broken global symmetries $`G`$ and $`G^{}`$ that produce monopoles. Let $`\mathrm{\Phi }`$ and $`\mathrm{\Phi }^{}`$ be the Higgs fields responsible for such breakings. We assume for simplicity that $`\mathrm{\Phi }`$ is trivial under $`G^{}`$ and so is $`\mathrm{\Phi }^{}`$ under $`G`$ respectively. Then the two fields produce independent monopoles which we shall refer to as $`\mathrm{\Phi }`$-monopoles and $`\mathrm{\Phi }^{}`$-monopoles respectively. If there is a $`GG^{}`$-invariant contact interaction in the potential $$\mathrm{\Phi }^{}\mathrm{\Phi }\mathrm{\Phi }^{{}_{}{}^{}}\mathrm{\Phi }^{}$$ (26) the monopoles of the two sorts will tend to create the bound-states. Such interaction can be easily arranged by introducing the following couplings in the superpotential $$W=Y(h\mathrm{\Phi }^2h^{}\mathrm{\Phi }^{}_{}{}^{}2)+\mathrm{}$$ (27) Where $`Y`$ is a chiral superfield and $`h,h^{}`$ are constants The equation of motion for $`Y`$ then forces $`h\mathrm{\Phi }^2=h^{}\mathrm{\Phi }^{}_{}{}^{}2`$ and encourages $`\mathrm{\Phi }`$ and $`\mathrm{\Phi }^{}`$ to have coincident zeros. Thus $`\mathrm{\Phi }`$-monopoles get bounded to $`\mathrm{\Phi }^{}`$-monopoles. The strength of the binding force is set by the parameters $`h,h^{}`$ and the magnitude of the two VEVs. Note that the binding interaction does not distinguish between monopoles or anti-monopoles in the opposite sectors. So formation of a monopole-monopole ($`MM`$) boundstate is as probable as the formation of a monopole-anti-monopole ($`M\overline{M}`$) one. Depending on the parameters of the theory, the Goldstone-mediated force between $`MM`$ and $`M\overline{M}`$ states may be either repulsive, attractive or zero. In the latter case the interaction will be dominated by a gravitational potential from the core. In addition global monopoles can have long-range gauge interactions. In particular, as we shall show below, this long range interaction can be due to their magnetic charge. ## D Global Monopoles with Magnetic Charges. Interestingly the global monopoles can carry magnetic charges under both Abelian or non-Abelian gauge groups. Let us consider first the Abelian case. For this, in addition to the spontaneously broken $`O(3)`$ global symmetry we introduce an unbroken $`U(1)_m`$ gauge symmetry in the theory. The question is whether a global monopole given by (7) can acquire a $`U(1)_m`$-magnetic charge. To answer this let us define an invariant two form $$M_{\mu \nu }=ϵ_{abc}\mathrm{\Phi }^a_\mu \mathrm{\Phi }^b_\nu \mathrm{\Phi }^c$$ (28) which determines a topological winding number $$n=\frac{1}{4\pi }_{S_2}\frac{M_{\mu \nu }}{|\mathrm{\Phi }|^3}𝑑x^\mu 𝑑x^\nu $$ (29) where integral is taken over a two-sphere surrounding the monopole. Let us now couple this two-form to a gauge-invariant $`U(1)_m`$-field strength $$\frac{\lambda }{2}M_{\mu \nu }F^{\mu \nu }$$ (30) We shell treat this term as the perturbation on the stable monopole background. The equation of motion then gives $$_\mu F^{\mu \nu }=\lambda _\mu M^{\mu \nu }$$ (31) This tells us that $`U(1)_m`$-magnetic field $`(𝐁)`$ is radial, but does not fix its magnitude. The magnitude is determined by minimizing the $`U(1)_m`$-magnetic energy of the system $$E_{magnetic}=\lambda \mathrm{𝐌𝐁}+\frac{𝐁^2}{2}$$ (32) where $`𝐌_𝐢=ϵ_{ijk}M_{jk}`$. The energy is minimized by $$𝐁=\lambda v^3\frac{𝐫}{r^3}$$ (33) In this way, the global monopole acquires $`U(1)_m`$ magnetic charge $$Q_m=\frac{1}{4\pi }_{S_2}F_{\mu \nu }𝑑x^\mu 𝑑x^\nu =Nv^3\lambda $$ (34) proportional to the topological charge of its own. For this configuration not to be singular $`U(1)_m`$ must be embedded in some non-Abelian group. To do this, let us, instead of $`U(1)_m`$, introduce $`SU(N)`$ gauge symmetry, spontaneously broken to $`SU(NM)SU(M)U(1)`$ by a Higgs field $`\mathrm{\Sigma }`$ in the adjoint representation. We can define a gauge-invariant two-form $$F_{\mu \nu }=\mathrm{Tr}\mathrm{\Sigma }𝐅_{\mu \nu }$$ (35) where $`𝐅_{\mu \nu }`$ is the $`SU(N)`$ field-strength. Then the coupling of the form (30) between (35) and (28) will induce a magnetic charge of the global monopole. Effectively what happens is that $`SU(N)`$-gauge t’Hooft-Polyakov monopole gets trapped in the core of the global monopole. Finally in case when gauge sector permits topologically stable monopoles (just like in above $`SU(N)`$ example), the global monopole can acquire magnetic charge by simply trapping the gauge monopoles inside its core due to a short-range Higgs interaction of the form similar to (27) $$W_{cross}=Y(h\mathrm{\Phi }^a\mathrm{\Phi }^ah^{}Tr\mathrm{\Sigma }^2)$$ (36) This interaction tries to bring zeros of the two Higgs fields together and thus creates a boundstate of global and gauge monopoles. For instance, we can choose $`SU(2)`$ gauge symmetry broken by triplet $`\mathrm{\Sigma }^a`$ and set its self-interaction term in the superpotential $$W_\mathrm{\Sigma }=Z(Tr\mathrm{\Sigma }^2v_\mathrm{\Sigma }^2)$$ (37) where $`Z`$ is an additional superfield. The asymptotic form of the solutions for the “composite” monopole is $$Y=Z=0,\mathrm{\Phi }^a=\pm \sqrt{\frac{h^{}}{h}}\mathrm{\Sigma }^a=v_\mathrm{\Sigma }\frac{r^a}{r},A_\mu ^a=ϵ_{\mu ab}\frac{r^b}{r^2}$$ (38) and has the same magnetic charge as the elementary $`SU(2)`$ monopole. ## E Some speculations. As we have seen, the global monopoles in $`N=1`$ $`D=4`$ supersymmetric theories exhibit some interesting properties. Due to the solid angle deficit at infinity they do not permit existence of conserved supercharges and, thus, there is no a priory reason for a Fermi-Bose degeneracy in the spectrum, although supersymmetry is unbroken and cosmological constant vanishes. Despite the fact that the total energy of the configuration diverges, in some respect they behave like point-like particles of a finite mass localized at the core. These localized masses can carry gauge quantum numbers and, in particular, both Abelian as well as non-Abelian magnetic charges. If this toy picture can have any relation with the observed smallness of the cosmological term, we should expect that in some “dual” theory these monopoles are related to the ordinary particles carrying electric Yang-Mills charges. It has been shown that the $`SU(3)SU(2)U(1)`$ magnetic charges of the gauge monopoles produced in the symmetry breaking $$SU(5)SU(3)SU(2)U(1)/Z_6$$ (39) are in correspondence with the electric gauge charges of the standard model fermions. This interesting observation naturally leads to the idea of the dual standard model. We have shown that global monopoles can carry any magnetic charge that can be carried by the local ones. Then the following toy scenario emerges. We take $`N=1`$ supersymmetric theory with a symmetry group $`G_gG_l`$, where $`G_g`$ is global and $`G_l`$ is gauged. We assume that these symmetries are spontaneously broken to $`H_gH_l`$ by Higgs fields $`\mathrm{\Phi }`$ and $`\mathrm{\Sigma }`$ respectively. The breaking is such that $`\pi _2(G_g/H_g)0`$. This ensures the existence of stable global $`\mathrm{\Phi }`$-monopoles. On the other hand $`\pi _2(G_l/H_l)`$ may or may not be empty, since as we have seen, the global monopoles can acquire magnetic charges even if the vacuum of the broken gauge symmetry is topologically trivial. If however, also $`\pi _2(G_l/H_l)0`$, then there will be local monopoles formed by $`\mathrm{\Sigma }`$. If $`G_l`$ is simply connected, then gauge monopoles are classified by non-contractable paths in $`H_l`$ and carry different magnetic charges. There can be a finer classification according to the representations of a dual magnetic group. Now the global monopoles can acquire magnetic charges under $`H_l`$ via one of the above discussed mechanisms, either by “kinetic” mixing (35) or by forming the boundstate due to the contact Higgs interaction (27). In the latter case for each stable gauge monopole there will be a composite stable global monopole with identical gauge charge. These objects carry a global topological charge (“dual baryon number”) and a local magnetic charge (“dual electric charge”). If $`H_l=SU(3)SU(2)U(1)/Z_6`$ then according to the magnetic charges of these monopoles will match the $`SU(3)SU(2)U(1)`$-electric charges of the standard model fermions. This choice however is not necessary since the dual “electric” group $`\stackrel{~}{H_l}`$ need not be isomorphic to $`H_l`$. They move in a cloud of Goldstone field which creates a solid angle deficit at infinity. No conserved supercharges can be defined on such a background. On the other hand supersymmetry and $`R`$ symmetry are unbroken in the vacuum and cosmological constant vanishes. Then one may expect that in the dual picture this monopoles are replaced by electrically charged particles (“baryons” and “leptons”) of some more conventional theory in which fermions and bosons will not be degenerate but vacuum energy will be zero because of (hidden) unbroken supersymmetry. Such a duality probably has to exchange the global topological winding number with the global Noether charge (baryon or lepton numbers), ## F Infinite-Volume Extra Dimensions. An alternative way of controlling the cosmological constant by supersymmetry may be by going to a “brane world” scenario. In the standard brane world picture the ordinary matter is localized on a brane embedded in $`N`$ large extra dimensions. These can be as large as millimeter with the fundamental Planck mass $`(M_{Pf})`$ around TeV. The volume of extra dimensions is finite, either because these dimensions are compact or because the warp factor decays exponentially fast away from the brane. The compactification combined with non-trivial warp factors in co-dimension two spaces were considered. Compactification may take place due to singularity at finite distance from the brane as, for instance, in . In either case there is an upper bound around $``$ mm on the volume of extra dimensions from gravitational measurements. We shall define this volume as $$V_{extra}=𝑑x_{extra}^N\sqrt{g_{extra}}g^{00}(x_{extra})$$ (40) where the integration is taken over an $`x_{extra}`$-dependent part of the metric. This volume factor sets the normalization of the 4D zero mode graviton and thus the relation between fundamental and observable Planck scales $$M_P^2=M_{Pf}^{2+N}V_{extra}$$ (41) In this set-up, it is very difficult to control the value of the cosmological constant by bulk supersymmetry. Indeed, since the extra volume is finite, the extra coordinates can be integrated over and at large distances, the effective four-dimensional description should be valid. Then by four-dimensional general covariance the brane and bulk states should be gravitationally connected through the zero mode graviton. Now, supersymmetry must be broken among the brane states at least at scale $`TeV`$. This breaking then will universally get transmitted to all bulk modes by gravity. This transmission is at best volume suppressed, so that by dimensional argument the mass splitting among the bulk states is at least $`\mathrm{\Delta }m^2TeV^4/M_P^2`$ and the resulting cosmological constant is at least $`\mathrm{\Delta }m^2\mathrm{\Lambda }_{cutoff}^2`$. The crucial point is that in theories with infinite $`V_{extra}`$ there is no a priory reason for the above inconsistency : since the volume is infinite, the theory is never four-dimensional and effective 4D description is never valid. In other words, supersymmetry broken on the brane may not get transmitted in the bulk. So the high-dimensional theory can be supersymmetric even though brane observers will not see any Fermi-Bose degeneracy. An interesting model of infinite-volume extra dimensions was invented in as modification of RS scenario. (Modification of gravity at large distances was suggested earlier in .) In their case the warp factor in five dimensional metric $$ds^2=A(y)ds_4^2dy^2$$ (42) instead of vanishing as $`y\pm \mathrm{}`$, was asymptotically approaching a tiny constant value. As a result the space is infinite and asymptotically flat in $`y`$ direction. This structure was achieved by expense of introducing a combination of positive and negative tension branes. Despite this fact a correct four-dimensional Newtons law is reproduced at intermediate scales on the brane. As was shown in , this can be explained by the existence of meta-stable resonant graviton localized on the brane. There are two potential problems with this scenario. One, to be put aside in the present paper, is the fact that 4D gravity on the brane is mimicked by massive spin-2 fields, which can not reproduce the predictions of Einsteins theory . This is true whenever dominant contribution to 4D gravity comes from a massive spin-2 state(s). Some way out of this problem in the context of was suggested in <sup>\**</sup><sup>\**</sup>\**In it was noted that the unwanted contributions to one-graviton exchanges may be canceled by unconventional states like ghost (we understand that this was essentially confirmed by ref), but then it is hard to make sense of the theory, even if this states only dominate exchanges at finite distances. This goes in contrast to the conclusions of and in which it was argued that such states can solve the problem (conclusion about their long-distance behavior is opposite in these two references). We think that existence of the physical ghost is a problem at any distances. For instance, even if canceled in all one-particle exchanges, it is not clear what can prevent their appearance in the final state. Second is the violation of the weak energy condition, which has to do with the existence of AdS portion bounded by negative tension branes embedded in Minkowski space. In this respect it is very important to obtain a source that quasi-localizes gravity as a solution of the underlying theory. We want to note that the branes which appear from spontaneous breaking of non-Abelian global symmetry may have this property. We consider co-dimension $`3`$ case. In this case the brane is a global monopole embedded in three extra dimensions<sup>††</sup><sup>††</sup>†† From a different perspective the higher co-dimension global defects were studied by Vilenkin and Olasagasti and by Cho. I am grateful to these authors for correspondence and discussions. Consider a theory in $`7`$ dimensions with a spontaneously broken global $`O(3)`$ symmetry, by the VEV of a triplet scalar field $`\mathrm{\Phi }_a`$. The potential is $$V=(\mathrm{\Phi }_a\mathrm{\Phi }_av^2)^2$$ (43) Due to topological arguments this theory has co-dimension $`3`$ objects, independent of four space-time coordinates, described by eq(7), where $`r`$ has to be understood as the radial coordinate perpendicular to the brane. The general spherically symmetric ansatz for the metric can be written $$ds^2=a^2(r)\eta _{\mu \nu }dx^\mu dx^\nu b^2(r)dr^2r^2(d\theta ^2+sin^2\theta d\varphi ^2)$$ (44) Much in the same way as for the global monopole, away from the core we can set $`f(r)=v^2`$ and the only contribution to the source will come from the angular derivatives. As a result $$T_\mu ^\mu =T_r^r=\frac{v^2}{r^2}+\mathrm{}$$ (45) and all other components zero. The ellipses stand for the sub-leading correction. With this source, there is a straightforward generalization of the Barriola-Vilenkin solution, which (in $`M_{fp}`$ units) reads: $$a^2=b^2=1v^2=\alpha $$ (46) As in the four-dimensional case, there is a deficit angle $`4\pi v^2`$. In principle, by tuning the parameter $`v`$, one can make angular deficit arbitrarily close to $`4\pi `$. Now let us study the issue of graviton localization. First let us ignore the subleading corrections in (45). That means assume that $`a`$ goes to it asymptotic constant value fast enough. Then the volume of the transverse space $$V_{extra}=4\pi 𝑑ra^2br^2$$ (47) is infinite unless $`a`$ vanishes asymptotically. This volume has two contributions $$V_{extra}=V_{core}+4\pi _{r_{core}}^{\mathrm{}}ar^2𝑑r$$ (48) The second term diverges, unless we tune $`\alpha =0`$ in which case the volume is finite and is given by $`V_{core}`$. There is no reason to expect in $`\alpha 0`$ limit any pathological behavior in the core. Assuming this the theory must have a finite volume and must support a localized zero mode graviton in the core of the monopole. The following fluctuation about the background metric $$ds^2=a^2(r)(\eta _{\mu \nu }+h_{\mu \nu })dx^\mu dx^\nu b^2(r)dr^2r^2(d\theta ^2+sin^2\theta d\varphi ^2)$$ (49) is a zero mode. This mode is normalizable if $`V_{extra}`$ is finite. Now, we can make $`\alpha `$ to be nonzero, but arbitrarily small. Now the volume becomes infinite and the zero mode becomes non-normalizable. However, if $`\alpha `$ is small enough, the system by continuity should still allow a localized meta-stable mode, with an arbitrarily long life-time. A problem with this argument is that it relies on an oversimplified anzats for the monopole core. For a realistic monopole, there are sub-leading corrections $`\frac{1}{r^2}`$ to eq(45) (recall that $`(f^2v^2)`$ dies away as $`\frac{1}{r^2}`$). So the volume for $`r\mathrm{}`$ is divergent, which makes existence of quasi-localized mode unclear. More precise conclusion depends on the model of monopole core which will not be discussed here. Summarizing, in order to avoid the violation of the weak energy condition one may try to quasi-localize gravity on a brane which is a solution of an underlying sigma model. Under certain assumption regarding the core structure, this brane may supports a meta-stable graviton because of the deficit angle at infinity. The next step would be to supersymmetrize the initial theory in such a way that supersymmetry gets restored in the bulk. However, this is not so easy, since due to the same deficit angle we will not be able to define globally conserved supercharges in such a background. However, this is not necessarily a problem, since as in four-dimensions, the bulk vacuum energy density is controlled by the order parameters that break supersymmetry. These are angular gradient densities that die away from the core as $`1/r^2`$. Thus all the energy densities in the bulk must be controlled by the distance from the brane and get zero at infinity. Although the total transverse energy is linearly divergent, this is not of any problem, since the four-dimensional metric on the brane is flat. An alternative approach one can try is to reintroduce Killing spinors, by adding gauge fields and canceling the deficit angle by Aharonov-Bohm-type phases, just as in the three dimensional case of ref. In this way some of the supercharges may be unbroken, but act trivially on the brane. ## G conclusions. It is not clear how the $`4D`$ scenario discussed in this paper can be related to an observable Universe. Yet it has some ingredients for this connection and is an example of $`4D`$ supersymmetric theory with vanishing cosmological constant, in the background that makes impossible definition of unbroken supercharges. Probably the way this framework may be related to more conventional ones is through some sort of a duality, which relates solitons with particles, magnetic charges with electric charges and in addition global topological charges (winding number) to conserved global Noether charges (such as baryon or lepton number). Among many open questions, not addressed in this letter, there are cosmological issues. For instance with a rough estimate the number of the global monopoles that would contribute $`\rho _{baryon}`$ energy density within the observable part of the Universe is much less than the number of baryons. So one may ask how this fast fits in the above picture. We do not see any reason why two theories related through duality should have similar cosmological history. We expect that duality relates spectra and quantum numbers, but not the actual number of occupied states, which is determined by the cosmological initial conditions. Alternatively infinite volume extra dimensional theories may provide example of unbroken supersymmetry compatible with four-dimensional Fermi-Bose non-degeneracy. We have suggested that gravitating sources similar to global topological defect, producing angular deficit in the extra space, may under certain conditions localize $`4D`$ meta-stable graviton. Acknowledgments It is pleasure to thank B. Bajc, G. Gabadadze, E. Poppitz, M. Porrati, M. Shifman, T. Vachaspati and A. Vilenkin for very useful discussions. This work is supported in part by a David and Lucile Packard Foundation Fellowship for Science and Engineering.
warning/0004/math0004179.html
ar5iv
text
# Log mirror symmetry and local mirror symmetry ## 1. Log mirror symmetry ### 1.1. A-model On the enumerative side, we counted curves of the lowest ‘log genus’ in \[T1\]. ###### Definition 1.1. A plane curve $`C`$ is said to satisfy the condition (AL) if it is irreducible and reduced and the normalization of $`C\backslash B`$ is isomorphic to the affine line $`𝔸^1`$. ###### Remark 1.2. If $`C`$ satisfies (AL), then $`CB`$ consists of one point. This point is a $`(3\mathrm{deg}C)`$-torsion for a group structure on $`B`$ whose zero element is an inflection point of $`B`$. We treat only ‘primitive’ cases. ###### Definition 1.3. Let $`P`$ be a point of *order* $`3d`$ for a group structure on $`B`$ whose zero element is a point of inflection. Then we define $`m_d`$ to be the number of curves $`C`$ of degree $`d`$ which satisfy (AL) and $`CB=\{P\}`$ (or, more precisely, the degree of the 0-dimensional scheme parametrizing such curves). The above definition implicitly assumes that the number $`m_d`$ does not depend on the choice of $`P`$. Although we haven’t proved it yet, it holds in the cases where we know $`m_d`$, i.e. when $`d8`$. ###### Theorem 1.4. (\[T1\]) We have the following table for $`m_d`$: | $`d`$ | $`1`$ | $`2`$ | $`3`$ | $`4`$ | $`5`$ | $`6`$ | | --- | --- | --- | --- | --- | --- | --- | | $`m_d`$ | $`1`$ | $`1`$ | $`3`$ | $`16`$ | $`113`$ | $`948`$ | Furthermore, under a technical hypothesis(see \[T1\]), we have $`m_7=8974`$ and $`m_8=92840`$. Our invariants are apparently related to the relative Gromov-Witten invariants defined in \[IP\] and \[LR\] for pairs of a symplectic variety and its subvariety of real codimension 2. In algebraic language: ###### Proposition-Definition 1.5. (1-pointed case of \[G\]) Let $`\overline{M}_{0,1}(^2,d)`$ be the moduli stack of $`1`$-pointed genus 0 stable maps of degree $`d`$ to $`^2`$ and $`\overline{M}_i^B(^2,d)`$ the closed subset consisting of points corresponding to $`f:(C,P)^2`$ such that (i) $`f(P)B`$ and (ii) $`f^{}BiPA_0(f^1B)`$ is effective. Then the virtual fundamental class $`[\overline{M}_i^B(^2,d)]^{virt}`$ is naturally defined and is of expected dimension $`3di`$. Considering that the number of $`3d`$-torsions is $`(3d)^2`$ and that the relative Gromov-Witten invariants take multiple covers into account, we expect the following to hold. ###### Conjecture 1.6. $`[M_{3d}^B(^2,d)]^{virt}=(3d)^2_{k|d}(1)^{dd/k}m_{d/k}/k^4`$. ###### Remark 1.7. (1) Although the factor $`k^4`$ looks unfamiliar, this is compatible with the factor $`k^3`$ for Gromov-Witten invariants, since $`m_d`$ is conjecturally equal to $`(1)^{d1}n_d/(3d)`$, where $`n_d`$ is the local Gromov-Witten invariant (see Remark 2.2). (2) A. Gathmann informed the author that this is true for $`d8`$(i.e. where $`m_d`$ is calculated). ### 1.2. B-model The result of \[T2\] suggests that the mirror manifold of $`^\times `$ is $`^\times `$. In A-model, we associate a parameter to each point over $`0`$ or $`\mathrm{}`$ on a cover of $`^\times `$, and in B-model this corresponds to the choice of points on $`^\times `$: the former can be considered as “Kähler moduli” and the latter as “complex moduli”. In B-model, we used the points as the boundary of integrals to define coordinate functions. With this in mind, we look for a function which has an expansion with coefficients $`m_d`$. Although the calculation below is essentially the same as in \[CKYZ\], here we see it from the point of view of Mirror Symmetry of log Calabi-Yau surfaces: we claim that the mirror of ($`^2\backslash (\text{smooth cubic})`$ & Kähler moduli) is ($`^\times \times ^\times `$ & cubic), where the latter cubic has complex moduli and we take it as the boundary of integrals, just as in the case of $`^\times `$. We consider homogeneous coordinates $`X,Y,Z`$ and inhomogeneous coordinates $`x,y`$ on $`^2`$. Let $`\mathrm{\Omega }`$ be the logarithmic 2-form $`dx/xdy/y`$ on $`X_0:=^2\backslash \{XYZ=0\}`$ and write $`B_\varphi `$ for the plane cubic defined by $`XYZ\varphi (X^3+Y^3+Z^3)=0`$. Then we consider the integral $$I:=_{\mathrm{\Gamma }_\varphi }\mathrm{\Omega },$$ where $`\mathrm{\Gamma }_\varphi `$ is a 2-chain in $`X_0`$ whose boundary has support on $`B_\varphi `$. ###### Lemma 1.8. We have $$\varphi \frac{dI}{d\varphi }=_\mathrm{\Gamma }𝑑x/(x3\varphi y^2).$$ ###### Proof. If we fix $`x`$ and differentiate $`xy\varphi (x^3+y^3+1)=0`$, we have $`dy/d\varphi =xy/\varphi (x3\varphi y^2)`$. ∎ So, if we set $`z:=\varphi ^3`$ and $`\theta :=z\frac{d}{dz}`$, we have $$\{\theta ^33z\theta (3\theta +1)(3\theta +2)\}I=0.$$ The following functions form a basis of the space of solutions: $`I_1`$ $`=`$ $`1,`$ $`I_2`$ $`=`$ $`\mathrm{log}z+I_2^{(0)},`$ $`I_3`$ $`=`$ $`I_2\mathrm{log}z{\displaystyle \frac{(\mathrm{log}z)^2}{2}}+I_3^{(0)},`$ where $`I_2^{(0)}`$ $`=`$ $`6z+45z^2+560z^3+{\displaystyle \frac{17325}{2}}z^4+{\displaystyle \frac{756756}{5}}z^5+2858856z^6`$ $`+{\displaystyle \frac{399072960}{7}}z^7+{\displaystyle \frac{4732755885}{4}}z^8+\mathrm{},`$ $`I_3^{(0)}`$ $`=`$ $`9z+{\displaystyle \frac{423}{4}}z^2+1486z^3+{\displaystyle \frac{389415}{16}}z^4+{\displaystyle \frac{21981393}{50}}z^5+{\displaystyle \frac{16973929}{2}}z^6`$ $`+{\displaystyle \frac{8421450228}{49}}z^7+{\displaystyle \frac{1616340007953}{448}}z^8+\mathrm{}.`$ The monodromy of $`I_3`$ for $`ze^{2\pi i}z`$ is $`2\pi i(I_2\pi i)`$, which we denote by $`\stackrel{~}{I}_2`$. We denote the monodromy $`(2\pi i)^2`$ of $`\stackrel{~}{I}_2`$ by $`\stackrel{~}{I}_1`$. Now we write $`I_3`$ in terms of $`q:=e^{2\pi i\stackrel{~}{I}_2/\stackrel{~}{I}_1}=e^{I_2}=ze^{I_2^{(0)}}`$: $`I_3`$ $`=`$ $`{\displaystyle \frac{(I_2)^2}{2}}+9z+{\displaystyle \frac{351}{4}}z^2+1216z^3+{\displaystyle \frac{319455}{16}}z^4+{\displaystyle \frac{18122643}{50}}z^5`$ $`+{\displaystyle \frac{35161224}{5}}z^6+{\displaystyle \frac{7009518168}{49}}z^7+{\displaystyle \frac{1350681750297}{448}}z^8+\mathrm{}`$ $`=`$ $`{\displaystyle \frac{(\mathrm{log}(q))^2}{2}}+09q+{\displaystyle \frac{135}{4}}q^2244q^3+{\displaystyle \frac{36999}{16}}q^4{\displaystyle \frac{635634}{25}}q^5`$ $`+307095q^6{\displaystyle \frac{193919175}{49}}q^7+{\displaystyle \frac{3422490759}{64}}q^8+\mathrm{}`$ $`=`$ $`{\displaystyle \frac{(\mathrm{log}(q))^2}{2}}3^2.1{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^k}{k^2}}+6^2.1{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^{2k}}{k^2}}9^2.3{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^{3k}}{k^2}}+12^2.16{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^{4k}}{k^2}}`$ $`15^2.113{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^{5k}}{k^2}}+18^2.948{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^{6k}}{k^2}}21^2.8974{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^{7k}}{k^2}}`$ $`+24^2.92840{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^{8k}}{k^2}}+\mathrm{}.`$ ###### Remark 1.9. $`K:=(zdI_2/dz)^3d^2I_3/dI_2^2`$ satisfies $`dK/dz=(27/(127z))K`$, and therefore we have $`K=1/(127z)`$. Comparing the coefficients with the table in Theorem 1.4, we propose: ###### Conjecture 1.10. $$I_3=\frac{(\mathrm{log}(q))^2}{2}+\underset{d=1}{\overset{\mathrm{}}{}}(1)^d(3d)^2m_d\underset{k=1}{\overset{\mathrm{}}{}}\frac{q^{dk}}{k^2}.$$ ###### Remark 1.11. If we assume Conjecture 1.6, the previous conjecture is equivalent to $$I_3=\frac{(\mathrm{log}(q))^2}{2}+\underset{d=1}{\overset{\mathrm{}}{}}(1)^d[M_{3d}^B(^2,d)]^{virt}q^d,$$ and in fact this may be a more natural equality. According to A. Gathmann, his algorithm in \[G\] can be used to prove this. ## 2. Log Mirror and Local Mirror In \[CKYZ\], the generating function of the “numbers of rational curves in a local Calabi-Yau 3-fold” was given. Let $`^2`$ be embedded in a Calabi-Yau 3-fold and denote by $`n_d`$ the contribution of rational curves of degree $`d`$ in $`^2`$ to the number of rational curves in the Calabi-Yau 3-fold. Let $`\overline{M}_{0,0}`$ be the moduli of stable maps of genus 0 curves to $`^2`$ with degree $`d`$ images and $`U`$ the vector bundle over $`\overline{M}_{0,0}`$ whose fiber at the point $`[f:C^2]`$ is $`H^1(C,f^{}K_^2)`$. Then, the Chern number $`K_d:=c_{3d1}(U)`$ is equal to $`_{k|d}n_{d/k}/k^3`$. ###### Theorem 2.1. (\[CKYZ\]) $`I_3`$ $`=`$ $`{\displaystyle \frac{(\mathrm{log}(q))^2}{2}}{\displaystyle \underset{d=1}{\overset{\mathrm{}}{}}}3dK_dq^d`$ $`=`$ $`{\displaystyle \frac{(\mathrm{log}(q))^2}{2}}{\displaystyle \underset{d=1}{\overset{\mathrm{}}{}}}3dn_d{\displaystyle \underset{k=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{q^{dk}}{k^2}}.`$ ###### Remark 2.2. Thus Conjecture 1.10 is equivalent to $`n_d=(1)^{d1}3dm_d`$, and this equality holds for $`d8`$ by Theorem 1.4 and the $`q`$-expansion of $`I_3`$ in the previous section. We give a heuristic argument as to why this should hold, although it contains serious gaps as explained later. By Serre duality, the dual of $`U`$ is isomorphic to the vector bundle $`V`$ over $`\overline{M}_{0,0}`$ whose fiber at $`[f:C^2]`$ is $`H^0(C,K_Cf^{}𝒪_^2(B))`$. Since the rank of the bundles is $`3d1`$, we have $`c_{3d1}(V)=(1)^{d1}K_d`$. Let $`\overline{M}_{0,1}`$ be the moduli of stable maps of 1-pointed genus 0 curves to $`^2`$ with degree $`d`$ images, $`M_{0,1}`$ the open subset of $`\overline{M}_{0,1}`$ representing $`f:(^1,P)^2`$ and $`\pi :\overline{M}_{0,1}\overline{M}_{0,0}`$ the projection. Further, let $`E_1`$ be the line bundle over $`\overline{M}_{0,1}`$ whose fiber at $`f:(C,P)^2`$ is $`H^0(C,f^{}𝒪_^2(B)𝒪_P)`$ and $`L`$ the line bundle whose fiber is $`H^0(C,𝒪_P(P))`$. Then, we have $`c_{3d}(\pi ^{}VE_1)=3dc_{3d1}(V)`$, for the zero set of a section of $`E_1`$ induced by a defining equation of $`B`$ is the set of the points corresponding to $`f:(C,P)^2`$ such that $`f(P)B`$, and there are $`3d`$ such points for any $`f:C^2`$. We have an exact sequence $$0K_CK_C(P)\stackrel{\text{residue}}{}𝒪_C0.$$ If $`C`$ is irreducible, i.e. isomorphic to $`^1`$, we obtain exact sequences $`0`$ $``$ $`H^0(C,K_C((k+1)P)f^{}𝒪_^2(B))H^0(C,K_C(kP)f^{}𝒪_^2(B))`$ $``$ $`H^0(C,O_P((k+1)P)f^{}𝒪_^2(B))0`$ for $`0k3d2`$. We also have $`H^0(C,K_C((3d1)P)f^{}𝒪_^2(B))=0`$. Thus, on $`M_{0,1}`$, we have a filtration of $`\pi ^{}VE_1`$ such that the associated graded module is isomorphic to $`_{i=0}^{3d1}E_1L^i`$. On the other hand, there are about $`(3d)^2m_d`$ plane curves of degree $`d`$ satisfying (AL), since the number of $`3d`$-torsions on $`B`$ is $`(3d)^2`$. They are in one-to-one correspondence with points $`(f:(^1,P)^2)M_{0,1}`$ such that $`f`$ is birational onto the image and that $`f^{}B=3dP`$. Consider the vector bundle $`E_{3d}`$ of rank $`3d`$ over $`\overline{M}_{0,1}`$ whose fiber at $`f:(C,P)^2`$ is $`H^0(C,f^{}𝒪_^2(B)𝒪_{3dP})`$. On $`M_{0,1}`$, the zero set of a section of $`E_{3d}`$ induced by a defining equation of $`B`$ is the set of the points $`f:(C,P)^2`$ such that $`f^{}B=3dP`$. Now, from the exact sequences $$0𝒪_P(kP)𝒪_{(k+1)P}𝒪_{kP}0,$$ we see that $`E_{3d}`$ has a filtration such that the associated graded module is isomorphic to $`_{i=0}^{3d1}E_1L^i`$. Thus we may expect $`(3d)^2m_d(1)^{3d1}3dK_d(1)^{3d1}3dn_d`$. Rigorously, however, this argument makes little sense. First, the section of $`E_{3d}`$ induced by a defining equation of $`B`$ has undesirable zeros in $`\overline{M}_{0,1}\backslash M_{0,1}`$. For example, if $`C=C_1C_2C_3`$(a chain in this order), $`PC_2`$ and $`C_2`$ maps to a point $`Q`$ in $`B`$, we may take any rational curves through $`Q`$ as the images of $`C_1`$ and $`C_3`$. Thus the number $`(3d)^2m_d`$ may be much different from $`c_{3d}(E_{3d})`$. Second, we have a filtration of $`\pi ^{}VE_1`$ merely on $`M_{0,1}`$. Let $`\overline{M}_i^B(^2,d)`$ be as in Definition 1.5. The section of the line bundle $`(E_1L^i)|_{\overline{M}_i^B(^2,d)}`$ induced by a defining equation of $`B`$ vanishes when $`f^1B(i+1)P`$ is satisfied. Then, \[G\] describes the difference between $`c_1(E_1L^i).[\overline{M}_i^B(^2,d)]^{virt}`$ and $`[\overline{M}_{i+1}^B(^2,d)]^{virt}`$. This should account for the difference between $`c(_{i=0}^{3d1}E_1L^i)`$ and $`c(\pi ^{}VE_1)`$.
warning/0004/cond-mat0004413.html
ar5iv
text
# Superconducting 𝑑_{𝑥²-𝑦²}±𝑖⁢𝑑_{𝑥⁢𝑦} phase glass ## I INTRODUCTION It has been known for a long time that magnetic ions in a superconductor can form a spin glass (SG) state. There is growing experimental evidence for the formation of a SG phase in the high-$`T_c`$ materials once magnetic ions, such as Fe and Ni, are doped into the system. Since magnetic ions interact with the superconducting condensate it is natural to expect that this interaction will cause frustration of the underlying condensate and eventualy might lead to a superconducting glass or phase glass (PG). In all of the above discussions of the role of impurity spins it was assumed that magnetic scattering frustrates and suppresses the superconductivity. There are additional physical effects that were not addressed in previous work: namely how frustrated localized spins in an unconventional superconductors can distort the condenstate and produce patches of secondary components of the order parameter near the localized impurity, see Fig. 1. It has been shown that a magnetic impurity in a $`d_{x^2y^2}`$ superconductor with order parameter $`\mathrm{\Delta }^0`$ induces a local complex $`d_{xy}`$ component of the order parameter which we designate $`\mathrm{\Delta }^1`$. Locally near each impurity site, on the scale of the coherence length $`\xi `$, there is a patch of $`\mathrm{\Delta }^0+i\mathrm{\Delta }^1`$ order paramter for $`S^z=+1`$ and $`\mathrm{\Delta }^0i\mathrm{\Delta }^1`$ for $`S^z=1`$, depending on the sign of the impurity spin. Hereafter we assume classical spins with $`S=1`$. This is a reasonable assumption taken the fact that magnetic ions substituted into high-$`T_c`$ superconductors have a large spins, such as Ni (S=1). Here we will explore the coupling between different patches due to the Josephson coupling. Phase coherence between these patches of $`\mathrm{\Delta }^0\pm i\mathrm{\Delta }^1`$ would lower the kinetic energy of the condensate and would tend to align all the patches into globally coherent $`\mathrm{\Delta }^0+i\mathrm{\Delta }^1`$, or its conjugate. This would imply ferromagnetic ordering of the impurity spins $`S^z`$. On the other hand, the dipolar and RKKY spin-spin interaction terms, which we show are mainly antiferromagnetic, would frustrate this ferromagnetic order. In fact, we find that the spin-spin interaction is frustrating and produces a SG phase at $`T<T_{SG}`$ in the absence of a coupling between the superconducting condenstate and spin degrees of freedom. When this coupling is considered the result depends on the relative strength of the two interactions. We find that since the spin-dependent part of the Josephson coupling is small the magnetic susbsystem will drive the phase transition, at least at low and intermediate impurity concentrations. Given that the impurity spins form a SG phase we are led to the question how the phase of the induced component $`\pm i\mathrm{\Delta }^1`$ is affected by spin frustration. We find that the spin frustration in the SG phase will frustrate the phase of the $`d_{xy}`$ component and a PG will form at low temperatures $`T<T_{PG}<T_{SG}`$. More generally we will discuss the possible phases of the coupled spin-phase model, and at higher impurity concentrations we will consider the formation of a ferromagnetic phase. When discussing glassy phases, we will consider states where the spatial average $`S^z=0`$. However at any given patch, $`i`$, the time averaged $`S_i^z_\tau 0`$. Similarly we will consider site and time averaged induced component $`\mathrm{\Delta }^1=0`$ and $`\mathrm{\Delta }_i^1_\tau 0`$ respectively. The additional new order parameter we find relevant is the product $`S^z\mathrm{\Delta }^1`$. The new order in this case arises from the possibilty of having fully disordered phases $`\mathrm{\Delta }^1=S^z=0`$ and still having the true long range order in $`S^z\mathrm{\Delta }^z0`$. Physically it corresponds to phase locking of the patches to the local value of impurity spin $`S^z`$. This order parameter has no analog in purely magnetic spin glass systems and is a direct consequence of the spin-phase order parameter coupling discussed above. In all of the discussion hereafter we assume that the “backbone” order parameter $`\mathrm{\Delta }^0`$ has true long range order as it is robust at low temperatures $`TT_c100K`$. The plan of this paper is as follows: in the first section we explore the effective spin-spin coupling and integrate out the superconducting degrees of freedom by performing an RKYY calculation. The effective spin model is found to give rise to a SG phase. In the second section of the paper the Josephson coupling between different patches is considered. It is found that the interaction favors a FM phase. In the last section we dicuss the PG phase, where the induced superconducting order parameter is locked to the impurity spin. Furthermore we will discuss the possibilities of para- and ferromagnetic phases. In related recent work Simon and Varma treat the magnetic impurity problem by a variational approach. They concentrate on the single impurity, but conclude by arguing that the formation of a ferromagnetic phase is unlikely and that a spin glass state is much more likely. ## II EFFECTIVE SPIN MODEL In this section we will consider the effective spin model that describes the interaction between impurity spins surrounded by patches of induced complex $`d_{xy}`$ order parameter, as shown in Fig. 1. The most relevant sources of interaction between the impurity spins are the electron-mediated RKKY-interaction and the direct dipolar magnetic interaction. The standard RKKY interaction is mediated through an interaction of the form $`JSI`$, where $`S`$ referes to the spin of the conduction electron, and $`I`$ denotes the spin of the impurity. Of particular interest to this work, however, is another term that is directly related to the induced order parameter. The $`LI`$ interaction, where $`L`$ refers to the angular momentum of a conduction electron, scatters an electron into the complex $`d_{xy}`$ phase. A second order RKKY calculation where the interaction potential is taken to be of the from $`LI`$ will therefore be relevant in this context. The effective spin Hamiltonian is thus of the form $$H=H^M+H^{SI}+H^{LI},$$ (1) where the dipolar term $`H^M`$ and the electron mediated interactions $`H^{SI}`$ and $`H^{LI}`$ are given by $`H^M`$ $`=`$ $`{\displaystyle \underset{ij}{}}{\displaystyle \frac{1}{r_{ij}^3}}[S_iS_j3(S_i\widehat{r}_{ij})(S_j\widehat{r}_{ij})]`$ (2) $`H^{SI}`$ $`=`$ $`{\displaystyle \underset{ij}{}}J_{ij}^{SI}S_iS_j`$ (3) $`H^{LI}`$ $`=`$ $`{\displaystyle \underset{ij}{}}J_{ij}^{LI}S_i^zS_j^z`$ (4) The size of the dipolar term for two spins separated by $`1`$ Åcan be estimated to be about $`10^4`$ eV, while the RKKY interaction depends strongly on the coupling between the conduction electrons and the impurity spin. For Mn ions in alloys is has been estimated to be on the order of $`10^1`$ eV. We are not aware of direct estimates of the RKKY interaction in high-$`T_c`$ materials, but it is likely to be greater than the dipolar interaction. The spatial decay of the last two terms are explicitly calculated in the appendix and are of the form $`H^{LI}`$ $``$ $`{\displaystyle \frac{2J^2I_1^zI_2^zk_F^2}{vr^3}}[0.37+0.33\mathrm{sin}(2Kr)]`$ (5) $`H^{SI}`$ $``$ $`{\displaystyle \frac{J^2I_1I_2}{2\pi vr^3}}[0.37+0.33\mathrm{sin}(2Kr)]`$ (6) where $`v`$ is the gap velocity and K is a momentum cut-off. The RKKY terms thus share a cubic decay with the dipolar interaction, but are found to be completely antiferromagnetic. However, independently of the specific form of the terms, general consideration tells us that as long as there are sufficient amounts of disorder and frustration present the SG phase should be realized. The impurity spins we are considering are randomly distributed, and antiferromagnetic interactions will therefore lead to frustration. So as long as the effective interactions are not primarily ferromagnetic the model should have a SG groundstate. The dipolar interaction is antiferromagnetic in the $`z`$-component of the spins in an $`xy`$ plane, and ferromagnetic in the plane. As mentioned above the RKKY terms are completely antiferromagnetic. We therefore conclude that the effective spin model, consisting of the combined dipolar and RKKY terms, exhibits a spin glass phase at low temperatures. For large enough quantum fluctuations it would also be possible to realize a quantum paramagnet phase, where any freezing is destroyed by quantum fluctuations. We focus here on large classical spins and assume that quantum fluctuations are negligible. In principle similar considerations can be given for $`S=1/2`$ impurity spins, in which case the paramagnetic phase could be realized. Note in particular that the impurity spins would form a spin glass even without the $`H^{LI}`$ term, and that this term is second order in $`\mathrm{\Delta }^1\mathrm{\Delta }^0`$ . This is the motivation why we have first considered the effective spin model and determined the spin configuration. So although the spin and superconducting degrees of freedom are connected, as we will see in the next section, we will there take the spin configuration as given, and determine the phase of the superconducting order parameter based on this. Note also that the importance of the dipolar term may increase as the $`\mathrm{\Delta }_1`$ component is induced, since this will open up a fully gapped state which will tend to exponentially decrease the long-range RKKY interaction. The short-range RKKY interaction should not be significantly affected by the opening of the superconducting gap. ## III JOSEPHSON COUPLING In this section we will consider the Josephson coupling between the different patches of induced order parameter. Let us first consider the order parameter around an impurity spin $$\mathrm{\Psi }_i=(\mathrm{\Delta }_i^0+e^{i\frac{\pi }{2}S_i^z}\mathrm{\Delta }_i^1)e^{i\theta _i},$$ (7) where the real parameter $`\mathrm{\Delta }^0`$ denotes the $`d_{x^2y^2}`$ order parameter, and the likewise real quantity $`\mathrm{\Delta }^1`$ denotes the induced $`d_{xy}`$ component. Hereafter we assume that $`\mathrm{\Delta }^0`$ is a robust order parameter that develops true long-range order at $`T<T_c90`$ K and remains ordered in all of the phases we discuss. The impurity spin $`S_i^z`$ uniquely determines the relative phase of the induced order parameter to be $`+\pi /2`$ for $`S^z=+1`$ and $`\pi /2`$ for $`S^z=1`$. The main result of the previous section was that the impurity spins, considered independently of the coupling to the induced order parameter, will exhibit a spin glass phase at low temperatures. Let us next consider what effects the inter-patch Josephson interaction may have. In order to address this question we examine the Josephson coupling between different patches given by $$H^J=\underset{ij}{}|I_{ij}|(\mathrm{\Psi }_i^{}\mathrm{\Psi }_j+\mathrm{\Psi }_i\mathrm{\Psi }_j^{}),$$ (8) Using the local order parameter Eq. (7) this leads to an interaction of the form $$H^J=2\underset{ij}{}|I_{ij}|[(\mathrm{\Delta }^0)^2+(\mathrm{\Delta }^1)^2S_i^zS_j^z]\mathrm{cos}(\theta _i\theta _j)$$ (9) We have here assumed that $`\mathrm{\Delta }_i^0=\mathrm{\Delta }_j^0=\mathrm{\Delta }^0`$, and likewise for $`\mathrm{\Delta }_1`$. The first term, which is zeroth order in $`\mathrm{\Delta }_1`$, wants to align the phase of $`\mathrm{\Delta }_i^0`$ at different patches by setting $`\theta _i=\theta _j`$. The second term wants to align the impurity spin in an ferromagnetic phase. The Josephson coupling thus favors a ferromagnetic spin configuration, while the effective RKKY and dipolar spin model favors a spin glass. In the next section we will consider some possible outcomes of this competition. We would, however, like to point out that the RKKY and Josephson effects are not as independent as they may at first seem. Both are mediated by electrons and while the dominant part of the RKKY interaction in the superconducting state is zeroth order in $`\mathrm{\Delta }^1`$ there is a part that is second order in $`\mathrm{\Delta }^1`$, corresponding to the Josephson coupling. The Josephson interaction physically expresses an effect arising from electron pair tunneling, while the GG part of the RKKY interaction expresses an electron-hole channel. Here G and F are the normal and anomalous propagators in superconducting state. The FF part of the RKKY interaction is closer to the Josephson interaction since it expresses an electron-electron process. Furthermore, there is a second order contribution to the RKKY interaction which we have not evaluated in this work. This contains two explicit $`d_{xy}`$ propagators, with two exchange interactions at the impurity spins. This part of the interaction should be similar to the Josephson interaction since explicit information about the patches, such as spatial decay, should be contained in the propagators. The physical effects of this term should, however, be included in the Josephson coupling considered above. ## IV POSSIBLE PHASES In this section we will consider the different phases that could occur as a result of the competition between the RKKY and the Josephson terms. First we will enumerate the different phases and thereafter we will discuss them in some more detail. The simplest phase is a ferromagnetic (FM) phase, favored by the Josephson coupling of different patches. The FM phase is characterized by a finite spatial average of the magnetization and the induced phase; $`S^z0`$ and $`\mathrm{\Delta }_10`$. The phase favored by the effective spin Hamiltonian, on the other hand, is a SG phase, where the spatial average of the magnetization vanishes $$S^z=0,$$ (10) but the time averaged local magnetization remains finite; $$S_i^z_\tau 0.$$ (11) Assuming that the phase of $`\mathrm{\Delta }^1`$ is determined by the impurity spin this gives rise to similar ordering for the induced phase; $`\mathrm{\Delta }^1=0`$ and $`\mathrm{\Delta }^1_\tau 0`$. In this case the combination $`S^z\mathrm{\Delta }^1`$ will also have a non-vanishing spatial average $$S^z\mathrm{\Delta }^10$$ (12) This particular order parameter is unique for the coupling of SG and SC degrees of freedom and is not present in a purely magnetic system. It describes the phase-locking of the superconducting phase and the impurity spin. There also exists the possibility that the induced order parameter $`\mathrm{\Delta }^1`$ does not phase-lock with the impurity spin, but prefers to vary over length scales greater than the inter-patch distance. In that case the spatial average $`S^z\mathrm{\Delta }^1`$ would vanish, but the local time average $`S_i^z\mathrm{\Delta }_i^1_\tau `$ would still remain finite. If quantum fluctuations are strong enough, then it would also be possible to realize a paramagnetic spin phase. In this case also the time averages $`S^z_\tau `$ and $`\mathrm{\Delta }^1_\tau `$ would vanish, and only local $`S^z\mathrm{\Delta }^1`$ would remain finite, for as long as the phase-locking is maintained. In this work we will, however, neglect this possibility since we focus on large classical spins. After having considered the different options let us consider the interplay between the different terms. If there is a low concentration of impurity spins (the patches do not overlap), then the ferromagnetic Josephson term is bound to be exponentially small, and the ground state should be a spin glass, characterized by a non-vanishing spatial average $`S^z\mathrm{\Delta }^1`$. If the impurity concentration becomes larger, but not large enough to kill the $`\mathrm{\Delta }^0`$ superconductivity, then there may be a region where the ferromagnetic Josephson terms are predominant and the FM phase is formed. The FM phase was considered in detail in a previous Ginzburg-Landau description . In Fig. 2 we present a phase diagram as a function of impurity concentration and temperature. Note that this PG phase is different from previously proposed superconducting phase glasses in that the glassy behavior is only displayed in the induced component of the order parameter, and not in the robust $`d_{x^2y^2}`$ part. Considering the phase diagram we note that the disorder suppresses the critical temperature for the superconductor-metal insulator, as is well known. The impurity spins form a spin-glass phase at low temperatures, and this phase is independent of the electronic order parameter and persists also in the metallic phase. The PG phase is induced by the SG and SC phases and hence it must only exist within the boundaries of these two phases. The possible FM phase is induced by a large impurity concentration. As the $`d_{x^2y^2}`$ superconducting order parameter gets suppressed by disorder the induced component will also vanish, and hence the FM phase gets strongly suppressed as we approach $`x_c`$. Experimentally the proposed phase locked state can be observed in scanning tunneling microscope measurements, where the particle spectrum would develop a full gap near impurity site even though the phase of $`\mathrm{\Delta }^1`$ remains uncertain. The ac magnetic susceptibility in the superconducting state also should show features upon crossing the SG and PG lines. Another experiment, that would be sensitive to the appearance of the $`S^z\mathrm{\Delta }_1`$, would be the penetration depth that would become exponential $`\delta 1/\lambda ^2\mathrm{exp}[|\mathrm{\Delta }^1|/T]`$ in the PG and FM phases. ## V CONCLUSION We have examined the role of magnetic impurities in a d-wave superconductors. In particular we have studied the effective impurity spin model arising from electron mediated RKKY and magnetic dipolar terms and argued that these terms lead to a SG phase at low temperatures.. Furthermore we have analyzed the coupling between the spin- and superconductor order parameter that arises from the Josephson interaction of patches of induced order parameter around the impurity spins. The Josephson interaction favors a FM phase. At high impurity concentrations the FM phase may be realized, while at low concentrations the SG phase would be prefered. Due to the coupling between the spin and superconducting order parameters the SG phase induces a superconducting PG at low temperatures. The glassy behavior is a property of the induced $`d_{xy}`$ component of the order parameter, while the primary component $`d_{x^2y^2}`$ is assumed robust. The superconducting PG phase is characterized by an order parameter of the form $`S^z\mathrm{\Delta }^1`$, which describes the phase-locking of the induced order parameter and the local impurity spin. In addition to the PG and FM phases we have also discussed a possible paramagnetic phases. We are grateful to D. Agterberg, N. Bonesteel, M. Graf and I. Martin, for useful discussions. This work was supported by US DOE and NSF Grant No. DMR-9629987. ## A The RKKY interaction The RKKY interaction describes a second-order process, where an electron of momentum k interacts with an impurity spin at $`r=R_1`$, is scattered to a state with momentum q and interacts with a another impurity spin at $`r=R_2`$, where it is scattered back into the original state. This process is described by the following diagram: Using Feynman rules we get the following expression for the effective Hamiltonian $`iH`$ $`=`$ $`(1){\displaystyle }{\displaystyle \frac{d^dk}{(2\pi )^d}}{\displaystyle }{\displaystyle \frac{d^dq}{(2\pi )^d}}{\displaystyle }{\displaystyle \frac{d\omega }{(2\pi )}}\times `$ (A2) $`iG^0(\omega ,k)(iV_{k\alpha ,q\beta }^{R^1})iG^0(\omega ,q)(iV_{q\beta ,k\alpha }^{R^2})`$ $`iH`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^dk}{(2\pi )^d}}{\displaystyle }{\displaystyle \frac{d^dq}{(2\pi )^d}}{\displaystyle }{\displaystyle \frac{d\omega }{(2\pi )}}\times `$ (A4) $`G^0(\omega ,k)V_{k\alpha ,q\beta }^{R^1}G^0(\omega ,q)V_{q\beta ,k\alpha }^{R^2}`$ Using the Nambu formalism for the superconductor this expression transforms to $`H`$ $`=`$ $`{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}{\displaystyle }{\displaystyle \frac{d^2q}{(2\pi )^2}}{\displaystyle }{\displaystyle \frac{d\omega }{(2\pi i)}}\times `$ (A6) $`\text{Tr}\left\{[G(\omega ,k)][V_{k,q}^{R^1}\tau _0][G(\omega ,q)][V_{q,k}^{R^2}\tau _0]\right\},`$ where $`G(\omega ,k)`$ $`=`$ $`\left[\begin{array}{cc}G_{11}(\omega ,k)& F(\omega ,k)\\ F(\omega ,k)& G_{22}(\omega ,k)\end{array}\right]`$ (A9) $`\tau _0`$ $`=`$ $`\left[\begin{array}{cc}1& 0\\ 0& 1\end{array}\right]`$ (A12) $`F(\omega ,k)`$ $`=`$ $`{\displaystyle \frac{\mathrm{\Delta }_k}{\omega ^2E_k^2+i\delta }}`$ (A13) $`G_{11}(\omega ,k)`$ $`=`$ $`{\displaystyle \frac{\omega +ϵ_k}{\omega ^2E_k^2+i\delta }}`$ (A14) $`G_{22}(\omega ,k)`$ $`=`$ $`{\displaystyle \frac{\omega ϵ_k}{\omega ^2E_k^2+i\delta }},`$ (A15) where $`E_k=\sqrt{ϵ_k^2+\mathrm{\Delta }_k^2}`$. For a tight binding model $`ϵ_k=t\left[\mathrm{cos}(k_xa)+\mathrm{cos}(k_ya)\right]`$, and the $`d_{x^2y^2}`$ gap is given by $`\mathrm{\Delta }_k=\mathrm{\Delta }_0\left[\mathrm{cos}(k_xa)\mathrm{cos}(k_ya)\right]`$. Next we will consider a few specific forms of the interaction $`V_{k\alpha ,q\beta }^R`$, which is the Fourier transform of the electron-impurity spin interaction. For free electrons $`|k=e^{ikx}`$ and we get $$V_{k\alpha ,q\beta }^R=k\alpha |V^R|q\beta =d^dxe^{i(kq)x}\alpha |V^R|\beta $$ (A16) Let us first consider a contact potential of the form $`V^R=J\delta (xR)SI`$, where $`S`$ denotes the electron spin and $`I`$ the impurity spin. This results in $$V_{k\alpha ,q\beta }^R=Je^{i(kq)R}\alpha |SI|\beta $$ (A17) Next we will consider a $`LI`$ interaction, where $`L`$ is the angular momentum of the electron, and $`I`$ is the spin of the impurity at location $`R`$. The electron moves relative to the impurity spin, which sees a magnetic field of the form $`B\frac{L}{|xR|^3}`$, where $`x`$ is the location of the electron. We will consider two spatial dimensions, and since the angular momentum of the electron will have only a $`z`$-component we consider an interaction of the form $$V^R=J\frac{LI}{|xR|^3}=J\frac{L^zI^z}{|xR|^3},$$ (A18) where $`L=(xR)\times p=i(xR)\times _x`$. We get the following expression for the matrix element: $$V_{k\alpha ,q\beta }^R=Jd^2xe^{ikx}\frac{[i(xR)\times _x]^z}{|xR|^3}I^ze^{iqx}\alpha |\beta $$ (A19) After performing the integrals we arrive at $$V_{k\alpha ,q\beta }^R=2\pi iJI^ze^{i(kq)R}\frac{(k\times q)^z}{|kq|}\delta _{\alpha ,\beta }$$ (A20) We are now in a position to evaluate the effective interaction $`H`$ $`=`$ $`i{\displaystyle }{\displaystyle \frac{d^2k}{(2\pi )^2}}{\displaystyle }{\displaystyle \frac{d^2q}{(2\pi )^2}}{\displaystyle }{\displaystyle \frac{d\omega }{(2\pi )}}V_{k,q}^{R^1}V_{q,k}^{R^2}\times `$ (A23) $`[G_{11}(\omega ,k)G_{11}(\omega ,q)+G_{22}(\omega ,k)G_{22}(\omega ,q)`$ $`+2F(\omega ,k)F(\omega ,q)].`$ We will begin with the frequency integral for the FF contribution $$I_F=\frac{d\omega }{(2\pi )}2F(\omega ,k)F(\omega ,q).$$ (A24) There are poles at $`\omega =\pm \sqrt{E_k^2i\delta }=E_k\pm i\delta `$. Closing the integral in the upper half plane leads to contributing poles at $`\omega =E_k+i\delta `$ and $`\omega =E_q+i\delta `$. It follows that $$I_F=i\frac{\mathrm{\Delta }_k\mathrm{\Delta }_q}{E_kE_q(E_k+E_q)}$$ (A25) Next we will consider the GG contribution $$I_G=\frac{d\omega }{(2\pi )}G_{11}(\omega ,k)G_{11}(\omega ,q)+G_{22}(\omega ,k)G_{22}(\omega ,q)$$ (A26) Proceeding as above it follows that $$I_G=i\frac{(ϵ_kϵ_qE_kE_q)}{E_kE_q(E_k+E_q)}.$$ (A27) We will start by considering an interaction of the $`L\mathrm{cot}I`$ kind. We then have $$V_{k,q}^{R^1}V_{q,k}^{R^2}=4\pi ^2J^2I_1^zI_2^ze^{i(kq)r}\left|\frac{k\times q}{kq}\right|^2,$$ (A28) where $`r=R_1R_2`$. For the effective interaction we then get $`H`$ $`=`$ $`{\displaystyle \frac{J^2I_1^zI_2^z}{4\pi ^2}}{\displaystyle }d^2k{\displaystyle }d^2qe^{i(kq)r}\times `$ (A30) $`\left|{\displaystyle \frac{k\times q}{kq}}\right|^2{\displaystyle \frac{(\mathrm{\Delta }_k\mathrm{\Delta }_q+ϵ_kϵ_qE_kE_q)}{E_kE_q(E_k+E_q)}}`$ In order to solve the above integral we introduce the nodal point approximation. The dominant contribution to the integral should come from each nodal point, and we perform a rotation and translation to transform the origin to the nodal points, with the $`x`$-axis along the tight binding Fermi surface, see Fig. 4. The transformation is given by $$\left(\begin{array}{c}k_x^{}\\ k_y^{}\end{array}\right)=\left(\begin{array}{cc}\mathrm{cos}\theta & \mathrm{sin}\theta \\ \mathrm{sin}\theta & \mathrm{cos}\theta \end{array}\right)\left(\begin{array}{c}k_x\\ k_y\end{array}\right)\left(\begin{array}{c}0\\ k_F\end{array}\right),$$ (A31) where $`k_F=\pi /\sqrt{2}a`$ and $`\theta =\{\pi /4,3\pi /4,5\pi /4,7\pi /4\}`$ for the four nodes. Next we need to apply these transformations to all quantities in the effective interaction. The energy of a tight-binding model, $`ϵ_k=t\left[\mathrm{cos}(k_xa)+\mathrm{cos}(k_ya)\right]`$, will transform according to $`ϵ_k=v_Fk_y^{}`$ for all nodes, and the gap function, $`\mathrm{\Delta }_k=\mathrm{\Delta }_0\left[\mathrm{cos}(k_xa)\mathrm{cos}(k_ya)\right]`$ will transform according to $`\mathrm{\Delta }_k=\pm v_\mathrm{\Delta }k_x^{}`$, with a positive sign for nodes 1 and 3, and a negative sign for nodes 2 and 4. Assuming an isotropic superconductor $`v=v_F=v_\mathrm{\Delta }`$ this leads to $`E_k=vk^{}`$. The rotation and translation cannot affect $`|kq|`$, and therefore $`|kq|=|k^{}q^{}|`$. Furthermore $`(k\times q)^z`$ $`=`$ $`k_xq_yk_yq_x`$ (A32) $`=`$ $`k_x^{}q_y^{}k_y^{}q_x^{}+k_F(k_x^{}q_x^{})`$ (A33) for all the four nodes. The second term is linear in the momentum and will be retained. The final term to be transformed, $`(kq)r`$, is considered next. This term depends on the direction of $`r`$, and different nodes will give different contributions. Assume that $`r`$ is fixed in a direction $`\theta _r`$, and the contribution to the effective interaction from node 1 has been calculated. Since all other terms are identical, the contribution from the other nodes must be given by substituting $`\{\theta _r+\pi /2,\theta _r+2\pi /2,\theta _r+3\pi /2\}`$ for $`\theta _r`$ in the expression obtained for node 1. Therefore we will look at how the transformation is done for node 1, and the results for the other nodes will follow. For node 1 we get $`(kq)r`$ $`=`$ $`r_x(k_xq_x)+r_y(k_yq_y)`$ (A34) $`=`$ $`{\displaystyle \frac{r|k^{}q^{}|}{\sqrt{2}}}[\mathrm{cos}\theta _{k^{}q^{}}(\mathrm{cos}\theta _r+\mathrm{sin}\theta _r)`$ (A36) $`+\mathrm{sin}\theta _{k^{}q^{}}(\mathrm{cos}\theta _r+\mathrm{sin}\theta _r)]`$ As a summary we have thus arrived at the following transformations $`ϵ_k`$ $`=`$ $`vk_y^{}`$ (A37) $`\mathrm{\Delta }_k`$ $`=`$ $`vk_x^{}`$ (A38) $`E_k`$ $`=`$ $`vk^{}`$ (A39) $`|kq|`$ $`=`$ $`|k^{}q^{}|`$ (A40) $`(k\times q)^z`$ $`=`$ $`k_F(k^{}q^{})^x`$ (A41) $`(kq)r`$ $`=`$ $`{\displaystyle \frac{r|k^{}q^{}|}{\sqrt{2}}}[\mathrm{cos}\theta _{k^{}q^{}}(\mathrm{cos}\theta _r+\mathrm{sin}\theta _r)`$ (A43) $`+\mathrm{sin}\theta _{k^{}q^{}}(\mathrm{cos}\theta _r+\mathrm{sin}\theta _r)]`$ Dropping the primes, and using these results we can thus linearize the effective interaction around the nodes: $`H`$ $`=`$ $`{\displaystyle \frac{J^2I_1^zI_2^z}{4\pi ^2}}{\displaystyle }d^2k{\displaystyle }d^2qe^{i(kq)r}\times `$ (A45) $`{\displaystyle \frac{k_F^2|kq|^2\mathrm{cos}^2\theta _{kq}}{|kq|^2}}{\displaystyle \frac{vk_xvq_x+vk_yvq_yvkvq}{vkvq(vk+vq)}}`$ We start by integrating over the angles. First we fix the relative angle $`\theta `$ between $`k`$ and $`q`$ and integrate over $`\theta _k`$, thereafter we integrate over $`\theta `$. Integrating over $`\theta _k`$ we get $`H`$ $`=`$ $`{\displaystyle \frac{J^2I_1^zI_2^zk_F^2}{4\pi ^2v}}{\displaystyle }dk{\displaystyle }dq{\displaystyle \frac{kq}{k+q}}{\displaystyle }d\theta (\mathrm{cos}\theta 1)\times `$ (A47) $`\pi \left[J_0(r|kq|)\mathrm{sin}(2\theta _r)J_2(r|kq|)\right]`$ The angular dependence will, however, vanish, since summing up the contributions from the four nodes gives us $`\mathrm{sin}(2\theta _r)+\mathrm{sin}(2(\theta _r+\pi /2))`$ (A48) $`+\mathrm{sin}(2(\theta _r+\pi ))+\mathrm{sin}(2(\theta _r+3\pi /2))]=0,`$ (A49) and this tells us that the effective interaction is isotropic, even though the gap is anisotropic. Integrating out the relative angle we find $`H`$ $`=`$ $`{\displaystyle \frac{2J^2I_1^zI_2^zk_F^2}{v}}{\displaystyle 𝑑k𝑑q}\times `$ (A51) $`{\displaystyle \frac{kq}{k+q}}\left[J_0(kr)J_0(qr)+J_1(kr)J_1(qr)\right]`$ This integral is oscillatory, and we introduce a momentum cut-off $`K`$ and make the integration variables dimensionless by letting $`krk`$ and $`qrq`$: $`H`$ $`=`$ $`{\displaystyle \frac{2J^2I_1^zI_2^zk_F^2}{vr^3}}{\displaystyle _0^{Kr}}𝑑k{\displaystyle _0^{Kr}}𝑑q`$ (A53) $`{\displaystyle \frac{kq}{k+q}}\left[J_0(k)J_0(q)+J_1(k)J_1(q)\right],`$ This integral can be solved numerically, and the behavior for large $`r`$ is given by $$H\frac{2J^2I_1^zI_2^zk_F^2}{vr^3}[0.37+0.33\mathrm{sin}(2Kr)].$$ (A54) This results represents an anti-ferromagnetic $`r^3`$ part that is independent of the cut-off, and an oscillating $`\mathrm{sin}(2Kr)r^3`$ part. Due to the relative sizes of the two terms the interaction is always positive and hence completely anti-ferromagnetic. The $`SI`$ interaction will lead to a very similar result, differing only in the prefactor. This can be seen, since the result, after performing the first angular integral will be given by $`H`$ $`=`$ $`{\displaystyle \frac{J^2I_1I_2}{4\pi ^3v}}{\displaystyle }dk{\displaystyle }dq{\displaystyle \frac{kq}{k+q}}\times `$ (A56) $`{\displaystyle 𝑑\theta (\mathrm{cos}\theta 1)\pi J_0(r|kq|)}`$ and the only difference compared to the $`LI`$ interaction is the prefactor. The final result for the $`SI`$ coupling will be $$H^{SI}\frac{J^2I_1I_2}{2\pi vr^3}[0.37+0.33\mathrm{sin}(2Kr)]$$ (A57) So the two electron-spin couplings lead to the same functional form of the RKKY interaction. We have therefore showed that both the $`SI`$ and $`LI`$ interactions give rise to an effective anti-ferromagnetic model. We have used the nodal approximation and assumed that the superconductor is isotropic $`v_F=v_\mathrm{\Delta }`$, and these approximations may change the final result somewhat, but it appears unlikely that they would make the effective spin-spin interactions predominantly ferromagnetic, and therefore the randomly distributed spins will form a spin-glass phase, as discussed in the main part of the paper.
warning/0004/cond-mat0004268.html
ar5iv
text
# Micro-canonical Statistical Mechanics of some Non-Extensive Systems *footnote **footnote *Invited talk at the International Workshop on Classical and Quantum Complexity and Non-extensive Thermodynamics, Denton-Texas, April 3-6, 2000 ## I Introduction This conference is addressed to the extension of thermo-statistics to non-extensive systems. This is a new realm of thermo-statistics which came into focus by the pioneering work of Tsallis . Non-extensive systems are defined by the following property: If they are divided into pieces, their energy and entropy is not the sum of the energies and entropies of their parts in contrast to conventional extensive systems where this is assumed at least if the pieces are themselves macroscopic. This is the case if the forces in the systems have a long range comparable with or larger than the linear dimensions of the system like for nuclei, atomic clusters and astrophysical objects. However, also inhomogeneous systems, e.g. systems with separated phases are non-extensive. Although the largest possible systems like clusters of galaxies belong to this group I call these systems “small” to stress the fact that the thermodynamic limit either does not exist or makes no sense. For systems with short range forces does the entropy of the surfaces separating the different phases not scale with the volume of the system. The entropy per particle $`s=S(E)/N`$ shows a convex intruder with a depth $`N^{1/3}`$. As long as one cares about this non-concavity also these systems are to be considered as non-extensive. Elliot Lieb, Boltzmann laureate from 1998, claims “ Extensivity is essential for thermodynamics to work !” This is certainly true for the original statistical foundation of thermodynamics by Gibbs . For the extension of thermo-statistics to non-extensive, “small” systems one should, however, remember that the original formulation by Boltzmann, even though he did presumably not think of non-extensive systems, does not rely on the use of the thermodynamic limit nor any assumption of extensivity and concavity of the entropy, see below. Hence, before introducing any major deviation from standard equilibrium statistics one should explore its original and fundamental Boltzmann, or micro-canonical, form. This is what I will do in the following and demonstrate that Lieb’s claim is wrong and contradicts to Boltzmann’s view of thermodynamics. Entropy as defined by Boltzmann does not invoke the thermodynamic limit and, consequently, does not demand extensivity. Moreover, it will even turn out that the non-extensivity of inhomogeneous systems with separated phases gives just a clou to illuminate the physics of phase transitions explicitly and sharply. There is a huge world of non-extensive systems which can only be described by Boltzmann’s micro-canonical ensemble written in its most condensed form on the epitaph of his gravestone which covers all equilibrium thermodynamics: $`𝑺\mathbf{=}𝒌\mathbf{}𝒍𝒏𝑾`$ where $`W(E,N,V)`$ $`=`$ $`ϵ_0tr\delta (EH_N)`$ (1) $`tr\delta (EH_N)`$ $`=`$ $`{\displaystyle \frac{d^{3N}pd^{3N}q}{N!(2\pi \mathrm{})^{3N}}\delta (EH_N)}.`$ (2) is the volume of the $`3N1`$ dimensional manifold at given sharp energy, the micro-canonical ensemble. ($`ϵ_0`$ is a suitable small energy constant to make the $`W`$ dimensionless.) In his famous book “Elementary Principles in Statistical Physics” Gibbs deduced the canonical ensemble from the fundamental micro-canonical. He showed that the canonical one becomes equivalent to the micro-canonical ensemble in the thermodynamic limit if the system is homogeneous. Otherwise, the canonical and grand-canonical are not correct. On page 75, chapter VII of he gives explicitly the example of the separation of the liquid and gas phase for which the canonical fluctuations of the energy per particle do not vanish even in the thermodynamic limit and the canonical ensemble, and with it the “Boltzmann-Gibbs” distribution loose their validity. The reason for the special fundamental role of the micro ensemble comes of course from the fact that the internal dynamics of a many-body system conserves energy and does not mix different energy shells. In an open system embedded in a heat bath the mechanism of energy violation operates via the surface between the system and the bath and is of the same order in the particle number $`N^{2/3}`$ as any other internal surface energies, which are also to be ignored in the canonical treatment. Boltzmann defines the entropy in eq.(1) as a measure of the mechanical N-body phase space. Thermodynamics has thus a geometrical interpretation and can be read off from the topology of $`W(E,N,\mathrm{})`$, the volume of its constant energy manifold. No probabilistic interpretation must be invoked like $$S=\underset{i}{}p_i\mathrm{ln}p_i,$$ (3) where $`p_i`$ is the probability to find the N-particle configuration $`i`$ in the ensemble. This form of entropy can be investigated for any finite even “small” system without any reference to the thermodynamic limit. I will show here that all thermodynamical features of such a system including the whole “zoo” of phase transitions and critical phenomena can be read off from this topology. This proves that in contrast to what is written in most textbooks of statistical mechanics phase transitions do exist and can be sharply defined in finite even “small” and non-extensive systems. Conventional thermo-statistics, however, relies heavily on the use of the thermodynamic limit ($`V\mathrm{}|_{N/V\text{, or }\nu \text{ const.}}`$) and extensivity, c.f. e.g. the book of Pathria . This is certainly not allowed for our systems. That the micro-canonical statistics works well also for “small” systems without invoking extensivity will be demonstrated here for finite normal systems which are also non-extensive at phase transitions of first order. The use of the thermodynamic limit and of extensivity, however, is closely intervowen with the development of thermodynamics and statistical mechanics since its beginning more than hundred years ago. When we extend thermodynamics to “small” systems we should establish the formalism of thermodynamics starting from mechanics in order to remain on a firm basis. We will see how this idea guides us to more and deeper insight into the most dramatic phenomena of thermodynamics, phase transitions. Moreover, it gives the most natural extension of thermo-statistics to many non-extensive systems without invoking any modification of the entropy like that proposed by Tsallis . This discussion may further help to illuminate the domain of physical situations where the Tsallis formalism is relevant: systems that do not populate the energy manifold of phase space densely perhaps in a fractal way, perhaps at the edge of chaos c.f.. I will sketch a deduction of thermo-statistics from the principles of mechanics alone. Nothing outside of mechanics must be invoked. This was the starting point of Boltzmann , Gibbs , Einstein and the Ehrenfests at the beginning of the last century. They all agreed on the logical hierarchy of the micro-canonical as the most fundamental ensemble from which the canonical, and grand-canonical ensembles can be deduced under certain conditions. According to Gibbs the latter two approximate the micro ensemble in the thermodynamic limit of infinitely many particles interacting by short range interactions if the system is homogeneous. Then surface effects and fluctuations can be ignored relatively to the bulk mean values. This is the main reason why the thermodynamic limit became basic in the statistical foundation of macroscopic thermodynamics. However, it was Gibbs who stressed that the equivalence of the three ensembles is not true at phase transitions of first order, even in the thermodynamic limit. The link between the micro and the grand ensemble is established by the double Laplace transform: $$Z(T,\nu ,V)=\text{ }_0^{\mathrm{}}\frac{dE}{ϵ_0}𝑑Ne^{[E/T+\nu NS(E)]}=\frac{V^2}{ϵ_0}\text{ }_0^{\mathrm{}}𝑑e𝑑ne^{V[e/T+\nu ns(e,n)]}$$ (4) Globally $`s(e,n)=S(e=E/V,n=N/V)/V`$ is concave (downwards bended). If $`s(e,n)`$ is also locally concave then there is a single point $`e_s`$,$`n_s`$ for given $`T,\nu `$ as shown in figure (1) with $`{\displaystyle \frac{1}{T}}=\beta `$ $`=`$ $`{\displaystyle \frac{S}{E}}|_s`$ (5) $`{\displaystyle \frac{P}{T}}`$ $`=`$ $`{\displaystyle \frac{S}{V}}|_s`$ (6) $`\nu ={\displaystyle \frac{\mu }{T}}`$ $`=`$ $`{\displaystyle \frac{S}{N}}|_s,`$ (7) and a one to one mapping is generated by the Laplace transform eq.(4) from the micro variables $`E,N`$ to the grand canonical $`T,\nu `$. This is illustrated by figure (1) for the case of a single conserved variable $`E`$. In Gaussian approximation $`s(e)`$ remains under the tangent line and in the thermodynamic limit ($`V\mathrm{}`$) only the immediate neighborhood of the stationary point contributes. The importance of the curvature of $`s(e)`$ for the bijective mapping is evident. An equilibrated many-body system is characterized by few macroscopic quantities: 1. Its energy $`E`$, mass (number of atoms) $`N`$, volume $`V`$, 2. its entropy $`S`$ eq.(1), 3. its temperature $`T`$ eq.(5), pressure $`P`$ eq.(6), and chemical potential $`\nu `$ eq.(7). There are important qualitative differences between these three groups: All variables of the first group have a clear mechanical significance. They are conserved and well defined at each point of the N-body phase space. The internal dynamics of the system cannot leave the shell in phase space which is defined by these variables. Also entropy as the most important quantity within thermodynamics has with eq.(1) a clear mechanical foundation since Boltzmann. The set of points on this surface defines the micro-canonical ensemble. In contrast to the conserved quantities which are defined at each phase space point, entropy refers to the whole micro-canonical ensemble. Remark: For a system with discrete energies $`E_i`$ with some degeneration $`n(E_i)`$ e.g. a lattice or a quantum system one should define the micro-canonical partition sum by the number $`W=n(E_i)`$ of states at this energy. When we discuss derivatives of $`W`$ we imagine a suitable smoothing of this. It is important to notice that Boltzmann’s and also Einstein’s formulation allows for defining the entropy entirely within mechanics by $`S_{micro}:=ln[W(E,N,V)]`$. It is a single valued, non-singular, in the classical case multiply differentiable function of all “extensive”, conserved dynamical variables. No thermodynamic limit must be invoked and this definition applies to non-extensive like our “small” systems as well. The third group of quantities which characterize the thermodynamical state of an equilibrated many-body system, temperature $`T`$, pressure $`P`$ and chemical potential $`\nu `$ have no immediate mechanical significance. From the mechanical point of view they are secondary, derived quantities. This difference to the two other groups of variables will turn out to be significant for “small” systems. Again, like entropy itself, these quantities refer to the whole micro-canonical ensemble, not to an individual point in the N-body phase space. Starting from this point, the conventional thermo-statistics assumes extensivity and explores the thermodynamic limit ($`V\mathrm{}|_{\text{N/V, or }\nu \text{ const.}}`$) c.f. . This procedure follows Gibbs . He introduced the canonical ensemble, which since then became the basic of all modern thermo-statistics. ## II Phase transitions micro-canonically This talk addresses phase transitions in “small” systems. Conventionally phase transitions are thought to exist only in the thermodynamic limit ($`V\mathrm{}|_{\text{N/V, or }\nu \text{ const.}}`$). Yang and Lee define them by the singularities of the grand-canonical partition sum as function of the fugacity $`z=e^\nu `$. As the partition sum $`Z(T,\nu )`$ is analytical in $`z`$ for finite volumes $`V`$ the singularities can occur in the thermodynamic limit only ($`V\mathrm{}|_{e,n}`$). I will show how these singularities arise from points in the parameter space {$`e,n,\mathrm{}`$} where the micro-canonical (Boltzmann) entropy $`s(e,n)`$ has either vanishing or even positive curvature. These are the catastrophes of the Laplace transform from the micro-canonical to the grand-canonical partition sum (4). At phase transitions the inter-phase surface does not scale with the volume. Systems with phase separation are non-extensive. These configurations become exponentially suppressed in the canonical ensemble . Moreover, Schrödinger thought that Boltzmann’s entropy is not usefull to describe systems other than gases . However, today with the powerfull and cheap computers we can explore the micro-canonical ensemble in realistic situation. The clarification the basic role of Boltzmann’s statistics in the most dramatic situations of equilibrium thermodynamics is demanding. To realize the powerful application of Boltzmann’s \[not Boltzmann-Gibbs (!)\] statistics to non-extensive, finite systems having a non-concave entropy $`s(e,n,\mathrm{})`$ may also specify the cases where one has to go beyond and where a generalization like Tsallis entropy is needed. When there are no singularities in the partition sum $`Z(T,\nu )`$ for finite systems, are there no phase transitions in finite systems? There are phenomena observed in finite systems which are typical for phase transitions. Sometimes this is even so in astonishingly small systems like nuclei and atomic clusters of $`100`$ atoms . In chapter (IV) and in reference we show that their characteristic parameters as transition temperature, latent heat, and surface tension are – in the case of some metals – already for thousand atoms close, though of course not equal, to their known bulk values. Therefore, it seems to be fully justified to speak in these cases of phase transitions of first order. We need an extension of thermodynamics to “small” systems which avoids the thermodynamic limit. However, here is a severe problem: The non-equivalence of the three popular ensembles, the micro-canonical, the canonical, and the grand-canonical ensembles for “small” systems. The energy per particle fluctuates around its mean value $`<E/N>`$ in the (grand-)canonical ensemble whereas the energy fluctuations are zero in the micro-canonical ensemble. Moreover, the heat capacity is strictly positive in the canonical ensembles whereas it may become negative in the micro ensemble. To extend themodynamics to “small” systems it is certainly advisable to keep close contact with mechanics. It is helpful to realize that the fundamental micro-canonical ensemble as introduced by Boltzmann is the only one which has a clear mechanical definition for finite systems. To extend the definition of phase transitions of Yang and Lee to finite systems we must study which feature of the micro-canonical partition sum $`W(E,N,V)`$ leads to singularities of the grand-canonical potentials $`\frac{1}{V}ln[Z]`$ as function of $`z=e^\nu `$ by the Laplace transform eq.(4). In the thermodynamic limit $`V\mathrm{}|_{\nu \text{ const.}}`$ this integral can be evaluated by asymptotic methods as discussed above. If there is only a single stationary point then there is a one to one mapping of the grand-canonical ensemble to the micro-canonical one and energy-fluctuations disappear $`1/\sqrt{N}`$. This, however, is not the case at phase transitions of first order. Here the grand-canonical ensemble contains several Gibbs states (stationary points, c.f. the figure (6)) at the same temperature and chemical potential which contribute similarly to the integral eq.(4). The statistical fluctuations of $`e`$ and $`n`$ do not disappear in the grand-canonical ensemble even in the thermodynamic limit. Between the stationary points $`s(e,n)`$ has at least one principal curvature $`0`$. Here van Hove’s concavity condition for the entropy $`s(e,n)`$ is violated. In the thermodynamic limit these points get jumped over by the integral (4) and $`ln[Z]`$ becomes non-analytic. Consequently, we define phase transitions also for finite systems topologically by the points and regions of non-negative curvature of the entropy surface $`s(e,n)`$ as a function of the mechanical, conserved “extensive” quantities like energy, mass, angular momentum etc.. The central quantity of our further discussion, the determinant of the curvatures of $`s(e,n)=S(e=E/V,n=N/V)/V`$ is defined as $$d(e,n)=\begin{array}{cc}\frac{^2s}{e^2}& \frac{^2s}{ne}\\ \frac{^2s}{en}& \frac{^2s}{n^2}\end{array}=\begin{array}{cc}s_{ee}& s_{en}\\ s_{ne}& s_{nn}\end{array}=\lambda _1\lambda _2.$$ (8) The two curvature eigenvalues (main curvatures) are assumed to be ordered and $`\lambda _1>\lambda _2<0`$. Also critical fluctuations, i.e. abnormally large fluctuations of some extensive variable in the grand-canonical ensemble or the eventual divergence of some susceptibilities are micro-canonically connected to the vanishing of the curvature determinant, e.g. in the following examples of $`d(e,n)`$ or $`d(e,m)`$ respectively: The micro-canonical specific heat is given by : (9) $`c_{micro}(e,n)`$ $`=`$ $`{\displaystyle \frac{e}{T}}|_\nu ={\displaystyle \frac{s_{nn}}{T^2d(e,n)}},`$ (10) $`d`$ $`=`$ $`{\displaystyle \frac{d(\beta \nu )}{d(en)}}`$ (12) or the isothermal magnetic susceptibility by : $`\chi _{micro,T}(e,m)`$ $`=`$ $`{\displaystyle \frac{m}{B}}|_T={\displaystyle \frac{s_{ee}}{d(e,m)}},`$ (14) $`\text{with }s_{ee}={\displaystyle \frac{^2s}{ee}}\text{ etc.}`$ In the case of a classical continuous system $`s(e,n)`$ is everywhere finite and multiply differentiable. In that case the inverse susceptibilities $`[c_{micro}(e,n,V)]^1`$ and $`[\chi _{micro,T}(e,m,V)]^1`$ are well behaved smooth functions of their arguments even at phase transitions. Problems arise only if the susceptibilities are considered as functions of the “intensive” variables $`T`$, and $`\nu `$ or $`BS/m`$ . In the case of lattice systems we can only assume that the inverse susceptibilities are similarly well behaved. Experimentally one identifies phase transitions of first order of course not by the non-analyticities of $`\frac{1}{V}ln[Z]`$ but by the interfaces separating coexisting phases, e.g. liquid and gas, i.e. by the inhomogeneities of the system which become suppressed in the thermodynamic limit in the grand-canonical ensemble. This fact was early realized by Gibbs and he emphasized that using $`S`$ vs. volume, or density, at phase separation “has a substantial advantage over any other method (e.g. pressure) because it shows the region of simultaneous coexistence of the vapor, liquid, and solid phases of a substance, a region which reduces to a point in the more usual pressure-temperature plane.” That is also the reason why for the grand-canonical ensemble the more mathematical definition of phase transitions is needed. The main advantage of the micro-canonical ensemble is that it allows for inhomogeneities as well and thus we can keep much closer to the experimental criteria for finding phase transitions. Interfaces have three opposing effects on the entropy: * An entropic gain by putting a part ($`N_1`$) of the system from the majority phase (e.g. solid) into the minority phase (bubbles, e.g. gas) with a higher entropy per particle. However, this has to be paid by additional energy $`\mathrm{\Delta }E`$ to break the bonds in the “gas”-phase. As both effects are proportional to the number of particles $`N_1`$ being converted, this part of the entropy rises linearly with the additional energy. * With rising size of the bubbles their surfaces grow. This is connected to an entropic loss due to additional correlations between the particles at the interface(surface entropy) proportional to the interface area. As the number of surface atoms is $`N_1^{2/3}`$ this is not linear in $`\mathrm{\Delta }E`$ and leads to a convex intruder in $`S(E,N,V)`$, the origin of surface tension . This is also the reason why systems with phase separation are non-extensive c.f. chapter (VI). * An additional mixing entropy for distributing the $`N_1`$-particles in various ways over the bubbles. At a (multi-) critical point two (or more) phases become indistinguishable because the interface entropy (surface tension) and with it the inhomogeneity (interface) disappears. In order to demonstrate this we investigate in the following the 3-states diluted Potts model now on a finite 2-dim (here $`L^2=50^2`$) lattice with periodic boundaries in order to minimize effects of the external surfaces of the system. The model is defined by the Hamiltonian: $`H`$ $`=`$ $`{\displaystyle \underset{i,j}{\overset{n.n.pairs}{}}}o_io_j\delta _{\sigma _i,\sigma _j}`$ (15) $`n`$ $`=`$ $`L^2N=L^2{\displaystyle \underset{i}{}}o_i.`$ (16) Each lattice site $`i`$ is either occupied by a particle with spin $`\sigma _i=1,2,\text{ or }3`$ or empty (vacancy). The sum is over neighboring lattice sites $`i,j`$, and the occupation numbers are: $$o_i=\{\begin{array}{cc}1& \text{, spin particle in site }i\hfill \\ 0& \text{, vacancy in site }i\hfill \end{array}.$$ (17) This model is an extension of the ordinary ($`q=3`$)-Potts model to allow also for vacancies. At zero concentration of vacancies ($`n=1`$), the system has in the limit of infinite volume $`V`$ a continuous phase transition at $`e_c=1+\frac{1}{\sqrt{q}}1.58`$. With rising number of vacancies the probability to find a pair of particles at neighboring sites with the same spin orientation decreases. The inclusion of vacancies has the effect of an increasing effective $`q_{eff}3`$. This results in an increase of the critical energy of the continuous phase transition with decreasing $`n`$ and provides a line of continuous transition, which is supposed to terminate when $`q_{eff}`$ becomes larger than $`4`$, where the transition becomes first order. At smaller energies the system is in one of three ordered phases (spins predominantly parallel in one of the three possible directions). We call this the “solid” phase. This scenario gets full support by our numerical findings. In figure (2) the determinant of curvatures of $`s(e,n)`$: $$d(e,n)=\begin{array}{cc}\frac{^2s}{e^2}& \frac{^2s}{ne}\\ \frac{^2s}{en}& \frac{^2s}{n^2}\end{array}=\begin{array}{cc}s_{ee}& s_{en}\\ s_{ne}& s_{nn}\end{array}=\lambda _1\lambda _2$$ (18) is shown. On the diagonal we have the ground-state of the $`2`$-dim Potts lattice-gas with $`e_0=2n`$, the upper-right end is the complete random configuration (here without contour lines), with the maximum allowed excitation $`e_{rand}=\frac{2n^2}{q}`$ and the maximum possible entropy. In the region above the line $`\widehat{CP_mB}`$ we have the disordered, “gas”. Here the entropy $`s(e,n)`$ is concave ($`d>0`$), both curvatures are negative (we have always the smaller one $`\lambda _2<0`$). This is also the case inside the triangle $`A`$$`P_m`$$`C`$ (ordered, “solid” phase). In these regions the Laplace integral eq.(4) has a single stationary point. They correspond to pure phases. FIG. 4.: Cut through the determinant $`d(e,n)`$ along the line shown in figure (2) at const. $`n=0.57`$, slightly below the multi-critical region. There are several zero points of the determinant of curvatures: The left one is simultaneously a maximum with $`\mathbf{}𝒅\mathbf{=}\mathrm{𝟎}`$ and consequently critical as discussed above Below $`\widehat{AP_mB}`$ $`s(e,n)`$ is convex ($`d<0`$) corresponding to phase-separation, first order. At these {$`e,n`$} the Laplace integral (4) has no stationary point. Here we have a separation into coexisting phases, e.g. solid and gas. Due to the inter-phase surface tension or the negative contribution to the entropy by the additional correlations at the phase boundaries (surfaces), $`s(e,n)`$ has a convex intruder with positive largest curvature. In it is shown that the depth of the convex intruder in $`s(e,n)`$ gives the surface tension, c.f. chapter (IV). At the dark lines like $`\widehat{AP_mB}`$ we have $`d(e,n)=0`$. These are the termination lines of the first order transition. At these lines one of the two phases is depleted and beyond all particles are in the other phase (solid or gas respectively). Along the medium dark lines like $`\widehat{P_mC}`$ we have $`\text{v}_1\mathbf{}d=0`$, here the curvature determinant has a minimum in the direction of the largest curvature eigenvector $`\text{v}_1`$. The line $`\widehat{P_mC}`$ towards the critical point of the ordinary ($`q=3`$)-Potts model at $`e=1.58`$, $`n=1`$ correponds to a critical line of second order transition which terminates at the multicritical “point” $`P_m`$. It is a deep valley in $`d(e,n)`$ c.f. fig.3 which rises slightly up towards $`C`$. On the level of the present simulation we cannot decide whether this rise is due to our still finite, though otherwise sufficient, precision or is a general feature of finite size. (The largest curvature $`\lambda _1`$ of $`s(e,n)`$ has a local maximum with $`\lambda _1\stackrel{<}{}0`$, or $`d\stackrel{>}{}0`$). Because of our finite interpolation width of $`\mathrm{\Delta }e\pm 0.04`$, $`\mathrm{\Delta }n\pm 0.02`$ it might be that this valley of $`d(e,n)`$ gets a little bit filled up from its sides and the minimum of $`d(e,n)`$ is rounded, c.f. fig.3. The valley converts below the crossing point $`P_m`$ into a flat ridge inside the convex intruder of the first order lattice-gas phase-separation region see e.g. fig.(4). In the cross-region (light grey in fig.2) we have: $`d=0\mathbf{}d=\mathrm{𝟎}`$. This is the locus of the multi-critical point $`P_m`$ where the large curvature $`\lambda _10`$ in a two-dimensional neighborhood. Here the curvature determinant $`d(e,n)`$ is flat up to at least second order in both directions $`\mathrm{\Delta }e`$ and $`\mathrm{\Delta }n`$ and $`s(e,n)`$ is cylindrical. It is at $`e_m1`$, $`n_m0.6`$ or $`\beta _m=1.48\pm 0.03`$, $`\nu _m=2.67\pm 0.02`$. Naturally, $`P_m`$ spans a much broader region in {$`e,n`$} than in {$`\beta ,\nu `$}, remember $`d(e,n)`$ is flat near $`P_m`$. This situation reminds very much the well known phase diagram of a <sup>3</sup>He –<sup>4</sup>He mixture in temperature vs. mole fraction of <sup>3</sup>He c.f. fig.3. in ref.. ## III On the topology of curvatures The two eigenvalues of the curvature matrix (18) are: (19) $`\lambda _{1,2}`$ $`=`$ $`{\displaystyle \frac{s_{ee}+s_{nn}}{2}}\pm {\displaystyle \frac{1}{2}}\sqrt{(s_{ee}+s_{nn})^24d}`$ (21) and the corresponding eigenvectors are : $`\text{v}_\lambda `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{(s_{ee}\lambda )^2+s_{en}^2}}}\left(\begin{array}{c}s_{en}\hfill \\ s_{ee}\lambda \hfill \end{array}\right).`$ (24) At critical points the following conditions hold: $`d`$ $`=`$ $`{\displaystyle \frac{(\beta \nu )}{(en)}}=L^2D=0`$ (26) $`s_{ee}s_{nn}`$ $`=`$ $`s_{en}^2.`$ (29) Here the directions $`\beta =`$const. and $`\nu =`$ const. are parallel, the Jacobian vanishes and we have : $`{\displaystyle \frac{\beta }{e}}|_\nu `$ $`=`$ $`{\displaystyle \frac{d}{s_{nn}}}=0`$ (30) $`{\displaystyle \frac{\nu }{n}}|_\beta `$ $`=`$ $`{\displaystyle \frac{d}{s_{ee}}}=0.`$ (31) $`\lambda _1`$ $`=`$ $`0`$ (32) $`\lambda _2`$ $`=`$ $`s_{ee}+s_{nn}`$ (33) $`\text{v}_{\lambda =0}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{s_{ee}^2+s_{en}^2}}}\left(\begin{array}{c}s_{en}\hfill \\ s_{ee}\hfill \end{array}\right)`$ (36) $`\text{v}_{\lambda 0}`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{s_{nn}^2+s_{en}^2}}}\left(\begin{array}{c}s_{en}\hfill \\ s_{nn}\hfill \end{array}\right).`$ (39) The vanishing of $`d`$ alone is not sufficient for criticality. Physically, it means that the surface entropy (tension) and with it the interface separating coexistent phases disappears. This, however, can also signalize a depletion of one of the two phases in favor of the other. At a critical end-point, however, the interface disappears at a non vanishing number of atoms in each of the two phases. I.e. in an infinitesimal neighborhood of a critical point, $`d`$ must remain zero. In a topologically formulation a critical end-point of first order transition is at: $`d`$ $`=`$ $`0`$ (40) and (41) $`\text{v}_1\mathbf{}d`$ $`=`$ $`0,`$ (42) whereas at a multi-critical point we have $`\mathbf{}𝒅\mathbf{=}\mathrm{𝟎}`$. This is a generalization of the well known condition for a continuous transition in one dimension: the simultaneous vanishing of $`\beta ^{}(e)=0`$ and of the curvature of $`\beta (e)`$, $`\beta ^{\prime \prime }(e)=0`$. Figure (5) shows a map of some trajectories which follow the eigen-vector $`𝒗_\mathrm{𝟏}`$ with the largest curvature eigen-value $`\lambda _1`$. In the region of the convex intruder ($`\lambda _1>0`$) i.e. the region of phase-separation $`𝒗_\mathrm{𝟏}`$ is $``$ parallel to the ground state $`e=2n`$. Also the lines of $`\beta =`$const. and $`\nu =`$const. follow approximately this direction. Their Jacobian $`(\beta \nu )/(en)=d(e,n)`$ is negative but small. This reminds of the situation in the thermodynamic limit where this region of phase separation is flat ($`d(e,n)`$) or $`s(e,n)`$ cylindrical, both intensive variables are constant and the Jacobian $`d0`$. One can also see in fig.5 how the direction of the largest curvature $`𝒗_\mathrm{𝟏}`$ turns into the $`e`$-direction when one approaches the critical point $`C`$ of the ordinary ($`q=3`$)-Potts model at $`n=1`$. At $`n=1`$ we know that for an infinite system the ordinary ($`n=1`$) three state Potts model has a second order transition at $`e=1.58`$ where the curvature of $`s(e)`$ vanishes, $`s_{ee}=0`$. I.e. the component $`\text{v}_1\mathbf{}d`$ of $`\mathbf{}d`$ indicates nicely the locus of the second order “temperature driven” transition of the ordinary Potts model. This chapter was an overview on the power and extreme rich insight that the topology of curvatures of Boltzmann’s micro-canonical entropy $`s(e,n)`$ can give in a generic case of a “small” and non-extensive system. ## IV What is the physics behind a positive curvature? It is linked to the inter-phase surface tension. This is shown in a simulation of $`1000`$ sodium atoms at constant external pressure $`P=1`$atm. In figure (6) I show the micro-canonical entropy $`s(e)`$ for $`1000`$ sodium atoms with realistic interactions. Details of this calculation are given in . Here only a few remarks: The calculations were done at a constant volume $`V(E)`$ which was chosen for the whole ensemble at the given energy so that the pressure $`P=\frac{S}{V}/\frac{S}{E}`$ is given to be $`1`$ atm. The important and characteristic difference to Andersen’s constant pressure ensemble should be noticed! This is discussed in my book that will be published soon . The table (I) gives the 4 characteristic parameters classifying the transition of $`N=200\mathrm{}3000`$ Na-atoms at external pressure of $`1`$ atm. $`T_{tr}`$ is the transition temperature ($`T=(S/E)^1`$) in Kelvin, $`q_{lat}=e_3e_1`$ is the the latent energy per atom, $`s_{boil}`$ is the entropy gain of a single atom when converted from the liquid phase into the gas phase. $`\mathrm{\Delta }s_{surf}`$ is the entropy-loss per atom due to surface correlations, $`N_{eff}`$ is the average number of surface atoms of all coexisting clusters, and $`\sigma `$ is the surface tension per surface atom. These values are compared to their corresponding values of bulk sodium. ## V The information lost in the grand-canonical ensemble In this chapter I explain how and which part of the micro-canonical phase diagram is lost in the conventional canonical treatment. Figure (7) explains what happens if one plots the entropy $`s`$ vs. the “intensive” quantities $`\beta =S/E`$ and $`\nu =S/N`$ as one would do for the grand-canonical ensemble: As there are several points $`E_i,N_i`$ with identical $`\beta ,\nu `$, $`s_{micro}(\beta ,\nu )`$ is a multivalued function of $`\beta ,\nu `$. Here the entropy surface $`s_{micro}(e,n)`$ is folded onto itself. In the projection in fig.7, these points show up as a black critical line (dense region). Here this black line continues over the multi-critical point $`P_m`$ towards $`C`$ indicating the direction to the critical point of the ordinary $`q=3`$ Potts model at $`n=1`$ (zero vacancies). Between $`P_m`$ and $`C`$ the slopes $`{\displaystyle \frac{s}{\beta }}|_\nu `$ $`=`$ $`{\displaystyle \frac{1}{d}}[\beta s_{nn}\nu s_{ne}]`$ (43) or (44) $`{\displaystyle \frac{s}{\nu }}|_\beta `$ $`=`$ $`{\displaystyle \frac{1}{d}}[\beta s_{en}\nu s_{ee}]`$ (45) are negative large but finite. The information given by the projection would be all information which can be obtained from the conventional grand-canonical entropy $`s(T,\nu ,V)`$, if we would have calculated it from the Laplace transform, eq.(4). The shaded region will be lost. The upper part of figure (7) shows $`s_{micro}(\beta ,\nu )`$ in a three dimensional plot. The lines building the entropy surface are lines of equal $`\beta `$. The images of the points $`A,D,B,C`$ defined in fig.2 are roughly indicated. The back folded branches, the convex intruder of $`s(e,n)`$ between the lines $`\widehat{AP_mB}`$ and $`\widehat{ADB}`$, the region of phase separation, can here be seen from the side (shadowed). It is jumped over in eq. (4) and gets consequently lost in $`Z(T,\nu )`$. This demonstrates the far more detailed insight into phase transitions and critical phenomena obtainable by micro-canonical thermo-statistics not accessible to the canonical treatment, c.f. the similar arguments of Gibbs . In the next two figures the cross-section through $`s(\beta ,\nu )`$ at constant $`\beta `$ along the bold lines from fig.(7) is shown in figure (8) below the multi-critical point $`\beta _m=1.48`$ and in figure (9) above it. The latter clearly shows the back-bending of $`s(\beta ,\nu )`$. ## VI Convex entropy — Violation of the Second Law ? At this point it is worth-wile to spend some words on a popular misunderstanding connected with the eventual convexity of the entropy as function of “extensive” quantities like the energy: The convex parts of $`S(E,N)`$ violate van Hove’s concavity condition . One may believe that this is also a contradiction to the second law of thermodynamics: At a convex region of $`S(E,N)`$ a split of the system into two pieces with entropies $`S_1(E_1,N_1)`$ and $`S_2(E_2,N_2)`$ would have $`S_1(E_1,N_1)+S_2(E_2,N_2)>S(E_1+E_2,N_1+N_2)`$. So the system seems to gain entropy by splitting. This, however, is an error. The Boltzmann entropy as defined in eq.(1) is already the logarithm of the sum over all possible configurations of the system at the given energy. The split ones are a subset of these. Their partial phase space $`W_{split}`$ is of course $``$ the total $`W`$. The entropy $`S_{split}=ln(W_{split})`$ is $``$ the total entropy. Evidently, the split system looses some surface entropy $`S_{surf}`$ at the separation boundary due to additional correlations imposed on the particles at the boundary. The entropy after split is consequently: $`S_{split}`$ $`=`$ $`S_1(E_1,N_1)+S_2(E_2,N_2)S_{surf}`$ (46) $``$ $`S(E_1+E_2,N_1+N_2),`$ (47) It is a typical finite size effect. $`S_{surf}/V`$ vanishes in the limit $`V\mathrm{}`$ for interactions with finite range. The entropy is non-extensive for finite systems but becomes extensive in the limit, and van Hove’s theorem is fulfilled. This is of course only under the condition that $`lim_V\mathrm{}S_{surf}/V=0`$. So in the case of a self-gravitating system the convex intruder and the negative specific heat will not disappear . In general this is of course a trivial conclusion: An additional constraint like an artificial cut of the system can only reduce phase space and entropy. The Second Law is automatically satisfied in the Boltzmann formalism whether $`S`$ is concave or not, whether $`S`$ is “extensive” or not. A positive (wrong) curvature introduces problems to the geometrical interpretation of thermodynamics as formulated by Weinhold which relies on the non-convexity of $`S(E,N)`$. Weinhold introduces a metric like $`g_{ik}`$ $`=`$ $`{\displaystyle \frac{^2S}{X^iX^k}}`$ (49) where we identify : $`X^1`$ $`=`$ $`E`$ (50) $`X^2`$ $`=`$ $`N.`$ (52) The thermodynamic distance is defined as : $`\mathrm{\Delta }_{a,b}`$ $`=`$ $`\sqrt{[X^i(a)X^i(b)]g_{ik}[X^k(a)X^k(b)]}.`$ (53) Evidently, a negative metric $`g_{ik}`$ is here not allowed. Of course Weinhold’s theory does not apply to finite systems with phase transitions. ## VII Conclusion Micro-canonical thermo-statistics describes how the entropy $`s(e,n)`$ as defined entirely in mechanical terms by Boltzmann depends on the conserved “extensive” variables: energy $`e`$, particle number $`n`$, angular momentum $`L`$ etc. It is well defined for finite systems without invoking the thermodynamic limit. Thus in contrast to the conventional theory, we can study phase transitions also in “small” systems or other non-extensive systems. In this simulation we could classify phase transitions in a “small” system by the topological properties of the determinant of curvatures $`d(e,n)`$, eq.(18) of the micro-canonical entropy-surface $`s(e,n)`$: In the micro-canonical ensemble of a “small”, non-extensive system, phase transitions are classified unambiguously by the following topology of the curvature determinant: * A single stable phase by $`d(e,n)>0`$ ($`\lambda _1<0`$). Here $`S(E,N)`$ is concave in both directions. Then there is a one to one mapping of canonical$``$micro-ensemble. Then the last two terms in $$\frac{F(T,\mu ,V)}{V}e_s\mu n_sTs_s+T\frac{\mathrm{ln}(\sqrt{d(e_s,n_s)})}{V}+o(\frac{\mathrm{ln}V}{V})$$ (54) can be neglected for large volume. * A transition of first order with phase separation and surface tension is indicated by $`d(e,n)<0`$ ($`\lambda _1>0`$). $`S(E,N)`$ has a convex intruder in the direction $`𝒗_{𝝀_\mathrm{𝟏}}`$ of the largest curvature. Because of $`d0`$ the second last term in eq.(54) is complex or diverges. The whole convex area of {e,n} is mapped into a single point in the canonical ensemble. I.e. if the curvature of $`s(e,n)`$ is $`\lambda _10`$ both ensembles are not equivalent. * A continuous (“second order”) transition with vanishing surface tension, where two neighboring phases become indistinguishable, is indicated by lines (critical) with $`d(e,n)=0`$ and $`𝒗_{𝝀\mathbf{=}\mathrm{𝟎}}\mathbf{}\mathbf{}𝒅\mathbf{=}\mathrm{𝟎}`$. These are the catastrophes of the Laplace transform $`ET`$ * Finally a multi-critical point where more than two phases become indistinguishable is at the branching of several lines with $`d=0`$, $`\mathbf{}𝒅\mathbf{=}\mathrm{𝟎}`$. Our classification of phase transitions by the topological structure of the micro-canonical Boltzmann entropy $`s(e,n)`$ is close to the natural experimental way to identify phase transitions of first order by the inhomogeneities at phase separation boundaries. This is possible because the micro-canonical ensemble does not suppress inhomogeneities in contrast to the grand-canonical one, as was emphasized already by Gibbs . Inter-phase boundaries are reflected in “small” systems by a convex intruder in the entropy surface. With this extension of the definition of phase transitions to “small” systems there are remarkable similarities with the transitions of the bulk. Moreover, this definition agrees with the conventional definition in the thermodynamic limit (of course, in the thermodynamic limit the largest curvature $`\lambda _1`$ approaches $`0`$ from above at phase transitions of first order). The region of phase separation remains inaccessible in the conventional grand-canonical ensemble. We believe, however, that the various kind of transitions discussed here have their immediate meaning in “small” and non-extensive systems independently whether they are the same in the thermodynamic limit (if this then exist) or not. For systems like the Potts model that have a thermodynamic limit it might well be possible that the character of the transition changes towards larger system size. The great conceptual clarity of micro-canonical thermo-statistics compared to the grand-canonical one is clearly demonstrated. Not only that, we showed that the micro-canonical statistics gives more information about the thermodynamic behavior and more insight into the mechanism of phase transitions than the canonical ensemble: About half of the whole {E,N} space, the intruder of $`S(E,N)`$ or the region between the ground state and the line $`\widehat{AP_mB}`$ in figure (2), gets lost in conventional grand-canonical thermodynamics. Without any doubts this contains the most sophisticated and interesting physics of this system. We emphasized this point already in there, however, with still limited precision. Due to our refined simulation method this could be demonstrated here with uniformly good precision in the whole {$`E,N`$} plane. It turns out that not only are non-extensive systems a new and rich realm for thermodynamics but moreover non-extensivity makes phase transitions much more transparant which is no surprise as phase transitions of first order are coupled to situations where a system prefers to become inhomogeneous, i.e. non-extensive. Finally, we should mention that micro-canonical thermo-statistics allowed us to compute phase transitions and especially the surface tension in realistic systems like small metal clusters . Our finding clearly disproves the pessimistic judgement by Schrödinger who thought that Boltzmann’s entropy is only usefull for gases. A recent application of micro-canonical thermo-statistics to thermodynamically unstable, collapsing systems under high angular momentum is found at . Acknowledgment: First if all I thank to E.V.Votyakov for performing most of the numerical work. I thank M.E.Fisher for the suggestion to study the Potts-3 model and to test how the multicritical point is described micro-canonically.
warning/0004/hep-ph0004019.html
ar5iv
text
# 5-10 GeV Neutrinos from Gamma-Ray Burst Fireballs ## I Introduction Gamma-ray burst (GRB) sources are distributed throughout the universe and their energy output is measured to be a substantial fraction of a solar rest mass equivalent . A variety of observations support the interpretation that these events are caused by cataclysmic stellar collapse or compact mergers, producing a fireball with bulk expansion Lorentz factor $`\mathrm{\Gamma }10^210^3`$. In the standard GRB model a fireball made up of $`\gamma ,\text{e}^\pm `$ and magnetic fields with an admixture of baryons is produced by the release of a large amount of energy $`E\text{ }>10^{53}`$ ergs in a region $`r_o10^7r_{o7}`$ cm (e.g. ). The observations indicate that typical fireballs are characterized by a luminosity $`L10^{52}L_{52}`$ erg s<sup>-1</sup> and durations $`t_w=10t_{w1}`$ s in the observer frame, with a large spread in both quantities. The outflow is controlled by the value of the dimensionless entropy $`\eta =(L/\dot{M}c^2)`$ injected at $`r_o`$. Previous discussions of fireball models have generally focused on the charged particle components, since they determine directly the photon signal. However, consideration of a neutron component introduces qualitatively new effects . In a $`p,n`$ fireball, for values of $`\eta \text{ }>400`$, the neutrons and protons acquire a relative drift velocity causing inelastic $`n,p`$ collisions and creating neutrinos. We investigate here the neutrino and photon signals from $`n,p`$ collisions following decoupling in GRB. The $`10`$ GeV neutrinos from this mechanism depend upon the presence of neutrons in the original explosion, but the neutrinos are created in simple physical processes occurring in the later stages of the fireball. On the other hand, the $`10^5`$ GeV neutrinos discussed in refs. require the acceleration in shock waves of ultra-high energy protons interacting with photons. Thus the 10 GeV and the $`10^5`$ GeV neutrinos reflect very different astrophysical processes and uncertainties. Other processes, e.g. neutrinos from $`p,p`$ collisions also require shocks but have lower efficiencies, while 10-30 MeV neutrinos from the original explosion are much harder to detect due to the lower cross sections. We show (§III) that the 10 GeV neutrinos could be detectable by future km<sup>3</sup> size detectors. The associated $`10`$ GeV $`\gamma `$-ray fluences are compatible with current detection rates, and may be detectable with future space missions. The dependence of these signals on the neutron/proton ratio $`\xi `$ provides a new tool to investigate the nature of the GRB progenitor systems. Moreover the predicted neutrino event rate depends on the asymptotic bulk Lorentz factor of the neutrons, which is linked to that of the protons. The latter affects all the electromagnetic observables from the GRB fireball, including the photospheric and shock radii, as well as the particle acceleration and non-thermal photon production. ## II Dynamics, n-p Decoupling and Pions Above the fireball injection radius $`r_o`$ the outflow velocity increases through conversion of internal energy into kinetic energy, the bulk Lorentz factor $`\mathrm{\Gamma }`$ varying as $`\mathrm{\Gamma }T_o^{}/T^{}=r/r_o`$, where $`T^{}`$ is the comoving temperature and $`T_o^{}=1.2L_{52}^{1/4}r_{o7}^{1/2}`$ MeV is the initial temperature at $`r_o`$ (henceforth denoting with primes quantities measured in the comoving frame). The flow may be considered spherical, which is a valid approximation also for a collimated outflow of opening angle $`\theta _j>\mathrm{\Gamma }^1`$, for the conditions discussed here. In a pure proton outflow the linear growth of $`\mathrm{\Gamma }`$ saturates when it reaches an asymptotic value $`\mathrm{\Gamma }_f\eta `$ constant, the value $`\eta `$ being achieved when the fireball converts all its luminosity into expansion kinetic energy. For an $`n,p`$ fireball, beyond the injection radius $`r_o`$ the comoving temperature is low and nuclear reactions are rare, so the $`n/p`$ ratio $`\xi `$ remains constant. Since the thermal velocities are non-relativistic, decoupling of the $`n`$ and $`p`$ fluids is essential for high-energy neutrino production. At the base of the outflow the $`n`$ and $`p`$ components are coupled by nuclear elastic scattering. In terms of the CM relative energy $`ϵ_{rel}`$ and the relative velocity $`v_{rel}`$ between nucleons, $`\sigma _{el}^{}v_{rel}^{}\sigma _oc`$. The CM energy dependence $`\sigma _{el}ϵ^{1/2}v_{rel}^1`$ is approximately valid between energies $``$ MeV and the pion production threshold $`140`$ MeV, and $`\sigma _o\sigma _\pi 3\times 10^{26}`$ cm<sup>2</sup> is the pion formation cross section above threshold. The $`p`$ and $`n`$ are cold in the comoving frame, and remain well coupled until the comoving $`n,p`$ scattering time $`t_{np}^{}(n_p^{}\sigma _oc)^1`$ becomes longer than the comoving expansion time $`t_{exp}^{}r/c\mathrm{\Gamma }`$. Denoting the comoving neutron density $`n_n^{}=\xi n_p^{}`$ with $`\xi \text{ }<1`$, mass conservation implies $`n_p^{}=L/[(1+\xi )4\pi r^2m_pc^3\mathrm{\Gamma }\eta ]`$. The $`n,p`$ decoupling occurs in the coasting or accelerating regimes depending on whether the dimensionless entropy $`\eta `$ is below or above the critical value $`\eta _\pi =`$ $`\left(L\sigma _\pi /4\pi m_pc^3r_o(1+\xi )\right)^{1/4}`$ (1) $``$ $`3.9\times 10^2L_{52}^{1/4}r_{o7}^{1/4}\left([1+\xi ]/2\right)^{1/4}.`$ (2) Figure 1 shows the dependence of $`\mathrm{\Gamma }`$ on radius for different $`\eta `$. For low values, $`\eta \text{ }<\eta _\pi `$, the condition $`t_{np}^{}\text{ }>t_{exp}^{}`$ is achieved at a radius $`r_{np}/r_o=\eta _\pi (\eta _\pi /\eta )^3`$, which is beyond the saturation radius $`r_s/r_o\eta `$ at which both $`n`$ and $`p`$ start to coast with $`\mathrm{\Gamma }\eta =`$ constant. In this case, even after decoupling both $`n`$ and $`p`$ continue to coast together due to inertia, and their relative velocities never reach the threshold for inelastic collisions. For $`\eta \text{ }>\eta _\pi `$, on the other hand, the $`n,p`$ decoupling condition $`t_{np}^{}\text{ }>t_{exp}^{}`$ occurs while the protons (and neutrons) are still accelerating as $`\mathrm{\Gamma }_p(r/r_o)`$, at a radius $$(r_{np}/r_o)=\eta _\pi (\eta /\eta _\pi )^{1/3},\text{for}\eta \text{ }>\eta _\pi .$$ (3) Beyond this decoupling radius the $`p`$ can still continue to accelerate with $`\mathrm{\Gamma }_pr`$ (as long as they remain coupled to the photons). However the neutrons are no longer accelerated, since they only interact with the protons, and they continue to coast with the value of $`\mathrm{\Gamma }\mathrm{\Gamma }_{nf}`$ constant achieved up to that point, $`\mathrm{\Gamma }_{nf}=`$ $`(3/4)\eta _\pi (\eta /\eta _\pi )^{1/3}(\text{for}\eta \text{ }>\eta _\pi )`$ (4) $``$ $`3\times 10^2L_{52}^{1/4}r_{o7}^{1/4}\left([1+\xi ]/2\right)^{1/4}(\eta /\eta _\pi )^{1/3},`$ (5) where the (3/4) factor comes from a numerical solution of the coupling equations. When the $`n,p`$ decoupling condition $`\eta \text{ }>\eta _\pi `$ is satisfied, the relative $`n,p`$ drift velocity $`v_{rel}c`$ and the inelastic pion production threshold $`ϵ^{}>`$ 140 MeV is reached. Since $`\sigma _o\sigma _\pi `$, the condition $`t_{np}^{}t_{exp}^{}`$ implies that the optical depth to pion formation is of order unity. Thus, for $`\eta \text{ }>\eta _\pi `$ the radius $`r_{np}r_\pi `$ is not only a decoupling radius but also an effective “pionospheric” radius. The lowest energy threshold processes at $`r_\pi `$ are $`p+n`$ $`p+p+\pi ^{}`$ $`\mu ^{}+\overline{\nu }_\mu e^{}+\overline{\nu }_e+\nu _\mu +\overline{\nu }_\mu `$ (6) $`n+n+\pi ^+`$ $`\mu ^++\nu _\mu e^++\nu _e+\overline{\nu }_\mu +\nu _\mu `$ (7) $`p+n+\pi ^0`$ $`\gamma +\gamma ,`$ (8) which occur in approximately equal ratios and with near unit total probability. The corresponding $`p+p(n+n)\pi ^\pm ,\pi ^0`$ processes do not involve a relative drift velocity (as do the $`p+n`$), and are thus less probable. Processes leading to multiple baryons are also suppressed due to the higher threshold, and for simplicity we restrict ourselves to the above $`p+n`$ processes. ## III 10 GeV Neutrinos and $`\gamma `$-rays The total number of neutrons carried by the fireball is $`N_n=`$ $`\left({\displaystyle \frac{\xi }{1+\xi }}\right){\displaystyle \frac{E}{\eta m_pc^2}}`$ (9) $``$ $`0.83\times 10^{53}E_{53}\left({\displaystyle \frac{2\xi }{1+\xi }}\right)\left({\displaystyle \frac{400}{\eta }}\right),`$ (10) The comoving optical depth $`\tau ^{}n_p^{}\sigma r/\mathrm{\Gamma }\sigma /(r\mathrm{\Gamma })`$ has the same dependence for pion formation and photon scattering, but $`\sigma _\pi \sigma _T`$ (Thomson cross section), so the pionosphere $`r_\pi `$ occurs below the $`\gamma `$-photosphere $`r_\gamma `$. The $`\gamma `$-rays in equation (8) can only escape from a skin depth below the $`\gamma `$-sphere in the essentially laminar flow with probability $`P_\gamma \text{ }<\tau _\pi (r_\gamma )r_\pi /r_\gamma (\sigma _\pi /\sigma _T)(1+\xi /7)^21/25`$, for $`\eta \text{ }>\eta _\pi `$. Each $`n`$ leads to $`1`$ photon of CM energy $`ϵ_\gamma ^{}70`$ MeV and observer energy centered broadly around $`ϵ_\gamma 70\mathrm{\Gamma }_{nf}/(1+z)\text{MeV}10`$ GeV. Using a proper distance $`D_p=2.8\times 10^{28}h_{65}^1[11/\sqrt{1+z}]`$ cm with a Hubble constant $`h_{65}=H_o/65\text{Km/s/Mpc}`$, the number fluence at Earth is $`N_\gamma N_nP_\gamma /4\pi D_p^210^5`$ cm<sup>-2</sup>. This is below the sensitivity of the $`200`$ cm<sup>2</sup> area EGRET detector on the Compton Gamma Ray Observatory (e.g. ), but for rare nearby bursts it may be detectable by GLAST . The neutrinos originate at the pionospheric radius $`r_\pi r_\gamma `$ where $`\tau _\pi 1`$. In this region the stable charged products and $`\gamma `$-rays from the reactions (8) remain in the fireball, and each $`n`$ leads on average to one $`\nu `$ and one $`\overline{\nu }`$. We list below the average neutrino energies for pions and muons decaying at rest. The neutrinos from muon decay have a continuum spectrum. Also, the energies are Doppler broadened by $`v_{rel}/c0.5`$. $`ϵ_{\overline{\nu }_\mu }^{}`$ $`30\text{MeV},ϵ_{\nu _\mu }^{}30\text{MeV}\text{ from }\pi ^\pm `$ (11) $`ϵ_{\nu _e}^{}`$ $`30\text{MeV},ϵ_{\overline{\nu }_\mu }^{}50\text{MeV}\text{ from }\mu ^+,`$ (12) $`ϵ_{\overline{\nu }_e}^{}`$ $`30\text{MeV},ϵ_{\nu _\mu }^{}50\text{MeV}\text{ from }\mu ^{}`$ (13) The relevant cross section for detection averaged over $`\nu `$ and $`\overline{\nu }`$ is $`\sigma _\nu 0.5\times 10^{38}(ϵ/\text{GeV})`$ cm<sup>2</sup> at the observed energy $`ϵ`$ . The observer frame energy is $`ϵ=ϵ^{}\alpha \mathrm{\Gamma }_{nf}/(1+z)`$, where $`\alpha 1`$ near threshold. For the CM $`\nu \overline{\nu }`$ production energies of equation (11), the average $`\nu +\overline{\nu }`$ CM energy per neutron is $`ϵ^{}100`$ MeV. Taking $`\alpha 1`$ the observer $`\nu +\overline{\nu }`$ energy per neutron is $`ϵ0.1\mathrm{\Gamma }_{nf}/(1+z)`$ GeV, and the effective detection cross section per neutron is $`\overline{\sigma }_{\nu \overline{\nu }}5\times 10^{40}\mathrm{\Gamma }_{nf}(1+z)^1`$ cm<sup>2</sup>. Multiplying by a burst rate within a Hubble distance of $`_b10^3_{b3}`$/year, for a 1 km<sup>3</sup> detector containing $`N_t10^{39}N_{t39}`$ target protons, the rate $`R_{\nu \overline{\nu }}=(N_t/4\pi D_p^2)_bN_n\overline{\sigma }_{\nu \overline{\nu }}`$ is $`R_{\nu \overline{\nu }_\mu }`$ $`7E_{53}N_{t39}_{b3}\left({\displaystyle \frac{2\xi }{1+\xi }}\right)\left({\displaystyle \frac{\eta _\pi }{\eta }}\right)^{4/3}`$ (14) $`\times h_{65}^2\left({\displaystyle \frac{2\sqrt{2}}{1+z\sqrt{1+z}}}\right)^2\text{year}^1,`$ (15) events in the detector in coincidence with GRB electromagnetic flashes. The energies of the events are $$ϵ_{\nu _\mu \overline{\nu }_\mu }10\text{GeV},ϵ_{\nu _e\overline{\nu }_e}5\text{GeV},$$ (16) which scale $`E_{53}^{1/4}t_{w1}^{1/4}r_{o7}^{1/4}`$ $`(2/[1+\xi ])^{1/4}(2/[1+z])`$ $`(\eta _\pi /\eta )^{1/3}`$. Subsequent to decoupling and $`n,p`$ collisions, each neutron decay $`np+e^{}+\overline{\nu }_e`$ leads to an additional $`\overline{\nu }_e`$ of CM energy $`ϵ_{\overline{\nu }_e,d}^{}0.8`$ MeV, which boosted in the observer frame by $`\mathrm{\Gamma }_{nf}/(1+z)`$ is $`\text{ }<120MeV`$. The cross section is $`\sigma _{\overline{\nu }_e}2\times 10^{40}`$ cm<sup>2</sup> and the expected rate in a km<sup>3</sup> detector is less than one event per year. ## IV MeV $`\gamma `$-Rays The non-thermal MeV $`\gamma `$-rays are thought to be produced in collisionless shocks, which occur at a radius $`r_{sh}>r_\gamma >r_\pi `$, after the bulk Lorentz factor has saturated to its asymptotic value. For an $`n,p`$ outflow, shocks can occur in the original $`p`$, as well as in the $`n`$ component after the latter have decayed, and this can influence the external shock light curves. A separate and important consequence of neutron decay is that it should also affect the internal shock gamma-ray light-curves. In the proton component internal shocks occur at $`r_{sh}ct_v\mathrm{\Gamma }_{pf}^2`$, where $`t_v`$ is the variability timescale, and $`\mathrm{\Gamma }_{pf}`$ is the asymptotic proton Lorentz factor. From energy conservation, for $`\eta >\eta _\pi `$ this is $`\mathrm{\Gamma }_{pf}\eta (1+\xi )[1(\xi /[1+\xi ])(6/7)(\eta _\pi /\eta )^{4/3}]`$, and taking into account photon drag one can show that an upper limit is $`\mathrm{\Gamma }_{pf,max}\text{ }<8.3\times 10^2E_{53}^{1/4}t_{w1}^{1/4}r_{o7}^{1/4}`$ $`(1+8\xi /7)^{1/4}`$. The $`\gamma `$-rays start to arrive at an observer time $`tt_v\text{ }>10^3t_{v3}`$ s, lasting for a time $`t_w`$ (where $`10^3\text{ }<t_w\text{ }<10^3`$ s). For the $`n`$ component, $`r_{sh}ct_{min}\mathrm{\Gamma }_{nf}^2`$, with $`\mathrm{\Gamma }_{nf}`$ from equation (5) and $`t_{min}\mathrm{min}[t_v,t_n^{}\mathrm{\Gamma }_{nf}^1]`$, where $`t_n^{}10^3`$ s is the comoving frame neutron decay time. Taking $`\xi 1`$ in the estimates below, for $`20\text{ }<\eta \text{ }<\eta _\pi 400`$ the neutrons decay and shock beyond the proton shock for any $`t_v\text{ }<10^3\eta ^1`$, at observer times $`t_n=[50s,3s]`$, while for $`\eta \text{ }>\eta _\pi 400`$ the neutrons decay and shock beyond the proton shock for any $`t_v\text{ }<3(\eta _\pi /\eta )^{1/3}`$ s. The typical observed duration of the decay, including the blue shift due to the bulk motion towards the observer, is $`t_n10^3/\mathrm{\Gamma }_{nf}`$, where $`\mathrm{\Gamma }_{nf}\eta `$ for $`\eta <\eta _\pi `$ and $`\mathrm{\Gamma }_{nf}=(3/4)\eta _\pi (\eta /\eta _\pi )^{1/3}`$ for $`\eta \eta _\pi `$. Thus $`t_n`$ decreases from approximately 50 s to 3 s for $`20\text{ }<\eta \text{ }<\eta _\pi 400`$, and then slowly increases again as $`t_n3(\eta /\eta _\pi )^{1/3}`$ for $`\eta \text{ }>\eta _\pi 400`$, with both $`\eta _\pi `$ and $`t_n`$ scaling $`[(1+\xi )/2]^{1/4}`$. The number of neutron decays is $`1\mathrm{exp}(t/t_n)`$, so the envelope of the neutron-related light curve is the mirror image of a “fred” (fast rise - exponential decay), i.e. an “anti-fred” (or generally, slow rise - fast decay). In general, photon emission starts at $`t_v`$ from the proton-related component, which lasts a time $`t_w`$ with an arbitrary shape envelope, modulated by spikes of minimum duration $`t_v<t_w`$, depending on the chaotic behavior of the central engine producing the outflow. The neutron-related component starts at a later time $`t_n>t_v`$, and has an anti-fred shaped envelope modulated by spikes of $`t_v`$ and a total duration $`t_w`$. If $`t_w>t_n`$, the anti-fred component would be hard to distinguish because of the superposition of the ongoing $`p`$ and $`n`$ components. However, for short bursts with $`t_w<t_n3`$ s, the $`p`$ and $`n`$ components are separated: first there is a pulse of duration $`t_w`$ with a random envelope, followed after a time $`t_n`$ by a pulse with an anti-fred envelope of duration $`t_n`$, and characteristic photon energy softer than the previous by $`ϵ_n/ϵ_pt_w/t_n`$ (which if small could be below the BATSE band, but may be detectable with the Swift satellite). The latter pulse is a signature for neutron decay in the burst. ## V Discussion For characteristic parameters, GRB outflows produce 5-10 GeV $`\nu _\mu \overline{\nu }_\mu `$ and $`\nu _e\overline{\nu }_e`$ from internal inelastic $`p,n`$ collisions that create pions. The $`\nu \overline{\nu }`$ energy output $`E_{\nu \overline{\nu }}5\times 10^{51}E_{53}(2\xi /[1+\xi ])(2/[1+z])(\eta _\pi /\eta )^{4/3}`$ ergs depends on the total energy $`E`$ of the GRB and on the neutron fraction $`\xi `$ as well as on the dimensionless entropy $`\eta `$. For a km<sup>3</sup> detector, approximately 5 to 10 neutrino events above 10 GeV are predicted per year, for a neutron/proton ratio $`\xi =1`$. These events will be coincident with GRB electromagnetic flashes in direction and in time (to an accuracy of $`10`$ s), which can enable their separation from the atmospheric neutrino background. Underground water detectors of the type being planned by BAIKAL , NESTOR , ANTARES and the Antartic detector ICECUBE could potentially detect these relatively low energy neutrino events if a sufficiently high density of phototubes were used. About 80% of these neutrinos are $`\nu _\mu `$ and $`\overline{\nu }_\mu `$ (in approximately equal numbers) and the remainder are $`\nu _e`$ and $`\overline{\nu }_e`$. These 5-10 Gev $`\nu \overline{\nu }`$ are followed by $``$ 120 MeV $`\overline{\nu }_e`$ from neutron decay, but the event rate from neutron-decay neutrinos is very low. The higher energy neutrinos are produced for neutron/proton ratios $`\xi >0`$ when the dimensionless entropy $`\eta =L/\dot{M}c^2`$ exceeds $`\eta _\pi 4\times 10^2L_{52}^{1/4}r_{o7}^{1/4}`$ $`(2/[1+\xi ])^{1/4}`$, and are accompanied by $`10`$ GeV photons which may be detectable in low redshift cases with GLAST . For a typical GRB at redshift $`z1`$ the number fluences in 10 GeV neutrinos are $`N(\overline{\nu }_e+\nu _e)0.5N(\overline{\nu }_\mu +\nu _\mu )`$ $`10^4`$ cm<sup>-2</sup>, and one order of magnitude less for GeV photons. In all bursts where $`\xi >0`$ the lower energy ($`120`$ MeV) neutrinos are produced, and neutron decay occurs on an observer timescale $`t_n3L_{52}^{1/4}r_{o7}^{1/4}[(1+\xi )/2]^{1/4}(\eta /\eta _\pi )^{1/3}`$ s. For outflows of duration $`t_w`$, these decays will be associated with MeV electromagnetic pulses of duration $`\mathrm{min}[t_n,t_w]`$, which are additional to the MeV pulses expected from shocks in the original proton component. For short bursts with $`t_w\text{ }<3`$ s, the proton electromagnetic pulse appears first and is separated from a subsequent neutron electromagnetic pulse, the latter having a slow rise-fast decay envelope and a softer spectrum, which may be detectable with the Swift satellite. A systematic study of the time histories of GRB emission would be useful to search for evidence of delayed pulses that might be caused by neutron decay. The detection of 5-10 GeV $`\nu \overline{\nu }`$ in coincidence with GRB photon flashes will not be easy, but would provide unique astrophysical information. Constraints on the neutron fraction could provide information about the progenitor stellar system giving rise to GRB. For instance, core collapse of massive stars would lead to an outflow from an $`Fe`$-rich core with $`\xi 2/31`$, while neutron star mergers would imply $`\xi 1`$. Photodissociation during collapse or merger, as well as $`n,p`$ decoupling and inelastic collisions, would both drive $`\xi `$ toward unity, although this equalization process is likely to remain incomplete. For low $`\eta \text{ }<\eta _\pi `$, inelastic collisions are not expected and the 5-10 GeV $`\nu \overline{\nu }`$ are absent, producing only the harder to detect $`100`$ MeV $`\overline{\nu }_e`$ from neutron decay. An initially non-baryonic outflow of, e.g. $`e^\pm `$ and magnetic fields, would acquire a baryonic load by entrainment from the progenitor environment, with $`\xi 1`$ from massive stellar envelopes, but $`\xi \text{ }>1`$ for, e.g., compact mergers. Thus, lower values of $`\xi `$, leading to lower ratios of 5-10 GeV $`\nu \overline{\nu }`$ and a lower ratio of neutron decay MeV photons to total fluences would be expected from massive progenitors than from compact mergers. Acknowledgements Partial support was received by JNB from NSF PHY95-13835 and by PM from NASA NAG5-2857, the Guggenheim Foundation and the Institute for Advanced Study. We acknowledge valuable conversations with G. Fishman, V. Fitch, Vl. Kocharovsky, P. Kumar, R. Nemiroff, J. Norris, M.J. Rees, P. Vogel and E. Waxman.
warning/0004/astro-ph0004397.html
ar5iv
text
# Dark matter scaling relations ## 1. Introduction It is now well established that spiral galaxies have universal rotation curves (URC) that can be characterized by one single free parameter, the luminosity (e.g. Rubin et al. 1980; Persic & Salucci, 1991, Persic, Salucci & Stel 1996 (PSS)). For instance, low-luminosity spirals show ever-rising rotation curves (RC) out to the optical radius <sup>1</sup><sup>1</sup>1 $`R_{opt}=3.2R_d`$, with $`R_d`$ the exponential disk length-scale, while, in the same region, the RC of high-luminosity spirals are flat or even decreasing. It has been demonstrated by Persic & Salucci (1988, 1990) and Broeils (1992) that, as the galaxy luminosity decreases, the light is progressively unable to trace the radial distribution of the gravitating matter (see also Salucci, 1997). This discrepancy is, in general, interpreted as the signature of an invisible mass component (Rubin et.al. 1980; Bosma 1981). As pointed out by PSS the universality of the rotation curves, in combination with the invariant distribution of the luminous matter, implies an universal dark matter distribution with luminosity-dependent scaling properties. On the theoretical side, recent high-resolution cosmological N-body simulations have shown that cold dark matter halos achieve a specific equilibrium density profile (Navarro, Frenk & White 1996, NFW; Cole & Lacey 1997; Fukushige & Makino 1997; Moore et al. 1998; Kravtsov et al. 1998). This can be characterized by one free parameter, e.g. $`M_{200}`$, the dark mass enclosed within the radius inside which the average over-density is 200 times the critical density of the Universe. In the innermost regions the dark matter profiles show some scatter around an average profile which is characterized by a power-law cusp $`\rho r^\gamma `$, with $`\gamma =11.5`$ (NFW, Moore et al. 1998, Bullock et al., 1999). Until recently, due to both the limited number of suitable rotation curves and a poor knowledge of the exact amount of luminous matter present in the innermost regions of spirals, it has been difficult to investigate the internal structure of dark matter halos. The situation is more favorable for (low surface brightness) dwarf galaxies which are strongly dark matter dominated even at small radii. The kinematics of these systems shows an universality of the dark halo density profiles, but, it results in disagreement with that predicted by CDM, in particular because of the existence of dark halo density cores. (Moore 1994; Burkert 1995). The origin of these features is not yet understood (see e.g. Navarro, Eke & Frenk 1996; Burkert & Silk 1997, Gelato & Sommer-Larsen 1999), but it is likely that it involves more physics than a simple hierarchical assembly of cold structures. To cope with this observational evidence, Burkert (1995) proposed an empirical profile that successfully fitted the halo rotation curves of four dark matter dominated dwarf galaxies $$\rho _b(r)=\frac{\rho _0r_0^3}{(r+r_0)(r^2+r_0^2)}$$ (1) where $`\rho _0`$ and $`r_0`$ are free parameters which represent the central dark matter density and the scale radius. This sample has been extended, more recently, to 17 dwarf irregular and low surface brightness galaxies (Kravtsov et al. 1998, see however van den Bosch et al. 1999) which all are found to confirm equation (1). Adopting spherical symmetry, the mass distribution of the Burkert halos is given by $$M_b(r)=4M_0\left\{ln\left(1+\frac{r}{r_0}\right)tg^1\left(\frac{r}{r_0}\right)+\frac{1}{2}ln\left[1+\left(\frac{r}{r_0}\right)^2\right]\right\}$$ (2) with $`M_0`$, the dark mass within the core given by: $$M_0=1.6\rho _0r_0^3$$ (3) The halo contribution to the circular velocity is then: $$V_b^2(r)=GM_b(r)/r.$$ (4) Although the dark matter core parameters $`r_0`$, $`\rho _0`$ and $`M_0`$ are in principle independent, the observations reveal a clear correlation (Burkert 1995): $$M_0=4.3\times 10^7\left(\frac{r_0}{kpc}\right)^{7/3}M_{}$$ (5) which indicates that dark halos represent a 1-parameter family which is completely specified, e. g. by the core mass. The analysis of a recently published large sample of RC’s (Persic & Salucci, 1995) has provided a suitable framework to investigate the dark halo density distribution in spirals. The starting points of this study are: a) the mass in spirals is distributed according to the Inner Baryon Dominance (IBD) regime: there is a characteristic transition radius $`R_{IBD}2R_d\left(\frac{V_{opt}}{220km/s}\right)^{1.2}`$ for which, at $`rR_{IBD}`$, the luminous matter totally accounts for the mass distribution, while, for $`r>R_{IBD}`$, the DM rapidly becomes the dominant dynamical component (Salucci and Persic, 1999a,b; Salucci et al, 2000; Ratnam and Salucci, 2000; Borriello and Salucci, 2000). Then, although the dark halo might extend down to the galaxy center, it is only for $`r>R_{IBD}`$ that it gives a non-negligible contribution to the circular velocity. b) The dark matter is distributed in a different way with respect to any of the various baryonic components (PSS, Corbelli and Salucci, 2000), and c) The HI contribution to the circular velocity, for $`r<R_{opt}`$, is negligible (e.g. Rhee, 1996 ; Verheijen, 1997). The main aim of this letter is to expand the above results to derive the luminosity-averaged density profiles of the dark halos and to relate them with the Burkert profiles. Section 2 presents the analysis of a homogeneous sample of about 1100 rotation curves in which the dark halo contribution to the circular velocity is first derived and then matched to the Burkert halo mass models. In section 3 we discuss the results. We take $`H_0=75km/s/Mpc`$ and $`\mathrm{\Omega }_0=0.3`$, however no result depends on these choices. ## 2. DM halo profiles in Spirals PSS (see also Rubin et al. 1982) have derived, from $`15000`$ velocity measurements of 1000 rotation curves (RC), $`V_{syn}(\frac{r}{R_{opt}},\frac{L_I}{L_{}})`$, the synthetic rotation velocities of spirals sorted by luminosity (Figure 1, with $`L_I`$ the I-band luminosity and $`L_I/L_{}=10^{(M_I+21.9)/5})`$. Remarkably, individual RC’s have a very small variance with respect to the corresponding synthetic curves (PSS, Rhee 1996, Rhee & van Albada 1996, Roscoe 1999, Swaters, 1999): spirals sweep a very narrow locus in the RC-profile/amplitude/luminosity space. On the other hand, the galaxy kinematical properties do significantly change with luminosity (e.g. PSS), so it is natural to relate the mass distribution with this quantity. The whole set of synthetic RC’s has been modeled with the Universal Rotation Curve (URC), $`V_{URC}(r/R_{opt},L_I/L_{})`$ which includes: (a) an exponential thin disk term: $$V_{d,URC}^2(x)=1.28\beta V_{opt}^2x^2(I_0K_0I_1K_1)|_{1.6x}$$ (6) and (b) a spherical halo term: $$V_{h,URC}^2(x)=V_{opt}^2(1\beta )(1+a^2)\frac{x^2}{(x^2+a^2)},$$ (7) with $`xr/R_{opt}`$, $`\beta (V_{d,URC}(R_{opt})/V_{opt})^2`$, $`V_{opt}V(R_{opt})`$ and $`a`$ the halo core radius in units of $`R_{opt}`$. At high luminosities the contribution from a bulge component has been also considered (Salucci and Persic, 1999b). Let us stress that the halo velocity term of eq. (7) does not bias the mass model: it can account for maximum-disk, solid-body, no-halo, all-halo, CDM and core-less halo mass models. In practice, the values of the free parameters $`a`$ and $`\beta `$ that result from fitting the URC $$V_{URC}^2(x)=V_{h,URC}^2(x,\beta ,a)+V_{d,URC}^2(x,\beta )$$ (8) to the synthetic curves $`V_{syn}`$ select the actual model. Adopting $`\beta 0.72+0.44log(\frac{L_I}{L_{}})`$ and $`a1.5(\frac{L_I}{L_{}})^{1/5}`$ (see PSS) or, equivalently the corresponding $$a=a(\beta )\beta =\beta (logV_{opt})$$ (9) that we have plotted in Fig. (2), the URC mass models reproduce the synthetic curves $`V_{syn}(r)`$ within their r.m.s. (see Fig. (1)). More in detail, at any luminosity and radius, $`|V_{URC}V_{syn}|<2\%`$ and the $`1\sigma `$ fitting uncertainties on $`a`$ and $`\beta `$ are about 20% (PSS). We then compare the dark halo velocities obtained with eq. (7) and (9), with the Burkert velocities $`V_b(r)`$ of eqs. (2)-(4). We leave $`\rho _0`$ and $`r_0`$ as free parameters, i. e. we do not impose the relationship of eq. (5). The results are shown in Fig (3): at any luminosity, out to the outermost radii ($`6R_d`$), $`V_b(r)`$ is indistinguishable from $`V_{h,URC}(r)`$. More specifically, by setting $`V_{h,URC}(r)V_b(r)`$, we are able to reproduce the synthetic rotation curves $`V_{syn}(r)`$ at the level of their r.m.s. For $`r>>6R_D`$, i.e. beyond the region described by the URC, the two velocity profiles progressively differ. The values obtained for $`r_0`$ and $`\rho _0`$ for the URC agrees with the extrapolation at high masses of the scaling law $`\rho r_0^{2/3}`$ (Burkert, 1995) established for objects with core radii $`r_0`$ ten times smaller (see Fig 4). Let us notice that the core radii are very large: $`r_0>>R_d`$ so that an ever-rising halo RC cannot be excluded by the data. Moreover, the disk-mass vs. central halo density relationship $`\rho _0M_d^{1/3}`$, found in dwarf galaxies (Burkert,1995), where the densest halos harbor the least massive disks, holds also for disk systems of stellar mass up to $`10^{11}M_{}`$ (see Fig 4). The above relationship show a curvature at the highest masses/lowest densities that can be related to the existence of an upper limit in the dark halo mass $`M_{200}`$ <sup>2</sup><sup>2</sup>2The virial halo mass is given by; $`M_{200}200\times 4\pi /3\rho _cR_{200}^3\mathrm{\Omega }_0(1+z^3)g(z)`$ with $`z`$ the formation redshift, $`R_{200}`$ the virial radius, for $`g(z)`$ see e.g. Bullock et al., (1999); the critical density is defined as: $`\rho _c3/(8\pi )G^1H_0^2`$. which is evident by the sudden decline of the baryonic mass function of disk galaxies at $`M_d^{max}=2\times 10^{11}M_{}`$ (Salucci and Persic, 1999a), that implies a maximum halo mass of $$M_{200}^{max}\mathrm{\Omega }_0/\mathrm{\Omega }_bM_d^{max}$$ (10) where $`\mathrm{\Omega }_0`$ and $`\mathrm{\Omega }_b0.03`$ (e.g. Burles and Tytler, 1998) are the matter and baryonic densities of the Universe in units of critical density. From the definition of $`M_{200}`$, by means of eq. (2) and (3), we can write $`M_{200}`$ in terms of the ”observable” quantity $`M_0`$: $`M_{200}=\eta M_0`$. For $`(\mathrm{\Omega }_0,z)=(0.3,3)`$, $`\eta 12`$; notice that there is a mild dependences of $`\eta `$ on $`z`$ and $`\mathrm{\Omega }_0`$ which is irrelevant for the present study. Combining eq. (3) and (10) we obtain an upper limit for the central density, $`\rho _0<1\times 10^{20}(r_0/kpc)^3g/cm^3`$, which implies a lack of objects with $`\rho _0>4\times 10^{25}g/cm^3`$ and $`r_0>30kpc`$, as is evident in Figure 3. Turning the argument around, the deficit of objects with $`M_dM_d^{max}`$ and $`\rho _0>4\times 10^{25}g/cm^3`$, suggests that, at this mass scale, the total-to- baryonic mass fraction may approache the cosmological value $`\mathrm{\Omega }/\mathrm{\Omega }_b10`$. ## 3. Discussion Out to two optical radii, the Burkert density profile reproduces, for the whole spiral luminosity sequence, the DM halos mass distribution. This density profile, though at very large radii coincides with the NFW profile, approaches a constant, finite density value at the center, in a way consistent with an isothermal distribution. This is in contradiction to cosmological models (e.g. Fukushige and Makino 1997) which predict that the velocity dispersion $`\sigma `$ of the dark matter particles decreases towards the center to reach $`\sigma 0`$ for $`r0`$. After the result of this study, the dark halo inner regions, therefore, cannot be considered as kinematically cold structures but rather as ”warm” regions with size $`r_0\rho _0^{1.5}`$. The halo core sizes are very large: $`r_047R_d`$. Then, the boundary of the core region is well beyond the region where the stars are located and, as in Corbelli and Salucci (2000), even at the outermost observed radius there is not the slightest evidence that dark halos converge to a $`\rho r^2`$ (or a steeper) regime. We find that the dark halos around spirals are essentially an one-parameter family. It is relevant that the order parameter (the central density or the core radius) correlates with the luminous mass (see Fig 4). We do however not know how it is related to the global structural properties of the dark halo, like the virial radius or the virial mass. In fact, the RC out to $`6R_D`$ is completely determined by the core parameters, i.e. the central core density and the core radius, both of which are not defined in the CDM scenario. The location of spiral galaxies in the parameter space of virial mass, halo central density and baryonic mass is determined by different processes that occur on different scales and at different red-shifts. Yet, this 3D space degenerates into a single curve (see Figure 4,remind that: $`\rho _0=\frac{\pi }{24}M_{200}/r_0^3`$ and remind that: $`M_d=G^1\beta V_{opt}^2R_{opt})`$ which describes the dark-luminous coupling. Let us discuss the limitations of the present results. First, here we have considered the luminosity dependence of the dark halo structure. Although this is probably the most relevant one, other dependences (Hubble type and surface brightness) should also be investigated. Moreover, the existence of a (weak) cosmic variance in the halo structural properties cannot be excluded until we analyze individual objects (Salucci, 2000, Borriello and Salucci, 2000). Secondly, we have derived the profile of DM halos out to about six disk-scale lengths, i.e. out to a distance much smaller than the virial radius. To assess the global validity of the proposed mass model data at larger radii are obviously required.
warning/0004/hep-ph0004004.html
ar5iv
text
# Fat penguins and imaginary penguins in perturbative QCD ## Abstract We evaluate $`BK\pi `$ decay amplitudes in perturbative QCD picture. It is found that penguin contributions are dynamically enhanced by nearly 50% compared to those assumed in the factorization approximation. It is also shown that annihilation diagrams are not negligible, and give large strong phases. Our results for branching ratios of $`BK\pi `$ decays for a representative parameter set are consistent with data. preprint: KEK-TH-642, NCKU-HEP-00-01, KIAS-P00014, DPNU-00-13 Factorization assumption (FA) for nonleptonic two-body $`B`$ and $`D`$ meson decays pioneered by Stech and his collaborators has been extremely successful. It gives correct order of magnitude for branching ratios of most two-body $`B`$ meson decays. Why does it work so well ? Recent CLEO data of $`B\pi \pi `$ and $`BK\pi `$ branching ratios require not only order-of-magnitude predictions but quantitative predictions for these decay modes. As asymmetric $`B`$ factories, which are eventually capable of producing almost $`10^8`$ $`B`$’s per year, have started their operation, quantitative theoretical understanding will allow us to extract CP phases hidden in the above branching ratios. How can we go beyond FA ? Let us see what QCD can say about these questions. The fact that 5 GeV of energy is released and shared by two light mesons suggests that the basic interaction in two-body $`B`$ meson decays is mainly short-distance. In this letter we shall attempt to compute these decay amplitudes using as much information from the underlying theory, QCD, as possible. Our method is perturbative QCD (PQCD) factorization theorem, which has been worked out by Li and his collaborators based on the formalism developed by Brodsky and Lapage and by Botts and Sterman . Consider the specific $`\overline{B}^0K^{}\pi ^+`$ decay amplitude shown in Fig. 1, where the $`bs\overline{u}u`$ decay occurs. The pair of the $`s`$ and $`\overline{u}`$ quarks fly away and form the $`K^{}`$ meson. The spectator $`\overline{d}`$ quark of the $`\overline{B}^0`$ meson is more or less at rest and the $`u`$ quark is flying away. The probability that a quark and an antiquark with large relative velocity form the $`\pi ^+`$ meson is suppressed by the pion wave function. How big is the suppression ? It depends on the functional form of the wave function. It is safe to say that this suppression from the wave function is of the form $`\left(\mathrm{\Lambda }_{\mathrm{QCD}}/M_B\right)^n`$, where $`n`$ is likely to be large. Therefore, we expect that dominant contributions to the $`\overline{B}^0K^{}\pi ^+`$ decay come from the process, where a hard gluon is exchanged so that $`\overline{d}`$ quark momentum and $`u`$ quark momentum are aligned to form the pion. The rectangular dotted boxes in Fig. 1 enclose the part of interaction which is hard. The blobs represent wave functions giving amplitudes for a quark and an antiquark to form a meson. For the $`B`$ meson mass $`M_B\mathrm{\Lambda }_{\mathrm{QCD}}`$ and the kaon and pion masses $`M_KM_\pi 0`$, the $`\overline{B}^0K^{}\pi ^+`$ decay amplitude is then written as a convolution of four factors, $$M=d[x]d[𝐛]\varphi _B(x_1,𝐛_1)\varphi _K(x_2,𝐛_2)\varphi _\pi (x_3,𝐛_3)H([x],[𝐛],M_B),$$ (1) where the wave functions $`\varphi _{B,K,\pi }`$ for the $`\overline{B}^0,K^{},\pi ^+`$ mesons absorb nonperturbative dynamics of the process, the hard amplitude $`H([x],[𝐛],M_B)`$ can be calculated in perturbation theory, $`[x]`$ is a shorthand for the momentum fractions $`x_1`$, $`x_2`$, and $`x_3`$ associated with the $`\overline{d}`$ quark in the $`\overline{B}^0`$ meson, the $`u`$ quark, and the $`\overline{d}`$ quark in the $`\pi ^+`$ meson, respectively, and $`[𝐛]`$ is a shorthand for the two-dimensional vectors, i.e., the transverse extents, $`𝐛_1`$, $`𝐛_2`$, and $`𝐛_3`$ of the $`\overline{B}^0`$, $`K^{}`$ and $`\pi ^+`$ mesons, respectively. Here are some important questions: 1. Is $`H([x],[𝐛],M_B)`$ really dominated by short-distance contributions ? 2. Figure 1(a) is factorizable, since it can be written in terms of the $`B\pi `$ transition form factor $`F^{B\pi }`$ and the kaon decay constant $`f_K`$. This is the amplitude considered in FA. FA assumes that a nonfactorizable amplitude from Fig. 1(b), which can not be written in terms of a form factor and a decay constant, is negligible compared to Fig. 1(a). Are nonfactorizable amplitudes negligible compared to factorizable ones ? We have made a numerical study of this issue. While it is important to always check the relative magnitudes, we have found that nonfactorizable contributions are usually less than few percents of factorizable contributions. This is the reason FA has been so successful. However, there are exceptions: (1) In $`BD\pi `$ decays some nonfactorizable contributions can reach as much as 30% of factorizable ones. Actually, the experimental fact that the ratio $`a_2/a_10.2`$ in the Bauer-Stech-Wirbel model requires large nonfactorizable contributions. (2) In $`BJ/\psi K^{()}`$ decays nonfactorizable and factorizable contributions are of the same order of magnitude . 3. It is well known that the role of penguins is essential for explaining the observed $`BK\pi ,\pi \pi `$ branching ratios. How big are penguin amplitudes ? We shall show below that penguin amplitudes can be dynamically enhanced by 50% in PQCD compared to those assumed in FA. 4. Are annihilation diagrams in Fig. 2 really negligible ? 5. How big are final-state-interaction (FSI) effects ? It is impossible to compute FSI phases in FA. Effects from infinite soft gluon exchanges among mesons in two-body $`B`$ meosn decays have been analyzed quantitatively by means of renormalization-group methods and found to be small . This observation implies that effects from exchange of soft objects between the two final-state mesons are also small. Where then do strong phases come from ? We shall show that contrary to common belief, annihilation diagrams are important, and in fact, they contribute large strong phases. We present the factorizable PQCD amplitudes $`F_e`$, $`F_{e4}^P`$, and $`F_{e6}^P`$ corresponding to Fig.1(a) and $`F_a`$, $`F_{a4}^P`$, and $`F_{a6}^P`$ corresponding to Fig.2(a) from the four-quark operators $`O_{1,2}`$, $`O_{3,4,9,10}`$, and $`O_{5,6,7,8}`$, respectively , $`F_{e4}^P`$ $`=`$ $`16\pi C_FM_B^2{\displaystyle _0^1}𝑑x_1𝑑x_3{\displaystyle _0^{\mathrm{}}}b_1𝑑b_1b_3𝑑b_3\varphi _B(x_1,b_1)`$ (4) $`\times \{[(1+x_3)\varphi _\pi (x_3)+r_\pi (12x_3)\varphi _\pi ^{}(x_3)]E_{e4}(t_e^{(1)})h_e(x_1,x_3,b_1,b_3,M_B)`$ $`+2r_\pi \varphi _\pi ^{}(x_3)E_{e4}(t_e^{(2)})h_e(x_3,x_1,b_3,b_1,M_B)\},`$ $`F_{e6}^P`$ $`=`$ $`32\pi C_FM_B^2{\displaystyle _0^1}𝑑x_1𝑑x_3{\displaystyle _0^{\mathrm{}}}b_1𝑑b_1b_3𝑑b_3\varphi _B(x_1,b_1)`$ (7) $`\times r_K\{[\varphi _\pi (x_3)+r_\pi (2+x_3)\varphi _\pi ^{}(x_3)]E_{e6}(t_e^{(1)})h_e(x_1,x_3,b_1,b_3,M_B)`$ $`+[x_1\varphi _\pi (x_3)+2r_\pi (1x_1)\varphi _\pi ^{}(x_3)]E_{e6}(t_e^{(2)})h_e(x_3,x_1,b_3,b_1,M_B)\},`$ $`F_{a4}^P`$ $`=`$ $`16\pi C_FM_B^2{\displaystyle _0^1}𝑑x_2𝑑x_3{\displaystyle _0^{\mathrm{}}}b_2𝑑b_2b_3𝑑b_3`$ (10) $`\times \{[x_3\varphi _K(x_2)\varphi _\pi (x_3)2r_\pi r_K(1+x_3)\varphi _K^{}(x_2)\varphi _\pi ^{}(x_3)]E_{a4}(t_a^{(1)})h_a(x_2,x_3,b_2,b_3,M_B)`$ $`+[x_2\varphi _K(x_2)\varphi _\pi (x_3)+2r_\pi r_K(1+x_2)\varphi _K^{}(x_2)\varphi _\pi ^{}(x_3)]E_{a4}(t_a^{(2)})h_a(x_3,x_2,b_3,b_2,M_B)\},`$ $`F_{a6}^P`$ $`=`$ $`32\pi C_FM_B^2{\displaystyle _0^1}𝑑x_2𝑑x_3{\displaystyle _0^{\mathrm{}}}b_2𝑑b_2b_3𝑑b_3`$ (13) $`\times \{[r_\pi x_3\varphi _K(x_2)\varphi _\pi ^{}(x_3)+2r_K\varphi _K^{}(x_2)\varphi _\pi (x_3)]E_{a6}(t_a^{(1)})h_a(x_2,x_3,b_2,b_3,M_B)`$ $`+[2r_\pi \varphi _K(x_2)\varphi _\pi ^{}(x_3)+r_Kx_2\varphi _K^{}(x_2)\varphi _\pi (x_3)]E_{a6}(t_a^{(2)})h_a(x_3,x_2,b_3,b_2,M_B)\},`$ $`C_F`$ being a color factor. The expression of $`F_e`$ ($`F_a`$) for the $`O_{1,2}`$ contributions is the same as $`F_{e4}^P`$ ($`F_{a4}^P`$) but with the Wilson coefficient $`a_1(t_e)`$ ($`a_1(t_a)`$). The hard functions $`h`$’s in Eqs (4)-(13) are given by $`h_e(x_1,x_3,b_1,b_3,M_B)`$ $`=`$ $`K_0\left(\sqrt{x_1x_3}M_Bb_1\right)`$ (15) $`\times [\theta (b_1b_3)K_0\left(\sqrt{x_3}M_Bb_1\right)I_0\left(\sqrt{x_3}M_Bb_3\right)+(b_1b_3)],`$ $`h_a(x_2,x_3,b_2,b_3,M_B)`$ $`=`$ $`\left({\displaystyle \frac{i\pi }{2}}\right)^2H_0^{(1)}\left(\sqrt{x_2x_3}M_Bb_2\right)`$ (17) $`\times [\theta (b_2b_3)H_0^{(1)}\left(\sqrt{x_3}M_Bb_2\right)J_0\left(\sqrt{x_3}M_Bb_3\right)+(b_1b_3)].`$ The evolution factors $$E_{ei}(t)=\alpha _s(t)a_i(t)\mathrm{exp}[S_B(t)S_\pi (t)],E_{ai}(t)=\alpha _s(t)a_i(t)\mathrm{exp}[S_K(t)S_\pi (t)],$$ (18) arise from the summation of infinite infrared gluon emissions that give double (Sudakov) logarithms and single logarithms connecting the hard scales $`t`$ and the characteristic scales $`1/b`$ of the wave functions. For the explicit expressions of the Sudakov exponents $`S_B`$, $`S_K`$, and $`S_\pi `$, refer to . The hard scales $`t`$ are chosen as the virtualities of internal particles in hard $`b`$ quark decay amplitudes, $`t_e^{(1)}=\mathrm{max}(\sqrt{x_3}M_B,1/b_1,1/b_3),`$ $`t_e^{(2)}=\mathrm{max}(\sqrt{x_1}M_B,1/b_1,1/b_3),`$ (19) $`t_a^{(1)}=\mathrm{max}(\sqrt{x_3}M_B,1/b_2,1/b_3),`$ $`t_a^{(2)}=\mathrm{max}(\sqrt{x_2}M_B,1/b_2,1/b_3).`$ (20) It has been shown that this choice minimizes higher-order corrections to exclusive QCD processes . Equation (20) is consistent with the fact that the hard scales $`t`$ and the evolution effects related to running of $`t`$ should be process-dependent. The Wilson coefficients are $`a_1=C_2+{\displaystyle \frac{C_1}{N_c}},a_{4(6)}=C_{4(6)}+{\displaystyle \frac{C_{3(5)}}{N_c}}+{\displaystyle \frac{3}{2}}e_q\left(C_{10(8)}+{\displaystyle \frac{C_{9(7)}}{N_c}}\right),`$ (21) for the tree and the $`(VA)(VA)`$ penguins, respectively, $`N_c`$ being the number of colors. $`F_{e(a)6}^P`$ has a different integrand from $`F_{e(a)4}^P`$, reflecting the different helicity structures. The factors $`r_\pi `$ and $`r_K`$, $`r_\pi ={\displaystyle \frac{m_{0\pi }}{M_B}},m_{0\pi }={\displaystyle \frac{M_\pi ^2}{m_u+m_d}};r_K={\displaystyle \frac{m_{0K}}{M_B}},m_{0K}={\displaystyle \frac{M_K^2}{m_s+m_d}},`$ (22) are associated with the normalizations of the pseudoscalar wave functions $`\varphi ^{}`$, where $`m_u`$, $`m_d`$, and $`m_s`$ are the current quark masses of the $`u`$, $`d`$ and $`s`$ quarks, respectively, and $`M_\pi `$ and $`M_K`$ the pion and kaon masses, respectively. The pseudovector and pseudoscalar pion wave functions $`\varphi _\pi `$ and $`\varphi _\pi ^{}`$ are defined in terms of matrix elements of nonlocal operators $`0|\overline{d}\gamma ^{}\gamma _5u|\pi `$ and $`0|\overline{d}\gamma _5u|\pi `$, respectively. The kaon wave functions $`\varphi _K`$ and $`\varphi _K^{}`$ possess similar definitions. We employ the following set of meson wave functions as an illustrative example: $`\varphi _B(x,b)`$ $`=`$ $`N_Bx^2(1x)^2\mathrm{exp}\left[{\displaystyle \frac{1}{2}}\left({\displaystyle \frac{xM_B}{\omega _B}}\right)^2{\displaystyle \frac{\omega _B^2b^2}{2}}\right],`$ (23) $`\varphi _\pi (x)`$ $`=`$ $`{\displaystyle \frac{3}{\sqrt{2N_c}}}f_\pi x(1x)[1+0.8(5(12x)^21)],`$ (24) $`\varphi _K(x)`$ $`=`$ $`{\displaystyle \frac{3}{\sqrt{2N_c}}}f_Kx(1x)[1+0.51(12x)+0.3(5(12x)^21)],`$ (25) $`\varphi _\pi ^{}(x)`$ $`=`$ $`{\displaystyle \frac{3}{\sqrt{2N_c}}}f_\pi x(1x),\varphi _K^{}(x)={\displaystyle \frac{3}{\sqrt{2N_c}}}f_Kx(1x),`$ (26) with the shape parameter $`\omega _B=0.3`$ GeV and the decay constants $`f_\pi =130`$ MeV and $`f_K=160`$ MeV. The normalization constant $`N_B`$ is related to the $`B`$ meson decay constant $`f_B=190`$ MeV via $`\varphi _B(x,b=0)𝑑x=f_B/(2\sqrt{2N_c})`$. $`\varphi _K`$ is derived from QCD sum rules . All other meson wave functions and $`f_B`$ are determined from the data of the $`BD\pi `$, $`\pi \pi `$ decays and of the pion form factor . Note that we have included the intrinsic $`b`$ dependence for the heavy meson wave function $`\varphi _B`$ but not for the light meson wave functions $`\varphi _\pi `$ and $`\varphi _K`$. It has been shown that the intrinsic $`b`$ dependence of the light meson wave functions, resulting in only 5% reduction of the predictions for the form factor $`F^{B\pi }`$, is not important . We do not distinguish the pseudovector and pseudoscalar components of the $`B`$ meson wave functions under the heavy quark approximation. $``$ Is PQCD legitimate ? We show that PQCD allows us to compute two-body decay amplitudes by examining where dominant contributions to the form factor $`F^{B\pi }`$ come from. Figure 3 displays the fractional contribution as a function of $`\alpha _s(t)/\pi `$. It is observed that 97% of the contribution arises from the region with $`\alpha _s(t)/\pi <0.3`$. Therefore, our PQCD results are well within the perturbative region. This analysis implies that $`H([x],[𝐛],M_B)`$ is dominated by short-distance contributions, contrary to the viewpoint of Beneke, Buchalla, Neubert, and Sachrajda (BBNS) in , and that the physical picture we have given for what is happening in Fig. 1 is indeed valid. $``$ Fat penguins in PQCD: Let us have a careful look at the matrix elements of the penguin operators. It is noticed that unlike $`C_2`$, $`C_4`$ and $`C_6`$ have a steep $`\mu `$ dependence. In FA, amplitudes depend on the matching scale. Normally, it is taken to be $`m_b/2`$, $`m_b`$ being the $`b`$ quark mass, but there is no theoretical basis for this choice. One of the main advantages in PQCD is that it provides a prescription for choices of the hard scale $`t`$: $`t`$ should be chosen as the virtuality of internal particles in a hard amplitude in order to decrease higher-order corrections. A good fraction of contributions then come from $`t<m_b/2`$, and penguin contributions are enhanced. Numerically, this enhancement is given by: $$\frac{(F_{e6}^P)_{\mathrm{PQCD}}}{(F_{e6}^P)_{\mathrm{FA}}}=1.6,\frac{(F_{e4}^P)_{\mathrm{PQCD}}}{(F_{e4}^P)_{\mathrm{FA}}}=1.4,\frac{(F_e)_{\mathrm{PQCD}}}{(F_e)_{\mathrm{FA}}}=1.0,$$ (27) where $`(F)_{FA}`$ represent the form factors evaluated in PQCD but with the Wilson coefficients $`C(t)`$ set to $`C(M_B/2)`$. Equation (27) shows that penguin contributions are dynamically fattened by about 50%, and that the tree amplitudes from $`O_{1,2}`$ remains invariant. Other sources of penguin enhancement are referred to . The enhancement due to the increase of $`C_6(t)`$ with decreasing $`t`$ makes us worry that the contribution from the small $`t`$ region may be important. This will invalidate the perturbative expansion of $`H([x],[𝐛],M_B)`$. As a check, we examine the fractional contribution to $`F_{e6}^P`$ as a function of $`\alpha _s(t)/\pi `$. The results, similar to Fig. 3, indicate that about 90% (80%) of the contribution comes from the region with $`\alpha _s(t)/\pi <0.3`$ (0.2). Therefore, exchanged gluons are still hard enough to guarantee the applicability of PQCD. We emphasize that the penguin enhancement is crucial for the simultaneous explanation of the $`BK\pi `$, $`\pi \pi `$ data using a unitarity angle $`\varphi _390^o`$ . It has been shown that a simultaneous understanding of the data $`R=\mathrm{Br}(B^0K^\pm \pi ^{})/\mathrm{Br}(B^\pm K^0\pi ^{})1.0`$ and $`\mathrm{Br}(B^0\pi ^\pm \pi ^{})4.3\times 10^6`$ is difficult in FA . The former indeed leads to $`\varphi _390^o`$. However, the latter leads to $`\varphi _3130^o`$, even if $`m_0`$ is streched to $`m_04`$ GeV corresponding to $`m_d=2m_u=3`$ MeV. $``$ Imaginary annihilation penguins: There has been a widely spread folklore that the annihilation diagrams give negligible contribution due to helicity suppression, just as in $`\pi e\overline{\nu }`$ decay. That is, a left-handed massless electron and a right-handed antineutrino can not fly away back to back because of angular momentum conservation. However, this argument does not apply to $`F_{a6}^P`$. A left-handed quark and a left-handed antiquark, for which helicities are dictated by the $`O_6`$ operator, can indeed fly away back to back . These behaviors have been reflected by Eqs. (10) and (13): Eq. (10) vanishes exactly, if the kaon and pion wave functions are identical, while the two terms in Eq. (13) are constructive. Numerical results in Table I show that the strong phase associated with $`F_{a6}^P`$ is nearly $`90^{}`$. The large absorptive part arises from cuts on the intermediate state $`(s\overline{d})`$ in the decay $`\overline{B}^0s\overline{d}K^{}\pi ^+`$ shown in Fig. 2. The intermediate state $`(s\overline{d})`$ can be regarded as being highly inelastic, if expanded in terms of hadron states. On the issue of FSI, Suzuki has argued that the invariant mass of the $`s\overline{d}`$ pair in Fig. 2 is of order $`(\mathrm{\Lambda }_{\mathrm{QCD}}M_B)^{1/2}1.2`$ GeV . Hence, the $`BK\pi `$ decays are located in the resonance region and their strong phases are very complicated. We have computed the average hard scale of the $`BK\pi `$ decays, which is about 1.4 GeV, in agreement with the above estimate. Since the outging $`s\overline{d}`$ pair should possess an invariant mass larger than 1.4 GeV, the processes are in fact not so close to the resonance region. We could interpret that the $`BK\pi `$ decays occur via a six-fermion operator within space smaller than $`(1/1.4)`$ GeV<sup>-1</sup>. Though they are not completely short-distance, the fact that over 90% of contributions come from the $`x`$-$`b`$ phase space with $`\alpha _s(t)/\pi <0.3`$ allows us to estimate the decay amplitudes reliably. We believe that the strong phases can be computed up to about 20% uncertainties, which result in 30% errors in predictions for CP asymmetries. $``$ Br$`(BK\pi )`$: We present PQCD results of various $`BK\pi `$ branching ratios in Table II, which are well consistent with the CLEO data . These results are meant to be an example for a representative parameter set such as the wave functions in Eqs. (23)-(26), which are determined from the best fit to the data of the $`BD\pi `$, $`\pi \pi `$ and of the pion form factor, and $`m_{0\pi }=1.4`$ GeV and $`m_{0K}=1.7`$ GeV . When all other two-body decay modes are considered, we shall present an exhaustive study of the entire parameter space allowed by data uncertainties. $``$ Comparision with the BBNS approach: Here we compare our approach to exclusive nonleptonic $`B`$ meson decays with the BBNS approach . The differences between the two approaches are briefly summarized below. For more details, refer to . As stated before, the PQCD theory for exclusive processes was first formulated by Brodsky and Lepage . This formalism has been criticized by Isgur and Llewellyn-Smith , since involved perturbative evaluations, based on expansion in terms of a large coupling constant in the end-point region of momentum fractions, are not reliable. Li and Sterman pointed out that Sudakov resummation of large logarithms associated with parton transverse momenta, which was worked out by Botts and Sterman , causes suppression in the end-point region. This suppression is strong enough to render PQCD analyses self-consistent at energy scales of few GeV. The above approach was then extended by Li and his collaborators to exclusive $`B`$ meson decays in the heavy meson limit , where Sudakov suppression, cutting off the infrared singularity in heavy-to-light transition form factors , is even more important. Our calculation of two-body $`B`$ meson decays has followed the formalism developed by the above authors. Therefore, the $`B`$-to-light-meson transition form factors at maximal recoil are calculable in PQCD. In the BBNS approach, because Sudakov suppression is not considered, the transition form factors are not calculable and must be treated as inputs. Another crucial difference is that in the PQCD formalism annihilation diagrams are of the same order as factorizable diagrams in powers of $`1/M_B`$, which are both $`O(1/(M_B\mathrm{\Lambda }_{\mathrm{QCD}}))`$. The BBNS approach follows FA, in which it has been assumed that factorizable contributions, being $`O(1/\mathrm{\Lambda }_{\mathrm{QCD}}^2)`$, are leading, and all other contributions such as annihilation and nonfactorizable diagrams, being $`O(1/(M_B\mathrm{\Lambda }_{\mathrm{QCD}}))`$, are next-to-leading. Factorizable and nonfactorizable contributions are considered by BBNS, but annihilation are not. The difference is again traced back to the inclusion of parton transverse momenta and of the Sudakov form factor in our calculation. With Sudakov suppression, gluon exchanged in factorizable diagrams are as hard as those in annihilation diagrams. Because of parton transverse momenta, the internal particles in hard amplitudes may go onto the mass shell at nonvanishing momentum fractions . As a consequence, annihilation diagrams lead to large imaginary contributions, whose magnitudes are comparable to factorizable ones, and to large CP asymmetries in the $`BK\pi `$ decays. We emphasize that annihilation diagrams are indeed subleading in the PQCD formalism as $`M_B\mathrm{}`$. This can be easily observed from the hard functions in Eqs. (15) and (17). When $`M_B`$ increases, the $`B`$ meson wave funciton in Eq. (23) enhances contributions to $`h_e`$ from smaller momentum fraction $`x_1`$ as expected. However, annihilation amplitudes proportional to $`h_a`$, being independent of $`x_1`$, are relatively insensitive to the variation of $`M_B`$. Hence, factorizable contributions become dominant and annihilation contributions are subleading in the $`M_B\mathrm{}`$ limit. For $`M_B5`$ GeV, our calculaiton shows that these two types of contributions are comparable. We have shown that PQCD allows us to compute matrix elements of various four-quark operators. While FA gives reliable estimates for $`O_{1,2}`$, since their Wilson coefficients are nearly constant in the hard scale $`t`$, matrix elements of the penguin operators are another story. We have observed that PQCD results are larger than FA results by about 50% for the penguin operators, because of the $`t`$ dependence of the Wilson coefficients. With the penguin enhancement in PQCD, the CLEO data of the $`BK\pi `$, $`\pi \pi `$ branching ratios can be understood in a more self-consistent way. We have also pointed out that penguin annihilation diagrams are not negligible as claimed in FA. In fact, they contribute large strong phases, which are essential for predictions of CP asymmetries. ###### Acknowledgements. This work was supported in part by Grant-in Aid for Special Project Research (Physics of CP Violation) and by Grant-in Aid for Scientific Exchange from the Ministry of Education, Science and Culture of Japan. The works of H.N.L. and Y.Y.K. were supported by the National Science Council of R.O.C. under the Grant Nos. NSC-89-2112-M-006-004 and NSC-89-2811-M-001-0053, respectively. We thank H.Y. Cheng, M. Kobayashi, and members of our PQCD group for helpful discussions. Figure Captions Fig. 1: (a) Factorizable and (b) nonfactorizable tree or penguin contributions. Fig. 2: (a) Factorizable and (b) nonfactorizable annihilation contributions. A cut on the $`s\overline{d}`$ quark lines corresponds to the absorptive part. Fig. 3: Fractional contribution to the $`B\pi `$ transition form factor $`F^{B\pi }`$ as a function of $`\alpha _s(t)/\pi `$.
warning/0004/astro-ph0004306.html
ar5iv
text
# Resolving the radio nebula around 𝛽 Lyrae ## 1 Introduction $`\beta `$ Lyrae ($`3.4m_V4.4`$, $`P=12.91^\mathrm{d}`$) is a non-degenerate, semi-detached interacting binary system, located at a distance of 270$`\pm 38`$ pc (Perryman et al.perryman (1997)). The primary component is a B6-8 II star in contact with its Roche lobe resulting in rapid mass transfer ($`5\times 10^5M_{\mathrm{}}\mathrm{yr}^1`$) to its massive, unseen companion (Harmanec Harmanec90 (1990)). The nature of the secondary has been the subject of much speculation, from a flattened B5 star to a black hole with accretion disk. A B0V star embedded in a geometrically and optically thick accretion disk or a circumstellar shell, is currently the more favoured hypothesis (Hubeny $`\&`$ Plavec Hubeny (1991), Harmanec Harmanec92 (1992)). The presence of circumstellar plasma surrounding both components is indicated by stationary optical and UV emission lines (Batten & Sahade Batten (1973); Hack et al. Hack (1975)) as well as from the analysis of UV light curves (Kondo et al. Kondo (1994)). Although $`\beta `$ Lyrae has been observed extensively at other wavelengths for decades, radio observations are still too scarce to form a clear picture of its radio properties. The source was firstly detected by Wade & Hjellming (wade (1972)) at 2.7 and 8.1 GHz. During a successive monitoring program (Gibson 1975) the source always exhibited a spectrum consistent with thermal radio source with a radius $`50`$ AU, if a temperature of typical of an H II region ($`10^4`$ K) is assumed. However, Wright & Barlow (Wright (1975)) have noted that the observed slope of the radio spectrum ($`\alpha `$=0.96) is intermediate between that expected for a simple H II region and a stellar wind, implying that the physics underlying the radio emission of $`\beta `$ Lyrae is more complicated than assumed in either of these models. In an attempt to distinguish the various components of $`\beta `$ Lyrae which may be contributing to the radio emission, we obtained high-spatial resolution radio maps of $`\beta `$ Lyrae using MERLIN. In this paper we describe the results of these observations and briefly discuss their implications. ## 2 Observations and Results We observed $`\beta `$ Lyrae at 6 cm (4.994 GHz) using the Multi-Element Radio Linked Interferometer Network (MERLIN)<sup>1</sup><sup>1</sup>1 MERLIN is a national facility operated by the University of Manchester at the Nuffield Radio Astronomy Laboratories, Jodrell Bank, on behalf of the Particle Physics and Astronomy Research Council (PPARC) array on 1996 December 8, 9 and 16, from 04:00 to 22:00 UT. The observations were performed in phase referencing mode, which is the standard MERLIN observing mode for weak sources. This technique allows a better calibration of phase variation by observations of a bright phase calibrator interleaved with the observations of the target source. The phase calibrator 1846+322 was used and the flux density scale was determined by observing 3C286, whose flux at 4.994 GHz is assumed to be 7.086 Jy. Data were edited and amplitude calibrated using the OLAF software package. The phase calibration, mapping process and successive analysis were performed using the NRAO, Astronomical Image Processing System. To achieve the highest possible signal-to-noise ratio the mapping process was performed using natural weighting and the CLEAN algorithm was used to extract information as close as possible to the theoretical noise limit. We detected the source at all three epochs, corresponding to three different orbital phases. All three maps revealed a quite extended source, elongated slightly in the N-S direction. The noise level in the maps measured from several areas of blank sky is $`6\times 10^5`$Jy/beam, which is consistent with the expected theoretical noise. There are no significant differences between the maps observed at each epoch and so we combined the data from the three observing runs to form the radio map shown in Fig. 1. The source morphology is dominated by a compact structure, aligned approximately N-S, probably ($`<3\sigma `$ level), embedded in a more diffuse nebula. The profile of the radio nebula along its principal axes is shown in Fig. 2. From the figure is evident that most of the 6 cm emission comes from the central region, with negligible contribution from the more diffuse nebula, thus proving that a two dimensional Gaussian fit (dotted line) to the central part gives a fair description of the radio source. The position, flux density and angular size of the source are thus derived by fitting the two dimensional Gaussian brightness distribution to the map. The values and their uncertainties, derived following Fomalont (Fomalont (1989)), are given in Table 1. The optical position of $`\beta `$ Lyrae reported by Hipparcos, allowing for the proper motion, is also given. Differences between radio and optical positions ($`\mathrm{\Delta }\alpha =0.075\mathrm{},\mathrm{\Delta }\delta =0.061\mathrm{}`$) may be probably related to the overall position error to be associated to the source used as phase calibrator, reported to be of the order pf $`0.1\mathrm{}`$ (Wilkinson et al. 1998). MERLIN is sensitive to structures up to 3.5″ at 6 cm. No structure is seen outside of the nebula up to these scales, which would suggest that all of the flux from the nebula is detected in these maps. The difference between the flux measured here and previous measurements (e.g. Mutel et al. Mutel (1985); Leone, Trigilio & Umana Leone (1994)) is not surprising as $`\beta `$ Lyrae is known to show variability at radio wavelengths. ## 3 Discussion and Conclusions The radio observations here discussed are direct confirmation of a large plasma cloud surrounding $`\beta `$ Lyrae due to mass loss from the system. The angular size of the nebula corresponds to a linear size of approximately 40 AU, much larger than the binary system itself ($`0.5`$ AU). As the source has been resolved we can derive a brightness temperature $`T_\mathrm{B}`$ as follows: $$T_\mathrm{B}=\frac{1.36\times 10^6S\lambda ^2}{\theta _1\theta _2}\mathrm{K}=\mathrm{11\hspace{0.17em}000}\pm 700\mathrm{K}$$ (1) where $`S`$ is the flux in mJy at the 5 GHz, $`\lambda `$ is the observing wavelength in cm and $`\theta _1`$ and $`\theta _2`$ are the angular size of source in milliarcseconds. The derived brightness temperature implies a thermal origin for the radio emission. It can be argued that the source has, in reality, a much more complex morphology that the present MERLIN observations can not resolve. If, for example, most of the flux is coming from two small components, with thus a much higher brightness temperature, the simple one-component model, we are assuming, will be no suitable. Still, in the former case, a higher brightness temperature would imply a non-thermal mechanism operating in the system. The presence of non thermal component has been inferred in about 30$`\%`$ of the early type radio emitting stellar systems (Bieging et al. 1989). All of these stars always exhibited spectral indices $`\alpha 0`$ ($`S_\nu \nu ^\alpha `$), while past (Gibson, 1975) and more recent (Umana et al. 2000) radio observations of $`\beta `$ Lyrae point out a thermal spectrum ($`\alpha >0`$) for the system. Jameson & King (Jameson (1978)) modelled the radio source associated to $`\beta `$ Lyrae as a H II region. However, there is ample evidence, from both optical (Etzel & Meyer, 1983) and ultraviolet spectroscopy (Mazzali et al. 1992), of the presence of stellar wind in one or both components of the system, which can be responsible of the observed radio emission. Mazzali et al. (Mazzali (1992)) have developed a two-wind model to explain the peculiar UV spectrum of $`\beta `$ Lyrae , consisting of ”superionized” resonance lines, typical of hot-star winds, plus lines of moderately ionized species. The observed ultraviolet emission features can be modelled allowing the coexistence of two winds with different dynamical properties: a fast, but tenuous wind, associated to the hotter secondary component ($`\dot{M_\mathrm{h}}=5\times 10^8M_{\mathrm{}}\mathrm{yr}^1`$, $`v_\mathrm{h}=1470\mathrm{km}\mathrm{s}^1`$) and a denser, but slower wind, associated to the cooler B6-8 II primary ($`\dot{M_\mathrm{c}}=7.7\times 10^7M_{\mathrm{}}\mathrm{yr}^1`$, $`v_\mathrm{c}=390\mathrm{km}\mathrm{s}^1`$). Standard formulas for thermal radio emission from an expanding wind have been derived by Panagia & Felli (Panagia75 (1975)) and Wright & Barlow (Wright (1975)). They showed that the observed radio flux ($`S`$) is related to the dynamical parameters of the wind through the relation: $$S\left(\frac{\dot{M}}{v_{\mathrm{}}}\right)^{4/3}$$ (2) where $`\dot{M}`$ and $`v_{\mathrm{}}`$ represent the mass-loss rate and the terminal velocity of the wind. We may ask under which conditions the thermal radio emission from the winds of the two early-type components of a binary system has the same radio properties as a single symmetric wind and thus the relation (2) can be applied. The effects of binarity on thermal radio emission from early-type systems have been recently studied by Stevens (1995). Effects of possible interactions between winds, such as extra source of emission due to the shocked gas, are mostly function of the momentum ratio of the two winds and they would be negligible in systems, where one component has a dominant wind. The two-wind model outlined by Mazzali et al. (1992) foresees a case of a system when only one wind is dominant. Moreover, since the winds of the two stars will contribute to the measured flux according to the ratio of the relative factors $`(\frac{\dot{M}}{v_{\mathrm{}}})`$, allowing for difference in mass-loss and wind terminal velocity, this indicates that less than 0.5 % of the observed flux is attributable to the BOV star’s wind. There is a class of stellar objects, the symbiotics, whose radio emission is interpreted in terms of a binary model, and where one of the two components has a dominant wind. Taylor & Seaquist (Taylor84 (1984)) attributed the observed radio emission to the stellar wind of the cool component of the system, which is ionized by the ultraviolet flux of the hot companion. They also showed that the geometry of the ionized radio emitting region of the intersystem material is determined by a single parameter, which is function of the physical parameters of the binary: separation ($`a`$), Lyman continuum luminosity of the hot component ($`L_{uv}`$), mass loss rate ($`\dot{M}`$) and velocity ($`v`$) of the dominant wind. If the wind is completely ionized, the ionized region shows the same radio properties as a spherically symmetric ionized wind. In the following we will check if this scenario is suitable also for $`\beta `$ Lyrae . By applying equation 14 from Taylor & Seaquist (Taylor84 (1984)) and assuming a binary separation of 61 solar radii (Harmanec, 1990), we obtain that the wind of the B6-8 II is completely ionized if $$L_{uv}5.8\times 10^{62}(\frac{\dot{M}}{v})^21.8\times 10^{45}\mathrm{photons}\mathrm{s}^1$$ (3) where the values of $`\dot{M}`$ and $`v`$ for the B6-8 II as determined by Mazzali et al. (1992) have been adopted. This requirement is well satisfied by the flux of Lyman continuum photons of the bright secondary component B0V, $`L_{uv}=4.26\times 10^{47}\mathrm{photons}\mathrm{s}^1`$ (Panagia Panagia73 (1973)), on the contrary of the B6-8 II, and we can safely apply relation (2). This formula has been derived for the case of spherically symmetric winds. However, Schmid-Burgk (Burgk (1982)) showed that the mass loss rate derived from radio flux densities is quite insensitive to source geometry and only in very extreme cases of deviation from spherical symmetry a geometry-dependent correction factor must be applied. Thus even if the source morphology appears to be slightly elongated (axis ratio 1.45), we can use them quite confidently. We, therefore, can estimate the mass loss by using the relation: $$\dot{M}=6.7\times 10^4v_{\mathrm{}}S_{6\mathrm{c}\mathrm{m}}^{3/4}D_{\mathrm{kpc}}^{3/2}(\nu \times g_{\mathrm{ff}})^{0.5}M_{\mathrm{}}\mathrm{yr}^1$$ (4) where full ionization and cosmic abundances have been assumed, $`v_{\mathrm{}}`$ is the terminal velocity of the wind, $`S_{6\mathrm{c}\mathrm{m}}`$ the observed flux density at 6 cm in mJy, $`D_{\mathrm{kpc}}`$ the distance of the system, in kpc. $`g_{\mathrm{ff}}`$ represents the free-free Gaunt factor that, following Leitherer & Robert (1991), can be approximated with: $$g_{\mathrm{ff}}=9.77(1+0.13\mathrm{log}\frac{T^{3/2}}{\nu })$$ (5) where $`T`$, in Kelvin, is the wind temperature. By assuming the stellar wind velocity of the B6-8 II component ($`v400\mathrm{km}\mathrm{s}^1`$, Mazzali et al. Mazzali (1992)) and constant temperature equal to the brightness temperature, derived by the present observations ($`T=T_\mathrm{B}`$), we obtain a mass loss rate of $$\dot{M}=5.6\times 10^7M_{\mathrm{}}\mathrm{yr}^1.$$ (6) This result is close to the value obtained by Mazzali et al. (Mazzali (1992)) for the B6-8 II component by modelling UV spectral lines, i. e. $`7.16\times 10^7M_{\mathrm{}}\mathrm{yr}^1`$ and indicates that the wind associated to the primary is predominant in the replenishment of the circumsystem material and constitutes the bulk of the system’s mass loss. $`\beta `$ Lyrae is approaching the end of a phase of strong mass transfer between components, which started 26000 years ago (De Greve & Linnell Degreve (1994)). The presence of such a nebula is the proof that the system had evolved in a non conservative way, i. e., with a fraction of the material from the loser lost by the system. This was already foreseen by the evolutionary studies of De Greve & Linnell (Degreve (1994)) since their conservative solution lead to several discrepancies with the observed characteristics of the system. If we assume that during this phase of mass transfer the mass-loss rate was constant ($`5.6\times 10^7M_{\mathrm{}}\mathrm{yr}^1`$) we estimate that almost 0.015 $`M_{\mathrm{}}`$ should have been lost by the system. High resolution optical interferometry combined with extensive photometry and spectroscopy has led Harmanec et al. (Harmanec96 (1996)) to conclude that $`\beta `$ Lyrae contains jet-like structures perpendicular to the orbital plane which are responsible for the bulk of the H$`\alpha `$ and He I 6678 emission. This scenario, already proposed by Harmanec (Harmanec92 (1992)), is further supported by recent spectropolarimetric observations of Hoffman et al. (Hoff (1998)), who suggested the presence of a bipolar outflow in $`\beta `$ Lyrae , perpendicular to the orbital plane (P.A. = $`164\mathrm{°}`$) and probably associated to the accretion disk around the mass-gaining component. The radio image presented in this work gives a hint of structure within the nebula, even if the obtained angular resolution did not allow us to trace the structure of the star’s circumstellar matter at very small scale and thus to verify the association of the radio emitting features to the bipolar optical jets. One piece of information on this direction, provided by the present results, is that the radio nebula is almost exactly aligned with these optical jet-like structures (P.A. = $`156.5\mathrm{°}\pm 4\mathrm{°}`$). It may be more accurate to picture the outflow as a stellar wind collimated by the thick accretion disk. The appearance of the circumbinary material would thus consist of a more extended component, due to the mass-loss from the system as a whole, plus an inner more compact component, related to the accretion disk created in the process of mass transfer from the B star to the hidden companion. To establish the feasibility of this scenario high resolution radio observations, at frequencies higher than 5 GHz, aimed to probe the morphology of the inner structure of the radio nebula are necessary. This observations, by allowing to look more inside the binary’s outflow, would provide further insight into the formation of the circumstellar matter and are of particular importance in understanding the connection of the observed radio nebula to the bipolar optical jets reported by Harmanec et al. (Harmanec96 (1996)) and Hoffman et al. (Hoff (1998)). ###### Acknowledgements. We wish to thank the MERLIN staff for conducting the observations and, in particular, Dr. T. Muxlow for its help in the first stage of data reduction and some helpful suggestions.
warning/0004/astro-ph0004134.html
ar5iv
text
# A Dynamical Solution to the Problem of a Small Cosmological Constant and Late-time Cosmic Acceleration ## Abstract Increasing evidence suggests that most of the energy density of the universe consists of a dark energy component with negative pressure, a “cosmological constant” that causes the cosmic expansion to accelerate. In this paper, we address the puzzle of why this component comes to dominate the universe only recently rather than at some much earlier epoch. We present a class of theories based on an evolving scalar field where the explanation is based entirely on internal dynamical properties of the solutions. In the theories we consider, the dynamics causes the scalar field to lock automatically into a negative pressure state at the onset of matter-domination such that the present epoch is the earliest possible time, consistent with nucleosynthesis restrictions, when it can start to dominate. Introduction. Observations of large scale structure, searches for Type Ia supernovae, and measurements of the cosmic microwave background anisotropy all suggest that the universe is undergoing cosmic acceleration and is dominated by a dark energy component with negative pressure. The dark energy may consist of a cosmological constant (vacuum density) or quintessence, such as a scalar field with negative pressure. In either case, a key challenge is the “cosmic coincidence” problem: Why is it that the vacuum density or scalar field dominates the universe only recently? Until now, either cosmic initial conditions or model parameters (or both) had to be tuned to explain the low density of the dark energy component. In this paper, we explore a new class of scalar field models with novel dynamical properties that avoid the fine-tuning problem altogether. A feature of these models is that the negative pressure results from the non-linear kinetic energy of the scalar field, which we call, for brevity, $`k`$-field or $`k`$-essence. (This consideration is inspired by earlier studies of $`k`$-inflation, kinetic energy driven inflation.). As we will show, for a broad class of theories, there exist attractor solutions which determine the equation-of-state of $`k`$-essence during different epochs depending on the equation-of-state of the background. Effectively, the scalar field changes its speed of evolution in dynamic response to changes in the background equation-of-state. During the radiation-dominated epoch, $`k`$-essence is led to be subdominant and to mimic the equation-of-state of radiation. Hence, the ratio of $`k`$-essence to radiation density remains fixed. When the universe enters the dust-dominated epoch, though, $`k`$-essence is unable to mimic the dust-like equation-of-state for dynamical reasons. Instead, the energy decreases rapidly by several orders of magnitude and freezes at a fixed value. After a period (typically corresponding roughly to the current age of the universe), the field overtakes the matter density and drives the universe into cosmic acceleration. Ultimately, the $`k`$-essence equation-of-state slowly relaxes to an asymptotic value between 0 and -1. (The reader may wish to sneak a peek at Fig. 3 which illustrates the behavior in a specific numerical example.) The scenario bears some likeness to the quintessence “tracker models” discussed by Zlatev et al., but is in fact very different from them. For a certain class of tracker potentials, the quintessence scalar field converges to an attractor solution in which the energy density in the quintessence field mimics the equation-of-state of the background (matter or radiation) energy density. However, for tracker potentials, it does not make a difference what the equation-of-state of the background is. Only after the scalar field passes a certain critical value does the quintessence develop a negative pressure, and, then, the energy density becomes fixed. The weakness in this model is that the energy density for which the pressure becomes negative is set by an adjustable parameter, which has to be fine tuned to explain why cosmic acceleration is happening at the present epoch in the history of the universe. The distinctive feature of $`k`$-essence models is that tracking of the background energy density can only occur in the radiation epoch. At the matter-radiation equality, a sharp transition of $`k`$-essence from positive to negative pressure is automatically triggered by dynamics. The $`k`$-essence cannot dominate before matter-radiation equality because it is exactly tracking the radiation background. It also cannot dominate immediately after dust-domination because its energy density necessarily drops several orders of magnitude at the transition to dust-domination. However, since its energy density decreases more slowly than the matter density as the universe expands, $`k`$-essence must dominate not too long thereafter, at roughly the current epoch. The resolution of the cosmic coincidence problem boils down to the fact that we live at the “right time” after matter-radiation equality. As noted above, the remarkable behavior comes at the cost of introducing a non-linear kinetic energy density functional of the scalar field and adjusting it to obtain the desired attractor behavior. This kind of action may describe a fundamental scalar field or be a low-energy effective action. For example, in string and supergravity theories, non-linear kinetic terms appear generically in the effective action describing moduli and massless degrees of freedom (superpartners) due to higher order gravitational corrections to the Einstein action. The attractor behavior of our models relies on certain broad conditions on the form of these terms. Our initial examples are admittedly contrived for the purposes of numerical illustration. A systematic study of model-building will appear in a forthcoming paper, although, having seen here the relatively simple basic principles, the reader should be equipped to explore more attractive and better-motivated forms. Equations. In the theories we consider the Lagrangian density for $`\phi `$ is taken to be $$=\frac{1}{6}R+\frac{1}{\phi ^2}\stackrel{~}{p}_k(X)+_m$$ (1) where $`R`$ is the Ricci scalar, $`X\frac{1}{2}(\phi )^2`$ , $`_m`$ is the Lagrangian density for dust and radiation and we use units where $`8\pi G/3=1`$. The energy density of the $`k`$-field $`\phi `$ is $`\rho _k=(2X\stackrel{~}{p}_{,X}\stackrel{~}{p})/\phi ^2`$; the pressure is $`p_k=\stackrel{~}{p}/\phi ^2`$; and the speed of sound of $`k`$-essence is $`c_s^2=p_{k,X}/\rho _{k,X}`$, where the subscript means derivative with respect to $`X`$. The attractor behavior can be explained most easily by changing variables from $`X`$ to $`y=1/\sqrt{X}`$ and rewriting the $`k`$-field Lagrangian as: $$_k=\stackrel{~}{p}_k(X)/\phi ^2g(y)/\phi ^2y.$$ (2) In this case, the energy density and pressure are $`\rho _k=g^{}/\phi ^2`$ and $`p_k=g/\phi ^2y`$, where prime indicates derivative with respect to $`y`$. The equation-of-state is $$w_kp_k/\rho _k=g/yg^{}$$ (3) and the sound speed is $$c_s^2=\frac{p_k^{}}{\rho _k^{}}=\frac{gg^{}y}{g^{\prime \prime }y^2}.$$ (4) In order to have a sensible, stable theory, we require $`\rho _k>0`$ and $`c_s^2>0`$. These conditions are satisfied if $`g^{}<0`$ and $`g^{\prime \prime }>0`$ in the region where $`p_k^{}`$ is positive. Therefore, a general, convex, decreasing function $`g(y)`$, such as shown in Fig. 1, satisfies these necessary conditions. Using the Friedmann equation: $`H^2=\rho _{tot}=\rho _k+\rho _m`$ ,where $`\rho _m`$ is the energy density of matter (radiation and dust), and the energy conservation equations, $`\dot{\rho }_i=3\rho _i(1+w_k)`$ for the $`k`$-essence $`\left(ik\right)`$ and matter $`\left(im\right)`$ components, we obtain the following equations of motion $`\dot{y}`$ $`=`$ $`{\displaystyle \frac{3}{2}}{\displaystyle \frac{(w_k\left(y\right)1)}{r^{}\left(y\right)}}\left[r\left(y\right)\sqrt{{\displaystyle \frac{\rho _k}{\rho _{tot}}}}\right]`$ (5) $`\left({\displaystyle \frac{\rho _k}{\rho _{tot}}}\right)^.`$ $`=`$ $`3{\displaystyle \frac{\rho _k}{\rho _{tot}}}\left(1{\displaystyle \frac{\rho _k}{\rho _{tot}}}\right)\left(w_mw_k\left(y\right)\right),`$ (6) where $$r(y)\left(\frac{9}{8}g^{}\right)^{1/2}y\left(1+w_k\right)=\frac{3}{2\sqrt{2}}\frac{\left(gg^{}y\right)}{\sqrt{g^{}}},$$ (7) and dot denotes derivative with respect to $`N\mathrm{ln}a.`$ These are the master equations describing the dynamics of $`k`$-essence models. Once some general properties of $`g(y)`$ are specified, the attractor behavior described in the introduction follows from these coupled equations. Dynamics. We are seeking a tracker solution $`y(N)`$ in which the $`k`$-essence equation-of-state is constant and exactly equal to the background equation-of-state, $`w_k(y\left(N\right))=w_m`$, and the ratio $`\rho _k/\rho _{tot}`$ is fixed. Generically, this requires $`y(N)`$ be a constant $`y_{tr}`$ and therefore $`\rho _k/\rho _{tot}=r^2(y_{tr})`$. The last condition can only be satisfied if $`r(y_{tr})`$ is less than unity. Hence, given a convex function such as shown in Fig. 1, we can first identify those ranges of $`y`$ where $`r(y)`$ is greater than unity or less than unity. In ranges where $`r(y)`$ is less than unity, we can seek values of $`y`$ where $`w_k`$ in Eq. (3) is equal to $`w_m`$, the equation-of-state of the matter or radiation. The value of $`y_{tr}`$ changes depending on the epoch and $`w_m`$. These are the attractor solutions. In ranges where $`r(y)`$ exceeds unity, there are no attractor solutions. Although there is no dust attractor, it is quite possible for there to be a radiation attractor. The radiation attractor corresponds to positive pressure, so it can be located only at $`y<y_D`$. Hence, we must have $`g(y)`$ such that $`r(y)`$ is less than unity for some range at $`y<y_D`$ which includes some point $`y_R`$ where $`w_k(y_R)=1/3`$. During the radiation-dominated epoch, the ratio of $`k`$-essence to the total density remains fixed on this attractor and equal to $`\left(\rho _k/\rho _{tot}\right)=r^2(y_R).`$ In Fig. 1, the pressure $`p_k=g/\phi ^2y`$ is positive above the $`y`$-axis, and negative below the $`y`$-axis. The dust equation-of-state $`p_k=0`$ can only be obtained at $`y=y_D`$ where $`g(y)`$ goes through zero. However, this point can be an attractor, only if, the second condition, $`r(y_D)<1,`$ is satisfied. If it so happens that $`r(y_D)>1`$, then there is no dust attractor in the matter-dominated epoch. This is precisely what we want for our scenario, and this is possible for a broad class of functions $`g`$. If $`g`$ possesses a radiation attractor but no dust attractor, what happens at dust-radiation equality? To answer this question let us study the solutions of the master equations Eqs. (5-7) in two limiting cases, when the energy density of $`k`$-essence is either much smaller or much greater than the matter energy density. If $`\rho _k/\rho _m1,`$ one can neglect the last term in the equation (5) and it is obvious that $`y\left(N\right)y_S,`$ where $`y_S`$ satisfies the equation $`r\left(y_S\right)=0`$, is an approximate solution of the equations of motion. The point $`S`$ satisfies $`g(y_S)=g^{}(y_S)y_S`$, so the tangent of $`g`$ at $`y_S`$ passes through the origin, as shown in Fig. 1. Since $`r(1+w_k),`$ the equation of state of $`k`$-essence at $`y_S`$ corresponds to $`w_k\left(y_S\right)1;`$ we call this solution the de Sitter attractor and denote it by S in Fig. 1. From Fig. 1, it is clear that $`y_S`$ nearly always exists for convex decreasing functions $`g`$. We stress that $`y\left(N\right)y_S`$ is an approximate solution to the equations-of-motion only when matter strongly dominates over $`k`$-essence. So, if $`\rho _k`$ during the radiation dominated epoch is significantly less than the radiation density which it tracks, which is both typical and required to satisfy nucleosynthesis constraints, then $`k`$-essence proceeds to the de Sitter attractor immediately after dust-radiation equality. As the transition to dust-domination occurs, $`\rho _k`$ first drops to a small, fixed value, as can be simply understood. Suppose that $`\left(\rho _k/\rho _{tot}\right)_R=r^2\left(y_R\right)=\alpha <10^2`$ during the radiation dominated epoch, where the bound is set by nucleosynthesis constraints. From the equation of state, Eq. (3), we have the relation: $`g(y_R)=g^{}(y_R)y_R/3`$. The condition $`r(y_D)1`$ is required in order to have no dust attractor solution. Combining these relations, we obtain: $$\frac{g_R^{}y_R^2}{g_D^{}y_D^2}\frac{9}{16}\alpha <10^2$$ (8) On the other hand, it is apparent from Fig. 1 that $`g_{}^{}{}_{R}{}^{}>g_{}^{}{}_{D}{}^{}`$, so $`y_Ry_D`$ if $`\alpha 1`$. In particular, the tangent at $`y_D`$ falls below $`g(y_R)`$, so $`g_D^{}(y_Ry_D)y_Dg_D^{}g(y_R)=y_Rg_R^{}/3`$. Using this relation, we obtain $$\frac{y_R}{y_D}\frac{3}{16}\alpha <210^3\text{ and }\frac{g_D^{}}{g_R^{}}\frac{\alpha }{16}<710^4.$$ (9) Since $`\rho _k=g^{}/\phi ^2`$ and $`\left|g^{}\left(y_S\right)\right|\left|g^{}\left(y_D\right)\right|`$ we conclude that after radiation domination, when the $`k`$-field reaches the vicinity of the S-attractor, the ratio of energy densities in $`k`$-essence and dust does not exceed $`(\rho _k/\rho _{tot})_R\times g_D^{}/g_R^{}`$; that is, $`\rho _k/\rho _{dust}<\alpha ^2/16<710^6.`$ Hence, provided $`(\rho _k/\rho _{tot})_R10^2`$ at dust-radiation equality, the $`k`$-essence field loses energy density on its way to the S-attractor down to a value below $`7\times 10^6`$. By definition, the S-attractor is one in which $`w1`$ and the energy density is nearly constant. Hence, once $`\rho _k`$ has reached its small but non-zero value, it freezes. In the further evolution of the universe, the matter density decreases, but the $`k`$-essence energy density remains constant, eventually overtaking the matter density of the universe. Note that, as $`\rho _k`$ approaches $`\rho _m`$, the condition $`\rho _k/\rho _m1`$ is necessarily violated and a new attractor solution is found for the case where $`k`$ -essence itself dominates the background energy density. This attractor is denoted K in Fig. 1. To prove that that the K-attractor exists, we consider the master equations, Eqs. (5-7), in the limit where $`\rho _k/\rho _{tot}1`$. If $`y_K`$ satisfies the equation $`r\left(y_K\right)=1,`$ then $`y\left(N\right)y_K`$ is an approximate solution of the equations of motion. When dust is not a tracker, there always exists a unique attractor $`y_K`$ in the interval $`y_D<y<y_S`$. To prove this, note that, within this interval, the function $`r(y)`$ has a negative derivative. Recall that $`r(y_S)=0`$ (definition of S-attractor) and $`r\left(y_D\right)>1`$ (to avoid a dust attractor). Since $`r\left(y\right)`$ is a monotonically decreasing, continuous function, there exists a unique point $`y_K`$ ($`y_D<y_K<y_S)`$) where $`r(y)`$ becomes equal to unity. At $`y>y_D`$ the pressure of $`k`$-essence is negative. Hence, generically the K attractor, located near $`y_K,`$ describes a universe dominated by a negative pressure component which induces power-law cosmic acceleration. As acceleration proceeds, $`\rho _k`$ increasingly dominates and $`yy_K`$. Following along using Fig. 1, the dynamics can be summarized as follows: $`k`$-Essence is attracted to $`y=y_R`$ during the radiation dominated epoch; at matter-domination, the energy density drops sharply as $`k`$-essence skips past $`y=y_D`$, because there is no dust attractor, and heads towards $`yy_S`$. The energy density $`\rho _k`$ freezes and, after a period, overtakes the matter density. As it does so, $`y`$ relaxes towards $`y_K`$. In this scenario, our current universe would be making the transition from $`y_S`$ to $`y_K`$. All this occurs for generic $`g(y)`$ satisfying broad conditions on its first and second derivatives. If the ratio of $`\rho _k`$ to the radiation density is near the maximum allowed by nucleosynthesis (roughly equipartition initial conditions), the scenario predicts that the $`\rho _k`$ dominates by the present epoch. Numerical results. We have verified these analytic predictions numerically for a wide class of $`g(y)`$. As as a strategy, we look for forms which are roughly linear, $$g\left(y\right)\frac{1}{3}g_R^{}y_R+g_R^{}(yy_R)+O\left(\left(yy_R\right)^2\right)$$ (10) in the vicinity of radiation attractor R and parabolic $$g\left(y\right)\frac{g_D^{}y_D}{y_D^2y_S^2}\left(yy_D\right)\left(y\frac{y_S^2}{y_D}\right)+\mathrm{}$$ (11) in the region $`y_Dyy_S`$. One can easily check, that the points $`y_R`$, $`y_D`$ and $`y_S`$ here are, by construction, the places where the corresponding attractors are located and $`g_R^{},g_D^{}`$ are the derivatives of $`g`$ at the appropriate points. The results are not sensitive to the precise form of $`g`$ that interpolates between these regimes. The main constraints are that the attractor solution have a small ratio of $`\rho _k/\rho _{tot}`$ during the radiation epoch and that there is no dust attractor. For illustrative purposes, we have used these principles to obtain a sample $`g(y)`$ and have transformed it to a form for $`\stackrel{~}{p}(X)`$ in the action, Eq. (1): $$\stackrel{~}{p}_k(X)=2.01+2\sqrt{1+X}+0.03(aX)^3(bX)^4,$$ (12) where $`a=10^5`$ and $`b=10^6`$. For small values of $`X`$, after a field redefinition, this Lagrangian density reduces to one equivalent to a canonical scalar field with exponential potential (curiously, a tracker model). The distinctive dynamical attractor in our models relies deviations from linearity at large $`X`$. The results of a numerical integration are presented in Figs. 2 and 3. We see that $`k`$-essence tracks the radiation ($`w_k1/3`$) during the radiation-dominated epoch. Then, at the onset of matter-domination, $`w_k`$ starts to change and the energy density of $`k`$-essence suddenly drops by several orders of magnitude becoming of the order of 10<sup>-7</sup> times the critical density at red shifts about $`z`$ $`1000`$ as the S-attractor is approached and $`w1`$. At about red shift $`z35,`$ $`\rho _k`$ becomes non-negligible and $`w_k`$ starts to increase, ultimately reaching $`w_k0.77`$ at $`z=0`$. The ratio of the $`k`$-essence energy density to the critical density today is $`\mathrm{\Omega }_k0.74`$. In the future, $`w_k`$ in this model has to approach the value $`0.55`$, corresponding to the K-attractor solution, and the universe will enter the period of power law $`k`$-inflation. Summary. In this paper, we have presented a scenario in which cosmic acceleration occurs late in the history of the universe due to an inevitable sequence of events caused by attractor dynamics. We view the present work as a demonstration of principle; hence, we have emphasized general conditions and an analytic understanding of the scenario. The specific example illustrated in this paper is admittedly complex, composed to illustrate the concept, but we know of no fine-tuning or other requirement that poses a barrier to finding simpler and better-motivated forms. By changes of variables and other simple techniques, one can quickly enlarge the class of actions considered here, which may suggest other types of attractive models. A prediction of $`k`$-essence models that distinguishes them from models based on tracker potentials is that $`w_k`$ is in the process of increasing today from -1 towards its asymptotic value at the K attractor, whereas, for trackers, $`w_k`$ is undergoing a transition from $`w0`$ towards $`w=1`$. A consequence is that the effective value of $`w_k`$ for $`k`$-essence – that is, the $`\mathrm{\Omega }_k`$ weighted average of $`w_k`$ between the present and $`z=1`$ – can be significantly lower than for the tracker potential case, which is bounded below by $`w_{eff}0.75`$. In the numerical example above, the effective $`w_{eff}=0.84`$, for example. The current supernovae data suggest a lower value of $`w_k`$ more consistent with $`k`$-essence. Of course, the $`k`$-essence range for $`w_k`$ is more difficult to distinguish from a cosmological constant ($`w=1`$). In future work, we discuss model-building using further examples and generalizations. We also explore interesting variations of the dynamical scenario with different kinds of attractors, including some which can lead to different long-term future outcomes, such as a return to a pressureless, unaccelerated expansion in the long-term future. This work was initiated at the Isaac Newton Institute for Mathematical Sciences; we thank the organizers and staff of the Institute for their support and kind hospitality. This work was supported in part by the “Sonderforschungsbereich 375-95 für Astro-Teilchenphysik” der Deutschen Forschungsgemeinschaft (C.A.P. & V.M.) and by Department of Energy grant DE-FG02-91ER40671 (Princeton) (P.J.S.).