id
stringlengths
30
36
source
stringclasses
1 value
format
stringclasses
1 value
text
stringlengths
5
878k
no-problem/9906/cond-mat9906296.html
ar5iv
text
# Positive Magnetoresistance of Composite Fermions in Laterally Modulated Structures \[ ## Abstract Adopting the mean-field composite fermion picture, we describe the magneto-transport properties of a two-dimensional electron gas with laterally modulated density around filling factor $`\nu =1/2`$. The occurrence of a strong positive magnetoresistance at low effective magnetic fields as well as Weiss oscillations, which were observed in recent experiments in such systems, can be explained within a semi-classical Boltzmann equation approach, provided one goes beyond a second order approximation in the modulation strength. \] Transport properties of strongly correlated two-dimensional electron systems (2DES) in high magnetic fields, where the lowest Landau level is about half filled, seem to be surprisingly well described by the “Composite Fermion” (CF) picture. This model is motivated by a singular Chern-Simons gauge-field transformation that maps the electron system near filling factor $`\nu =1/2`$ onto a metal of quasi-particles (the CFs). They can be interpreted as electrons to which two magnetic flux quanta have been attached. In the usual mean-field approximation, the CFs are considered as having the same density $`n_{\mathrm{CF}}=n_{\mathrm{el}}n`$ as the electrons and as moving in the weak effective magnetic field $`B_{\mathrm{eff}}=B2\mathrm{\Phi }_0n`$, where $`B`$ is the external applied magnetic flux density and $`\mathrm{\Phi }_0=h/e`$ the flux quantum. Halperin, Lee, and Read predicted that, at low temperatures, these CFs fill in momentum space the spin-polarized states within a well defined Fermi circle of radius $`k_\mathrm{F}=\sqrt{4\pi n}`$. Therefore, one expects that transport properties, which are governed by elastic scattering at the Fermi energy, can be calculated without considering explicitly the mutual interaction between the CFs, by analogy with the Landau theory of the electron Fermi liquid in zero and weak magnetic fields. Indeed, interesting features of the magneto-resistance of 2DESs (near $`\nu =1/2`$) in lateral superlattices, such as two-dimensional anti-dot lattices or one-dimensional (1D) superlattices, created either dynamically by the application of surface acoustic waves or statically by surface etching, have been reproduced astonishingly well by calculations describing the CFs as classical, non-interacting particles moving in suitable effective fields. Here, we focus on static 1D density modulations and compare experimental dc-transport measurements with calculations based on the linearized Boltzmann equation (LBE). The purpose of the present letter is to demonstrate that all characteristic features of the magnetoresistance curve near $`B=B_{1/2}2\mathrm{\Phi }_0n`$ can be reproduced in the quasi-classical CF picture, provided one incorporates all modulating fields and solves the LBE beyond the Beenakker-type approximation (BA), which allows an analytical solution and has been employed in previous work. These features are: (i) a pronounced V-shaped minimum near $`B_{1/2}`$, explained by “channeled orbits” which are omitted in the BA, (ii) shoulders or minima related to commensurability effects (Weiss oscillations due to drifting cyclotron orbits), (iii) a steep increase of the resistance with $`|\overline{B}_{\mathrm{eff}}|`$ between the structures caused by commensurability effects and the oscillatory structures due to the fractional quantum Hall effect, and (iv) an asymmetry in the slope and magnitude of these steep resistance flanks with respect to $`\overline{B}_{\mathrm{eff}}=0`$, due to interference effects between the direct electrostatic density modulation and the induced modulation of $`B_{\mathrm{eff}}`$. The material parameters needed to achieve agreement with the experiment around $`\nu =1/2`$ are consistent with those needed to explain the commensurability effects at low magnetic fields ($`B<0.5`$T). Recent work by Mirlin *et al.*, based on an analytical solution of the LBE within the BA for a special model of anisotropic scattering, obtained reasonable results for the features (ii) and (iii), but not for (i) and (iv). Their approximation misses relevant physics at low $`\overline{B}_{\mathrm{eff}}`$, where the CF picture is expected to apply best, and gives a rather poor fit to the experiment for very small $`|\overline{B}_{\mathrm{eff}}|`$ (see Fig 1 below). To describe the resistance for $`B<0.5`$T, we follow Ref. for pure electric modulation, and treat the 2DES as a degenerate Fermi gas of non-interacting particles, with charge $`e`$ and effective mass $`m_{\mathrm{e}l}^{}`$, average density $`\overline{n}`$, Fermi energy $`E_\mathrm{F}^{\mathrm{e}l}=\overline{n}/D_0^{\mathrm{e}l}`$, and density of states $`D_0^{\mathrm{e}l}=m_{\mathrm{e}l}^{}/(\pi \mathrm{}^2)`$. The etching of grooves into the surface of GaAs-heterostructures is assumed to produce an external electrostatic potential energy $`V^{\mathrm{e}xt}(x)=V_0^{\mathrm{e}xt}\mathrm{cos}qx`$ of period $`a=2\pi /q`$ in the plane of the 2DES. It is screened by the 2DES and this leads within the Thomas-Fermi approximation to a potential energy $`V^{\mathrm{el}}(x)=V_0^{\mathrm{el}}\mathrm{cos}qx`$ with $`V_0^{\mathrm{e}l}=V_0^{\mathrm{ext}}/[1+2/(qa_B^{})]`$, where $`a_B^{}=1/[\pi e^2D_0^{\mathrm{e}l}/\kappa ]`$ is the effective Bohr radius. For GaAs, $`\kappa =12.4`$ and $`a_B^{}10`$nm $`a400`$nm. Consequently, the relative modulation strength is $`ϵ_{\mathrm{el}}=V_0^{\mathrm{e}l}/E_\mathrm{F}^{\mathrm{e}l}V_0^{\mathrm{e}xt}\kappa a/(e^2\overline{n})`$, and the modulated electron density is $`n(x)=\overline{n}[1ϵ_{\mathrm{e}l}\mathrm{cos}qx]`$, independent of $`m_{\mathrm{el}}^{}`$. To describe the resistance of CFs, we modify the approach of Ref. . We treat the CF system as a degenerate Fermi gas of non-interacting particles, with charge $`e`$, effective mass $`m_{\mathrm{C}F}^{}`$ , average density $`\overline{n}`$, Fermi energy $`E_\mathrm{F}^{\mathrm{C}F}=\overline{n}/D_0^{\mathrm{C}F}`$, and density of states $`D_0^{\mathrm{C}F}=m_{\mathrm{C}F}^{}/(2\pi \mathrm{}^2)`$, which obeys Newton’s equation, $`m_{\mathrm{C}F}^{}\dot{𝐯}=e[𝐅_{\mathrm{eff}}+𝐯\times (B_{\mathrm{eff}}𝐞_z)]`$, with effective electric and magnetic fields. We use the LBE to calculate the response to an external homogeneous electric field $`𝐄^{(0)}`$. In the absence of $`𝐄^{(0)}`$, $`𝐅_{\mathrm{e}ff}=V(x)/e`$ is determined by the screened modulation potential, which we parametrize as $`V(x)=ϵ_{\mathrm{C}F}E_\mathrm{F}^{\mathrm{C}F}\mathrm{cos}(qx)`$. Since the static Thomas-Fermi screening of the CFs should be equivalent to that of electrons (with $`D_0^{\mathrm{C}F}>D_0^{\mathrm{el}}`$ instead of $`D_0^{\mathrm{e}l}`$), we expect $`ϵ_{\mathrm{C}F}ϵ_{\mathrm{e}l}`$, so that the modulated density $`n(x)=\overline{n}[1ϵ_{\mathrm{CF}}\mathrm{cos}qx]`$ for $`B`$ near $`B_{1/2}`$ is the same as that for small $`B`$ within this approximation. This yields the effective magnetic field $`B_{\mathrm{eff}}(x)=B2\mathrm{\Phi }_0n(x)=\overline{B}_{\mathrm{eff}}+\delta B_{\mathrm{eff}}(x)`$, where $`\delta B_{\mathrm{eff}}(x)=ϵ_{\mathrm{C}F}B_{1/2}\mathrm{cos}qx`$ is the magnetic modulation. The linear response to $`𝐄^{(0)}`$ is carried by CFs at the Fermi edge, $`(m_{\mathrm{C}F}^{}/2)𝐯^2+V(x)=E_\mathrm{F}^{\mathrm{C}F}`$. Due to the translational invariance in $`y`$ direction, the (suitably scaled) zero-temperature distribution function $`\mathrm{\Phi }(x,\phi )`$ depends only on two variables, the position $`x`$ and the polar angle $`\phi `$ of the velocity $`𝐯(x,\phi )=v(x)(\mathrm{cos}\phi ,\mathrm{sin}\phi )`$, with $`v(x)=v_\mathrm{F}[1V(x)/E_\mathrm{F}^{\mathrm{C}F}]^{1/2}`$, $`v_\mathrm{F}=\mathrm{}k_\mathrm{F}^{\mathrm{C}F}/m_{\mathrm{C}F}^{}`$, and $`k_\mathrm{F}^{\mathrm{C}F}=[4\pi \overline{n}]^{1/2}`$. We normalize $`\mathrm{\Phi }(x,\phi )`$ so that the current density $`𝐣(x)`$ reduces to $$𝐣(x)=e^2D_0^{\mathrm{C}F}_\pi ^\pi \frac{d\phi }{2\pi }𝐯(x,\phi )\mathrm{\Phi }(x,\phi ).$$ (1) A current of CFs implies motion of flux tubes and produces the Chern-Simon electric field $$𝐄^{\mathrm{CS}}=(2h/e^2)𝐣\times 𝐞_z,$$ (2) which, in the linearized Boltzmann equation, $$𝒟\mathrm{\Phi }𝒞[\mathrm{\Phi }]=𝐯(x,\phi )\left(𝐄^{(0)}+𝐄^{\mathrm{CS}}(x)\right),$$ (3) adds to the external driving field $`𝐄^{(0)}`$. Therefore a self-consistent solution of Eqs. (1)-(3) is necessary. The drift operator $$𝒟=v(x)\mathrm{cos}\phi _x+\left(\omega _\mathrm{c}+\omega _\mathrm{m}(x)+\omega _\mathrm{e}(x,\phi )\right)_\phi ,$$ (4) contains the average value and the periodic component of the effective magnetic field via $`\omega _\mathrm{c}=e\overline{B}_{\mathrm{eff}}/m_{\mathrm{CF}}^{}`$, and $`\omega _\mathrm{m}(x)=e\delta B_{\mathrm{eff}}(x)/m_{\mathrm{CF}}^{}`$, respectively. The electric modulation enters in $`\omega _\mathrm{e}(x,\phi )=\mathrm{sin}\phi dv/dx`$ and $`v(x)`$ . The collision operator is taken to be the same as for the unmodulated system and is written in the form $$𝒞[\mathrm{\Phi }]=\frac{1}{\tau }_\pi ^\pi \frac{d\phi ^{}}{2\pi }P(\phi ^{}\phi )\left[\mathrm{\Phi }(x,\phi ^{})\mathrm{\Phi }(x,\phi )\right],$$ (5) where the differential cross section is parametrized by a relaxation time $`\tau `$ and a dimensionless kernel $`P(\phi )`$. For actual calculations we take $`P(\phi )=b+(1b)P_p(\phi )`$ with $`0b1`$ and $$P_p(\phi )=[(2^pp!)^2/(2p)!]\mathrm{cos}^{2p}(\phi /2),$$ (6) which has the finite Fourier expansion $`P(\phi )=_{n=0}^p\gamma _n\mathrm{cos}n\phi `$ with $`\gamma _0=1`$, and $`\gamma _n=2(1b)(p!)^2/((p+n)!(pn)!)`$ for $`n1`$. For $`b=1`$ and $`p=0`$, $`P(\phi )`$ describes isotropic scattering. For $`b<1`$ the fraction $`(1b)`$ of the total scattering cross section is due to anisotropic scattering. With increasing $`p`$, $`P_p(\phi )`$ is increasingly stronger peaked in forward direction. To solve Eqs. (1)-(3), we calculate the distribution function of the homogeneous, unmodulated CF system first. This yields the resistivity tensor $`\widehat{\varrho }^\mathrm{h}`$ with components given by $`\varrho _{xx}^\mathrm{h}=\varrho _{yy}^\mathrm{h}=\varrho _0^{\mathrm{CF}}`$ and $`\varrho _{xy}^\mathrm{h}=\varrho _{yx}^\mathrm{h}=\omega _{\mathrm{t}ot}\tau _{\mathrm{t}r}\varrho _0^{\mathrm{CF}}`$, where $`\varrho _0^{\mathrm{CF}}=m_{\mathrm{CF}}^{}/(e^2\overline{n}\tau _{\mathrm{t}r})`$ and $`\tau _{\mathrm{t}r}=\tau /(1\gamma _1/2)`$ is the relevant CF transport scattering time. Although the CFs move in the effective magnetic field $`BB_{1/2}`$, the Hall resistance $`\varrho _{xy}^\mathrm{h}`$ is determined by the cyclotron frequency $`\omega _{\mathrm{t}ot}=\omega _\mathrm{c}+\omega _{1/2}`$ for the total magnetic field $`B`$, with $`\omega _{1/2}=eB_{1/2}/m_{\mathrm{CF}}^{}`$, due to the inclusion of the Chern-Simons electric field. Introducing the mean free path $`\lambda _{\mathrm{CF}}=\tau _{\mathrm{t}r}\mathrm{}k_\mathrm{F}^{\mathrm{CF}}/m_{\mathrm{CF}}^{}`$, we obtain $`\varrho _0^{\mathrm{CF}}=\mathrm{}k_\mathrm{F}^{\mathrm{CF}}/(e^2\overline{n}\lambda _{\mathrm{CF}})`$ and $`\varrho _{xy}^\mathrm{h}=B/(\overline{n}e)`$, so that the resistivity tensor is independent of $`m_{\mathrm{CF}}^{}`$, and the Hall resistance is the same as that of the 2DES. Comparing with the corresponding Drude formulas for the 2DES at low $`B`$, we see that the ratio of the longitudinal resistances, $`\varrho _0^{\mathrm{CF}}/\varrho _0^{\mathrm{el}}=\sqrt{2}\lambda _{\mathrm{el}}/\lambda _{\mathrm{CF}}`$, directly reflects the ratio of the corresponding mean free paths. The macroscopic resistivity tensor $`\widehat{\varrho }`$ of the modulated system, relates the spatial average $`𝐣(x)`$ of the current density $`𝐣(x)`$ to the driving field, $`\widehat{\varrho }𝐣(x)=𝐄^{(0)}`$. Exploiting the solution for the homogeneous system and the continuity equation, which implies that $`j_x(x)`$ for the modulated system is independent of $`x`$, it can be shown that $`\widehat{\varrho }`$ differs from $`\widehat{\varrho }^\mathrm{h}`$ only in its $`xx`$-component. With some formal but exact manipulations (similar to those of ref. ) one finds $$(\varrho _{xx}\varrho _0^{\mathrm{CF}})/\varrho _0^{\mathrm{CF}}=G/[1G/(1+(\omega _{\mathrm{t}ot}\tau _{\mathrm{t}r})^2)].$$ (7) In this equation $$G=\frac{2\tau _{\mathrm{t}r}}{v_\mathrm{F}^2}_\pi ^\pi \frac{d\phi }{2\pi }g(x,\phi )\chi (x,\phi ),$$ (8) where $`g(x,\phi )=v_\mathrm{F}^2(dV/dx)/(2E_\mathrm{F}^{\mathrm{CF}})+\omega _\mathrm{m}v_y`$ is of first order in the modulation strength $`ϵ_{\mathrm{CF}}`$, and $`\chi (x,\phi )`$ is the solution of the modified Boltzmann equation $$𝒟\chi 𝒞[\chi ]=g(x,\phi )+\omega _{1/2}v_y\left(j_x/j_x^\mathrm{h}1\right).$$ (9) The basic differences between these results for the CF system and those for an electron system derive from inclusion of the Chern-Simons electric field in Eq. (3): It is responsible for the occurence of $`\omega _{\mathrm{t}ot}B`$ in the denominator of Eq. (7), instead of $`\omega _\mathrm{c}BB_{1/2}`$, and the last term in the right hand side of Eq. (9). This last term can be expressed in terms of $`G`$, $`j_x/j_x^\mathrm{h}1=G/(1+(\omega _{\mathrm{t}ot}\tau _{\mathrm{t}r})^2)`$, and requires a self-consistent solution of Eqs. (7)-(9). Inserting $`\tau _{\mathrm{t}r}=\lambda _{\mathrm{CF}}/v_\mathrm{F}`$ into Eq. (8) makes $`G`$ and $`\varrho _{xx}`$ independent of $`m_{\mathrm{CF}}^{}`$, as it should be. Numerical solution of Eqs. (7)-(9) for the CF system and of the corresponding equations of Ref. for the electron system allows to calculate the resistance of the modulated sample for strong applied fields $`BB_{1/2}`$ and in the low $`B`$ regime, respectively. To compare with experiment, we need to choose a reasonable set of parameters. Whereas in both $`B`$ regimes the average density $`\overline{n}`$ and the modulation period $`a`$ are the same, and also the modulation strengths $`ϵ_{\mathrm{CF}}ϵ_{\mathrm{el}}`$ should be similar, the parameters characterizing the collision operator (in our model $`\tau `$, $`b`$ and $`p`$) are expected to be different. To reproduce the characteristics of the Weiss oscillations observed at small $`B`$ , we have to assume a very anisotropic differential cross section with a sharp peak in forward direction and a zero-$`B`$ resistance of only a few $`\mathrm{\Omega }`$ ($`\lambda _{\mathrm{el}}50\mu `$m). This is reasonable since these high-mobility samples have a spacer thickness of about $`60`$nm, so that the donors produce a smooth random potential in the plane of the 2DES that scatters the electrons predominantly under small angles. The CFs, on the other hand, are scattered not only by these small-amplitude donor-induced random potential fluctuation, but also – and predominantly – by the large-amplitude random fluctuations of the effective magnetic field that originates from the donor-induced density fluctuation. This results in $`\lambda _{\mathrm{CF}}\lambda _{\mathrm{el}}`$ and a less pronounced forward scattering for the CFs. Figure 1 shows resistivity data for a sample with density $`\overline{n}=1.8210^{11}\mathrm{cm}^2`$ and period $`a=400`$ nm near filling factor $`\nu =1/2`$ . The thick short-dashed line is the experiment of Ref. . The thin dash-dotted line is the theoretical curve using the approximations of Ref. , based on the Beenakker-type approximation (BA) mentioned above. It neglects the modulation effects in the drift operator $`𝒟`$ of the LBE, that causes a resistance correction quadratic in $`ϵ_{\mathrm{CF}}`$. Furthermore, the direct effect of the electric modulation on the CFs was ignored in this fit. Neglecting this direct effect too and using the BA, we can closely reproduce this symmetric curve of Ref. using the same values for $`\overline{n}`$, $`\varrho _0^{\mathrm{CF}}`$, $`ϵ_{\mathrm{CF}}`$, and using $`b=0`$, $`p=6`$ for our model of the scattering cross section, Eqs. (5) and (6). If we include the direct electric modulation in addition to the magnetic one, we obtain the asymmetric thin dotted line. For small $`|\overline{B}_{\mathrm{eff}}|`$, this again closely reproduces the results of Ref. , but at large $`|\overline{B}_{\mathrm{eff}}|`$ the curve shows a distinct asymmetry, similar to the experimental curve, and similar to the asymmetry obtained for mixed electric and magnetic modulations in 2DESs at low $`B`$ . Solving Eqs. (7)-(9) for the same set of parameters but without any approximation of the drift operator, we obtain the thin solid curve of Fig. 1. This shows that approximating the drift operator $`𝒟`$ by the one of the homogeneous system is insufficient for $`|\overline{B}_{\mathrm{eff}}|<0.5`$T. Physically the BA means that the forces due to the modulation fields are neglected against the Lorentz force due to the average magnetic field $`\overline{B}_{\mathrm{eff}}`$. This omits the effect of channeled orbits, i.e. of “snake orbits” along the lines of vanishing $`B_{\mathrm{eff}}(x)`$ which, similar to the “open orbits” in the case of a pure electric modulation, cause a pronounced positive low-field magneto-resistance . Since $`B_{1/2}=15`$T , the effective magnetic modulation has an amplitude $`ϵ_{\mathrm{CF}}B_{1/2}0.45`$T, so that the effect of channeled trajectories should be important for $`|\overline{B}_{\mathrm{eff}}|<0.45`$T, and a second order approximation in $`ϵ_{\mathrm{CF}}`$ is not adequate for the $`\overline{B}_{\mathrm{eff}}`$ values shown in Fig. 1. Comparing the thin solid curve with the experimental one suggests that the pronounced V-shaped resistance minimum at $`\overline{B}_{\mathrm{eff}}=0`$ seen in the experiment is indeed due to CFs on channeled orbits. To achieve a better quantitative agreement, we choose more realistic parameters. First we take $`b=0.1`$ and $`p=2`$, since then our model (6) approximates closely (apart from the $`\phi =0`$ divergence) the differential cross section derived by Aronov *et al.* for random magnetic field scattering. This choice allows a good fit of the resistance near $`\nu =1/2`$ for all samples available to us, indicating the same scattering mechanism in those samples. Then, assuming a density modulation of about 3%, we have to take $`\varrho _0^{\mathrm{CF}}=650\mathrm{\Omega }`$ ($`\lambda _{\mathrm{CF}}1.3a`$), i.e. a larger value than the resistance of the unmodulated reference sample ($`\varrho _0^{\mathrm{CF}}=270\mathrm{\Omega }`$) taken in Ref. . This yields the thick solid curve from the exact solution of the LBE, which reproduces nicely all features of the experimental data. The larger $`\varrho _0^{\mathrm{CF}}`$ value is reasonable since the etching procedure, in addition to the intended modulation, introduces unintended defects which increase the resistance. For the sample discussed so far ($`a=400`$ nm), the mean free path of the CFs is only marginally larger than the modulation period, $`\lambda _{\mathrm{CF}}1.3a`$, so that commensurability effects are visible only as weak shoulders (thick lines of Fig. 1). To see these effects more clearly, we have investigated a sample with a smaller modulation period, $`a=275`$ nm, so that we expect $`\lambda _{\mathrm{CF}}1.9a`$. Experimental and theoretical results (with the same anisotropy parameters $`b`$ and $`p`$ as in Fig. 1) are plotted in Fig. 2. Indeed, the commensurability effects near $`|\overline{B}_{\mathrm{eff}}|=0.5`$T are more pronounced. Again, the low field resistance (inset of Fig. 2) and that near $`\nu =1/2`$ can be reproduced theoretically with a consistent set of parameters. In conclusion, our numerical solution of the linearized Boltzmann equation for composite Fermions without the Beenakker-type approximation used in previous work , and thus the inclusion of the effect of “channeled orbits”, reproduces the typical V-shaped resistance minimum at $`\nu =1/2`$ ($`\overline{B}_{\mathrm{eff}}=0`$) and yields good overall agreement with experiment, consistent with the low-field magnetoresistance oscillations. Including further the superposition of the original electric superlattice and the induced effective magnetic superlattice, we can also reproduce the asymmetry of the resistance curve around $`\overline{B}_{\mathrm{eff}}=0`$. We are grateful to J. Smet for providing the unpublished data of Fig. 2 and for helpful discussions. This work was supported by BMBF Grant No. 01BM622. Fig.1 Zwerschke Fig.2 Zwerschke
no-problem/9906/hep-th9906047.html
ar5iv
text
# Charged False Vacuum Bubbles and the AdS/CFT Correspondence ## I Introduction The conjectured correspondence between large $`N`$ superconformal field theories and string theory in anti-de Sitter backgrounds , provides us with a possible nonperturbative definition of string theory. The formulation is background dependent, since one is always left with the restriction that the geometry is asymptotically anti-de Sitter space. Nevertheless it is possible to use this framework to address some of the longstanding questions in cosmology by considering processes where a black hole forms with an interior resembling a realistic cosmological model. In this paper we take a first step in this direction. We perform a comprehensive analysis of the motion of a charged domain wall that separates an external Reissner-Nordstrøm region of spacetime (with small or vanishing cosmological constant) from an internal de-Sitter region. Our analysis rests on the discontinuity equations of Israel and Kuchar , and is very much in the spirit of that carried out by Blau, Guendelman and Guth, for the case in which the external metric is Schwarzschild (see also ), and by Boulware , for the case in which the internal metric is Minkowski. In the present study, with a charged shell, there exists a much richer spectrum of possibilities for the dynamics of the collapsing shell. We avoid an obstacle encountered in that initial conditions are necessarily singular. We present all possible allowed dynamics for the shell in four dimensions. In the situation considered here, it is impossible to predict, without imposing additional boundary conditions, the ultimate outcome of the interior region of such a composite spacetime , since the singularity theorems of General Relativity predict a singularity to the past of the core region. However, this lack of predictability can be dealt with by appealing to the AdS/CFT correspondence. Taking the point of view that this conjecture is correct, it becomes possible to resolve the singularities that appear and unambiguously determine the dynamics. Thus, we discuss the gravitational systems explored here in the framework of the AdS/CFT correspondence. This study also addresses a number of issues important for topological inflation , where super-heavy magnetic monopoles are argued to provide the seeds for inflation. In addition, when solitons of various different topological charge are present in one theory , it has been suggested that monopole collisions could give rise to inflating core regions . The lack of predictability at the classical level mentioned above is a severe obstacle in these approaches. The structure of the paper is as follows. In the next section we derive the equations of motion of our charged shell. In section III we then discuss the possible solutions to these equations, and present the trajectories on Penrose diagrams. Finally, in section IV we discuss the bubble solutions using the AdS/CFT correspondence and comment on their interpretation at the quantum level. ## II Equations of motion Consider a spherical domain wall in $`n+1`$ spacetime dimensions which separates an interior region of spacetime from an exterior region, described respectively by the static metrics, in spherical polar coordinates $`(r,\mathrm{\Omega })`$, $$ds^2=f_{in(out)}(r)dt^2+f_{in(out)}^1(r)dr^2+r^2d\mathrm{\Omega }_{n1}^2.$$ (1) In particular we will examine the case where the internal region is in a false vacuum state described by de Sitter metric $$f_{in}=1\chi ^2r^2,$$ (2) and the external is described by Reissner-Nordstrøm-anti de Sitter (R-N) metric $$f_{out}=1\frac{2m}{r^{n2}}+\frac{Q^2}{r^{2n4}}+\mathrm{\Lambda }r^2,$$ (3) where $`m`$ is the Schwarzschild parameter, $`Q`$ the charge, and $`\mathrm{\Lambda }`$ the absolute value of the cosmological constant. Further, we will only consider the case $`Qm`$, to preserve the horizon structure of the R-N black hole, and $`\mathrm{\Lambda }`$ much less than any other parameter, to ensure a relevant contribution only in the asymptotic region $`r\mathrm{}`$. We will parameterize the radius $`r`$ of the shell by the proper time $`\tau `$ measured by an observer comoving with the shell. The equations of motion are referred to as junction equations , and in our parameterization reduce to $$s_{in}\sqrt{f_{in}+\dot{r}^2}s_{out}\sqrt{f_{out}+\dot{r}^2}=kr,$$ (4) where $`k`$ is proportional to the tension of the wall, $`s_{in(out)}=\pm 1`$, and a dot represents a derivative with respect to proper time. Squaring (4) twice, we obtain $$\dot{r}^2+V(r)=1,$$ (5) which is the equation for the one dimensional motion of a particle in a potential $$V(r)=\chi ^2r^2\frac{1}{4k^2r^2}\left[(k^2\chi ^2\mathrm{\Lambda })r^2+\frac{2m}{r^{n2}}\frac{Q^2}{r^{2n4}}\right]^2.$$ (6) It can be shown that in regions where $`r`$ is a spacelike coordinate $`s_{in(out)}`$ are positive if the outward normal to the wall is pointing towards increasing radii (see ), and negative if the normal points towards decreasing radii. If we use Kruskal-Szekeres style coordinates $`(u,v)`$ we find that $`s_{in(out)}`$ determine also whether the angle $`\mathrm{arctan}(v/u)`$ increases or decreases as we move along the trajectory (see ). For the external metric the Kruskal-Szekeres coordinates have a rather complicated relationship with $`r`$ and $`t`$. For definiteness we give the explicit relations for $`n=3`$, $`\mathrm{\Lambda }=0`$ in region $`I`$ ($`u>0`$, $`u>|v|`$) when the formulas are still reasonably simple $$u=e^{\gamma r_{}}\mathrm{cosh}(\gamma t),v=e^{\gamma r_{}}\mathrm{sinh}(\gamma t),$$ (7) where we have defined $$r_{}=r+\frac{r_+^2}{r_+r_{}}\mathrm{log}(rr_+)\frac{r_{}^2}{r_+r_{}}\mathrm{log}(rr_{}),\gamma =\frac{r_+r_{}}{2r_{}^2},$$ (8) and where $`r_+`$ and $`r_{}`$ denote the positions of the outer and inner event horizons of the R-N geometry respectively. $`s_{out}=+1`$ implies $`\mathrm{arctan}(v/u)`$ increases along the trajectory, while $`s_{out}=1`$ implies the angle decreases. For the internal de Sitter metric, the Kruskal-Szekeres style coordinates in region $`I`$ ($`u>0`$, $`u>|v|`$) are $$u=\left(\frac{1\chi r}{1+\chi r}\right)^{1/2}\mathrm{cosh}(\chi t),v=\left(\frac{1\chi r}{1+\chi r}\right)^{1/2}\mathrm{sinh}(\chi t).$$ (9) In this case, $`s_{in}=+1`$ implies $`\mathrm{arctan}(v/u)`$ decreases along the trajectory, while $`s_{in}=1`$ implies the angle increases. We now use these relationships to determine the regions of the Penrose diagrams of the internal and external spaces in which the wall evolves. To begin, (4) yields $$s_{in}=+1f_{in}f_{out}+k^2r^2>0,$$ (10) and $$s_{out}=+1f_{in}f_{out}k^2r^2>0.$$ (11) These lead to the relations $`s_{out}=+1`$ $``$ $`s_{in}=+1`$ (12) $`s_{in}=1`$ $``$ $`s_{out}=1.`$ (13) For convenience, define the following two functions: $`P(r)`$ $``$ $`(\chi ^2+\mathrm{\Lambda }+k^2)r^{2n2}2mr^{n2}+Q^2`$ (14) $`N(r)`$ $``$ $`(\chi ^2+\mathrm{\Lambda }k^2)r^{2n2}2mr^{n2}+Q^2,`$ (15) with asymptotic behaviors $$\underset{r0}{lim}P(r)=Q^2,\underset{r\mathrm{}}{lim}P(r)=+\mathrm{},$$ (16) and $$\underset{r0}{lim}N(r)=Q^2,\underset{r\mathrm{}}{lim}N(r)=\mathrm{sign}(\chi ^2+\mathrm{\Lambda }k^2)\mathrm{}.$$ (17) First focus on $`P(r)`$. Substituting (2), (3) in (11) we obtain $$s_{out}=+1P(r)<0.$$ (18) However, $`dP/dr=0`$ at only one point, $`r=\overline{r}_{out}`$, with $$\overline{r}_{out}=\left[\frac{m(n2)}{(n1)(\chi ^2+\mathrm{\Lambda }+k^2)}\right]^{1/n},$$ (19) and hence there are two possibilities for the behavior of $`s_{out}`$: $$P(\overline{r}_{out})>0s_{out}=1,$$ (20) for all $`r`$, or $$P(\overline{r}_{out})<0\{\begin{array}{cc}s_{out}=+1\hfill & r_{out}^{(1)}<r<r_{out}^{(2)}\hfill \\ s_{out}=1\hfill & r<r_{out}^{(1)},r>r_{out}^{(2)}\hfill \end{array},$$ (21) where $`r_{out}^{(1)}`$ and $`r_{out}^{(2)}`$ are the real solutions of $`P(r)=0`$. Now we turn to $`N(r)`$. Substituting (2) and (3) in (10) we obtain $$s_{in}=+1N(r)<0.$$ (22) Because of (17), there are more possibilities for $`s_{in}`$ than for $`s_{out}`$. If $`\chi ^2+\mathrm{\Lambda }k^2<0`$, then $`dN/dr<0`$ for all $`r`$. In this case $`s_{in}=+1`$ $``$ $`r>r_{in}`$ (23) $`s_{in}=1`$ $``$ $`r<r_{in},`$ (24) where $`r_{in}`$ is the single solution of $`N(r)=0`$. However, if $`\chi ^2+\mathrm{\Lambda }k^2>0`$ then $`dN/dr=0`$ at only one point, $`r=\overline{r}_{in}`$, with $$\overline{r}_{in}=\left[\frac{m(n2)}{(n1)(\chi ^2+\mathrm{\Lambda }k^2)}\right]^{1/n},$$ (25) and the corresponding possibilities for the behavior of $`s_{in}`$ are $$N(\overline{r}_{in})>0s_{in}=1,$$ (26) for all $`r`$, or $$N(\overline{r}_{in})<0\{\begin{array}{cc}s_{in}=1\hfill & r<r_{in}^{(1)},r>r_{in}^{(2)}\hfill \\ s_{in}=+1\hfill & r_{in}^{(1)}<r<r_{in}^{(2)}\hfill \end{array}$$ (27) where $`r_{in}^{(1)}`$ and $`r_{in}^{(2)}`$ are the real solutions of $`N(r)=0`$. The possibilities enumerated above for the behavior of the sign parameters $`s_{in}`$ and $`s_{out}`$, are precisely what we need to perform an analysis of the trajectories of our charged domain wall. ## III Trajectories We will describe the trajectories of the spherically symmetric charged walls using Penrose diagrams. To keep things simple, we will present diagrams for the familiar asymptotically flat case. The features of the trajectories we consider will be essentially the same in asymptotically anti-de Sitter space. The wall traces out a set of points in a Reissner-Nordstrøm manifold from the point of view of an external observer, and a complementary set of points in a de Sitter manifold from the point of view of an internal observer. These trajectories will be represented on pairs of Penrose diagrams describing the external and internal spaces. For reference when doing this, and to orient ourselves, we provide the separate Reissner-Nordstrøm and de Sitter Penrose diagrams in Fig. (1). Note the Reissner-Nordstrøm spacetime contains timelike singularities and a Cauchy horizon, which implies the classical evolution of the shell is not predictable, without additional boundary conditions. For the moment, we will simply fix the additional boundary conditions by requiring analyticity of the metric outside the shell, and will comment further on this choice later. From the discussion of the previous section, it is evident that for a classically allowed trajectory, the zeros of $`N(r)`$, $`r_{in}`$ must satisfy $$r_{in}>r_H1/\chi .$$ (28) By considering the physically allowed behavior of the outward pointing normal vector of the shell, one finds the zeros of $`P(r)`$, $`r_{out}`$ satisfy $$r_{}<r_{out}<r_+.$$ (29) Now, the time evolution of the shell is completely determined by equation (5) and equations (10,11). The first of these equations determines the global properties of the trajectory; in particular whether there exist any inversion points. In addition, equations (10,11) for $`s_{in}`$ and $`s_{out}`$ determine in which regions of the separate Penrose diagrams Fig. (1) the trajectory develops. We will consider only the cases in which the trajectory starts from zero or infinite radius. What are the general features of the trajectories? First note that equation (16) implies that the trajectory must begin and end in the regions $`I_{}`$ or $`III_{}`$ of the R-N Penrose diagram, while eq. (17) tells us that for $`r<r_H`$ it must be in the right part of the de Sitter Penrose diagram. Since the equations that determine the turning points of the trajectories are high order polynomial equations, it is necessary to solve numerically for the turning points to determine their position relative to the other features of the trajectory, to check that the trajectory is physically consistent. In the remainder of this section we will focus on the case of four spacetime dimensions ($`n=3`$), in which the wall is a two dimensional shell and work out all the possible trajectories. We have checked that all these trajectories may be realized also when $`n=4`$, but will not attempt to check this for general $`n`$. Let us first consider the case in which $$V(r)<1,$$ (30) for all values of $`r`$, so that there are no inversion points, and in particular let us focus on the case in which the shell starts from zero radius and expands to infinity. We will call these “growing” trajectories. In this situation there are 5 possible trajectories depending on the values of $`m`$, $`Q`$, $`\chi `$, $`k`$. 1. $`\chi ^2+\mathrm{\Lambda }>k^2;N(\overline{r}_{in})>0P(\overline{r}_{out})>0`$ In this case $`s_{in}=s_{out}=1`$ during the whole evolution. In the de Sitter diagram the shell starts in the right-hand patch with an outward normal pointing towards decreasing radii, and reaches infinite radius keeping always an increasing angle. In the R-N diagram the trajectory starts in a $`III_{}`$ region ($`s_{out}=1`$ so that the outward normal points towards decreasing radii) then crosses $`r_{}`$ and evolves with a decreasing angle to end in a $`I_{}`$ region at infinity. This behavior is summarized by the Penrose diagrams in square A1 of Fig. (2). 2. $`\chi ^2+\mathrm{\Lambda }<k^2;P(\overline{r}_{out})>0`$ In this case $`s_{in}`$ changes sign once, at $`r=r_{in}`$, while $`s_{out}=1`$ for the whole trajectory. In the de Sitter diagram the shell starts on the right patch with zero radius and evolves with an increasing angle until it reaches $`r_{in}`$ turning then to the right. In the R-N diagram the shell has the same behavior as in the preceding case. This behavior is summarized by the Penrose diagrams in square A2 of Fig. (2). 3. $`\chi ^2+\mathrm{\Lambda }>k^2,N(\overline{r}_{in})<0;P(\overline{r}_{out})>0`$ In this case $`s_{in}`$ changes twice, at $`r_{in}^{(1)}`$ and at $`r_{in}^{(2)}`$, while $`s_{out}`$ never changes. In the de Sitter diagram the evolution is the same as the previous trajectory until the first turn at $`r_{in}^{(1)}`$ which is followed by a second turn at $`r_{in}^{(2)}`$ which lead the shell towards the left. The behavior in the R-N diagram is the same as in the previous case. This behavior is summarized by the Penrose diagrams in square A3 of Fig. (2). 4. $`\chi ^2+\mathrm{\Lambda }>k^2,N(\overline{r}_{in})<0;P(\overline{r}_{out})<0`$ In this case $`s_{in}`$ changes at $`r_{in}^{(1)}`$ and $`r_{in}^{(2)}`$ and $`s_{out}`$ changes at $`r_{out}^{(1)}`$ and $`r_{out}^{(2)}`$. The behavior in the de Sitter digram is the same as the preceding case. In the R-N diagram the trajectory starts in the $`III_{}`$ region, crosses $`r_{}`$, turn to the right at $`r_{out}^{(1)}`$ and to the left as it reaches $`r_{out}^{(2)}`$, to end in the $`I_{}`$ region at infinite radius. This behavior is summarized by the Penrose diagrams in square B3 of Fig. (2). 5. $`\chi ^2+\mathrm{\Lambda }<k^2;P(\overline{r}_{out})<0`$ In this case $`s_{in}`$ changes one time and $`s_{out}`$ twice. The behavior is of the same kind as case (2) in de Sitter diagram and as the previous case in the R-N spacetime. This behavior is summarized by the Penrose diagrams in square B2 of Fig. (2). Let us now consider the possible trajectories that start at an infinite radius and collapse to zero, and for which (30) still holds. Once again there are 5 different possible trajectories, one corresponding to each of those above. The common features for these trajectories are that in the de Sitter diagram the trajectories end in the right hand patch with an outward normal pointing towards decreasing radii while in the R-N diagram the trajectory starts in a $`I_{}`$ region and ends in a $`III_{}`$ region. 1. $`\chi ^2+\mathrm{\Lambda }>k^2;N(\overline{r}_{in})>0P(\overline{r}_{out})>0`$ In the de Sitter diagram the shell starts from infinite radius with an increasing angle ($`s_{in}=1`$) and ends in the right patch with an outward normal pointing towards decreasing radius. In the R-N diagram the trajectory starts from a $`I_{}`$ region and ends in the $`III_{}`$ region without any turning point, since $`s_{out}=1`$ throughout the evolution. This is shown by the Penrose diagrams in square A1 of Fig. (3). 2. $`\chi ^2+\mathrm{\Lambda }<k^2;P(\overline{r}_{out})>0`$ In the de Sitter diagram the shell starts at infinity with decreasing angle, turns at $`r_{in}`$ and ends in the right patch. In the R-N diagram the shell behaves as in the previous case having $`s_{out}=1`$ for the whole evolution. This is shown by the Penrose diagrams in square A2 of Fig. (3). 3. $`\chi ^2+\mathrm{\Lambda }>k^2,N(\overline{r}_{in})<0;P(\overline{r}_{out})>0`$ In the de Sitter diagram the shell starts at infinity with decreasing angle, turns at $`r_{in}^{(1)}`$ and $`r_{in}^{(2)}`$ to end in the right patch. In the R-N diagram the behavior is the same as in the previous case. This is shown by the Penrose diagrams in square A3 of Fig. (3). 4. $`\chi ^2+\mathrm{\Lambda }>k^2,N(\overline{r}_{in})<0;P(\overline{r}_{out})<0`$ In the de Sitter diagram the trajectory has the same features as in the previous case. In the R-N diagram the shell starts at infinity in a $`I_{}`$ region, crosses $`r_+`$ and has two turning points at $`r_{out}^{(1)}`$ and $`r_{out}^{(2)}`$ to end in a $`III_{}`$ region. This is shown by the Penrose diagrams in square B3 of Fig. (3). 5. $`\chi ^2+\mathrm{\Lambda }<k^2;P(\overline{r}_{out})<0`$ In the de Sitter diagram the shell starts at infinity with decreasing radius, turns at $`r_{in}`$ and reaches the origin in the right patch. In the R-N diagram the shell behaves as in the previous case. This is shown by the Penrose diagrams in square B2 of Fig. (3). We now turn to cases in which there exists some value of $`r`$ at which $`V(r)>1`$, so that there are inversion points in the trajectory. For a growing trajectory (starting from zero radius), the shell cannot cross either the de Sitter horizon $`r_H`$ or the inner R-N horizon $`r_{}`$. Thus, $`r_{max}`$ $`<`$ $`r_{}`$ (31) $`r_{max}`$ $`<`$ $`r_H,`$ (32) and the behavior is represented in Fig. (4). Things are more complicated for collapsing trajectories (starting from infinite radius). We will distinguish different behaviors by the position of the minimum radius with respect to the horizons and by the changes in $`s_{in}`$ and $`s_{out}`$, so we will describe separately the behavior in the de Sitter and in the R-N spacetimes, checking then which of these trajectories can be consistently glued together. In the de Sitter space there are five possible different behaviors 1. $$s_{in}=1$$ (33) for the whole trajectory. The shell starts from infinite radius, reaches $`r_{min}`$ and bounces back at infinite radius always with an increasing angle. This is shown in square A in Fig. (5). 2. $$s_{in}=+1$$ (34) for the whole trajectory. The shell starts from infinite radius, reaches $`r_{min}`$ and returns to infinite radius always with decreasing angle. This is shown in square B in Fig. (5). 3. $$s_{in}=\{\begin{array}{cc}+1\hfill & r_{in}<r\hfill \\ 1\hfill & r_{min}<r<r_{in}.\hfill \end{array}$$ (35) The shell starts from infinity with decreasing angle, turns to the right at $`r_{in}`$, reaches $`r_{min}`$ with increasing angle and return to infinity with time reversal behavior. This is shown in square C in Fig. (5). 4. $$s_{in}=\{\begin{array}{cc}+1\hfill & r_{min}<r<r_{in}\hfill \\ 1\hfill & r_{in}<r.\hfill \end{array}$$ (36) The shell starts from infinity with increasing angle, reaches $`r_{in}`$ turning to the left to go to $`r_{min}`$ with decreasing angle and bounces back at infinity. This is shown in square D in Fig. (5). 5. $$s_{in}=\{\begin{array}{cc}+1\hfill & r_{in}^{(1)}<r<r_{in}^{(2)}\hfill \\ 1\hfill & r_{min}<r<r_{in}^{(1)},r_{in}^{(2)}<r.\hfill \end{array}$$ (37) The shell starts with increasing angle at infinite radius, turn to the left to at $`r_{in}^{(2)}`$ and then to the right at $`r_{in}^{(1)}`$ to reach $`r_{min}`$ and go back to infinity. This is shown in square E in Fig. (5). In the R-N space we have three possible behaviors. 1. $$s_{out}=1$$ (38) for the whole evolution. The trajectory starts from infinite radius in a $`I_{}`$ region and reaches $`r_{min}`$ with decreasing angle in a $`III_{}`$ region to bounce back at infinite radius in a $`I_{}`$ region. This is shown in square 1 in Fig. (5). 2. $$s_{out}=\{\begin{array}{cc}+1\hfill & r_{min}<r<r_{out}\hfill \\ 1\hfill & r_{out}<r.\hfill \end{array}$$ (39) The shell starts in a $`I_{}`$ region with decreasing angle, turn at $`r_{out}`$ to reach $`r_{min}`$ in a $`III_+`$ region and bounces back at infinite radius. This is shown in square 2 in Fig. (5). 3. $$s_{out}=\{\begin{array}{cc}+1\hfill & r_{out}^{(1)}<r<r_{out}^{(2)}\hfill \\ 1\hfill & r_{min}<r<r_{out}^{(1)},r_{out}^{(2)}<r.\hfill \end{array}$$ (40) The shell starts with infinite radius in a $`I_{}`$ region with decreasing angle, then turns to the left at $`r_{out}^{(1)}`$ and to the right at $`r_{out}^{(2)}`$ to reach $`r_{min}`$ in a $`III_{}`$ region. This is shown in square 3 in Fig. (5). For trajectories with $`r_{min}<r_{}`$ it is impossible to glue consistent with eq. (12) the trajectories A, C and E in the de Sitter diagram with the 2 in the R-N. Nor can the A trajectory in the de Sitter diagram be glued with with the 3 trajectory in R-N. All the other configurations are possible and have been realized numerically. The only trajectories remaining are those with $`r_{min}>r_+`$. In the R-N conformal diagram they will always stay in the $`I_{}`$ region as we show in Fig. (6). This can be glued consistently with the A, B, and D trajectories in the de Sitter spacetime. The bubble solutions we have described in this section contain timelike singularities. It is interesting to consider whether these solutions could be patched onto solutions with smooth initial data, removing the past timelike singularities. Physically, this would correspond to the ability to build a machine that would be capable of producing a black hole with an inflating interior. Uncharged collapsing shells have previously been considered in . The problem encountered there was that whenever the bubble grew sufficiently large, anti-trapped regions would necessarily appear . By the Penrose singularity theorem this implies a past directed null geodesic must be incomplete . This amounts to the condition that initial conditions must be singular. Similar solutions have been considered in the context of colliding monopoles in . There a way of circumventing the problem of singular initial data was suggested. A timelike singularity would form behind an event horizon, and the anti-trapped regions that must necessarily form, would be to the future of this timelike singularity. In this way, the Penrose theorem is satisfied, and it can be possible to realize this solution with smooth initial data. This is exactly the state of affairs in the collapse and bounce bubble solutions we have constructed. In principle, therefore, it should be possible to patch these bubble solutions onto smooth initial data. However, another obstacle encountered in was that classically it is not possible to evolve the solutions past the timelike singularity without specifying additional boundary conditions on the singularity. In the next section, we describe how this problem is solved for analogous asymptotically anti-de Sitter bubble solutions by appealing to the duality between gravity in AdS backgrounds and large $`N`$ conformal field theory. ## IV AdS/CFT Interpretation In principle, the correspondence between gravity in an anti-de Sitter background and conformal field theory on the boundary , allows us to map any process in the gravity theory to a statement in the field theory. The duality is conjectured to hold for any spacetime dimension bigger than 2 (on the gravity side). The simplest case to consider is five-dimensional supergravity in an AdS background which is dual to four-dimensional $`SU(N)`$ Yang-Mills theory with 4 supersymmetries. The radius of curvature $`R`$ of AdS and the string coupling $`g_s`$ are related to parameters in the Yang-Mills theory by $$R=(g_{YM}^2N)^{1/4},g_s=g_{YM}^2.$$ (41) The large $`N`$ limit, with $`R`$ fixed, will correspond to the classical limit of the supergravity theory. For other dimensions, the CFT is an exotic superconformal field theory which is difficult to describe explicitly. The important point for us is that evolution in the CFT is unitary, which allows us to describe, in a completely well-defined way, processes in gravity which lead to classical singularities. The bubble solutions we have discussed, and ones previously found , carry over to backgrounds where the asymptotically flat backgrounds are replaced by asymptotically anti-de Sitter backgrounds. Furthermore they can be generalized to general spacetime dimensions (the relevant black hole solutions may be found in .) The qualitative properties of the trajectories will not change provided $`R`$ is not taken to be too small relative to the other parameters. We will take the point of view that the correspondence between gravity and conformal field theory is valid, and discuss what these solutions must correspond to from the CFT point of view. The AdS/CFT conjecture is most precisely formulated in Euclidean space . We will follow the procedure discussed in , to obtain a definition of the Lorentzian gravity theory. The logic is as follows: the conjecture asserts that all observables in the Euclidean gravity theory can be computed in terms of gauge invariant observables in the Euclidean gauge theory. In particular, we assume this gives us enough information to construct the metric in the gravity theory (modulo coordinate transformations), for any given process in the gauge theory. The Lorentzian gauge theory is defined by the usual Wick rotation from imaginary time in the correlation functions, which leads to a definition for the observables in the Lorentzian gravity theory in terms of a Wick rotation of the Euclidean observables. Reissner-Nordstrom black holes have been previously considered in the AdS/CFT context by Chamblin, Emparan, Johnson and Myers . In terms of the CFT, the black hole microstates correspond to a finite temperature plasma of charge associated with a $`U(1)`$ subgroup of the global R-symmetry. Let us consider the interpretation of the bubble trajectories in terms of the large $`N`$ CFT. A bubble in region I far from the event horizon will correspond to a charged shell in CFT. The description of analogous neutral shells in the AdS/CFT correspondence has recently been studied by . For spherically symmetric shells, the radial position is encoded in non-local correlations in the CFT state. For the cases of most interest for us, the allowed collapse and bounce trajectories of the charged bubbles are always behind the event horizon of a black hole from the point of view of an outside observer. These solutions therefore correspond to particular black hole microstates from the point of view of an outside observer. Thus in the CFT description, the bubbles correspond to particular microstates in the finite temperature plasma associated with AdS Reissner-Nordstrom black holes. It is difficult to describe these states explicitly without more direct control over calculations in the strongly coupled gauge theory. If it is possible to indeed patch the bubble solutions onto configurations with smooth initial data, it would be much more straightforward to map this into initial data in the CFT, using the current understanding of the mapping between states in the CFT to states in the gravity theory . The interior of the domain wall is de Sitter space. In order for this to be a solution of the gravity theory, we suppose that appropriate perturbations are turned on in the CFT to lead to a phase with broken supersymmetry inside the domain wall. A non-trivial potential may then be generated for one of the scalars in the gravity theory, allowing it to act as an inflaton field, generating a non-trivial vacuum energy in this region. It is possible to construct a unitary time evolution operator which evolves states across the event horizon and through the singularity by combining the usual field theory time translation operator with the conformal generators. In the case of a neutral black hole in $`AdS_3`$ this construction has been made explicit . Thus the CFT description resolves the classical singularity of the black hole geometry and leads to a prescription for evolving gravity solutions to the future of timelike singularities. The CFT implies definite boundary conditions on the timelike singularities which in general will depend on the initial state. In the large $`N`$ limit, and for sufficiently large bubbles, it should be possible to make these conditions arbitrary by an appropriate choice of the CFT initial state. This follows from the fact that an arbitrary configuration in the gravity theory can be mapped to a configuration in the CFT. The bubble solutions described here correspond to boundary conditions fixed by analyticity of the metric. Finite $`N`$ corrections in the CFT correspond to quantum corrections from the gravity point of view. Black holes will be unstable at the quantum level due to Hawking radiation. The supergravity theory contains charged, light fields, hence the Hawking radiation will tend to discharge the black hole . In line with the discussion of black hole complementarity in AdS/CFT of , the CFT implies the following picture for the quantum evolution of the bubble in the gravity theory. The gravity variables inside and outside the event horizon are redundant. Outside the horizon, the final state will be a set of light outgoing quanta which carry all the information about processes inside the black hole event horizon. This will now be a complicated time dependent state from the CFT viewpoint. The degrees of freedom inside the horizon, including those describing the eternally inflating region, are encoded in the same field theory degrees of freedom that describe physics outside. Note the most interesting bubble solutions contain eternally inflating regions. The description of these solutions in terms of the CFT implies that even though we have a large (and increasing) proper volume inside the bubble, only a finite number of states can exist in this region. The Bekenstein-Hawking entropy of the black hole is finite, and will bound the number of field theory microstates corresponding to such configurations. The picture of the quantum evolution of the bubble described above alleviates another problem with the classical solution we have thus far not addressed. A Cauchy horizon is classically unstable to small perturbations, which grow producing weak curvature singularities in the vicinity of the Cauchy horizon . A free-falling observer crossing the Cauchy horizon sees the entire history of the external universe in the moments prior to crossing. The picture of the quantum evolution of our solutions we have discussed above solves this problem. The black hole evaporates in a finite time, which avoids the problem of an observer inside seeing the entire history of the external universe. Furthermore, because the CFT description gives a resolution of the classical singularity of the black hole, the Cauchy horizon is no longer present quantum mechanically. What is less clear is precisely how this singularity is resolved. One possibility is that a spacelike region of strong curvature forms outside the would-be Cauchy horizon, the system behaves in the same way as a Schwarzschild black hole, and the “tunnel” to the inflating region collapses. Another possibility is that the curvature remains finite in the vicinity of the inner horizon, and we do indeed have an eternally inflating region inside the bubble. We propose, at least for special choices of initial conditions, that the second option is viable. It seems likely that the constraints on these initial conditions are very strong, and it would be interesting to study further the experimental feasibility of creating an inflating universe in the laboratory. ## Acknowledgments We would like to thank Stephon Alexander, Roberto Balbinot, Cyrus Taylor and Tanmay Vachaspati for helpful conversations. The work of M.T. was supported by the Department of Energy (D.O.E.). The research of D.L. is supported in part by DOE grant DE-FE0291ER40688-Task A.
no-problem/9906/cond-mat9906221.html
ar5iv
text
# Planar cyclotron motion in unidirectional superlattices defined by strong magnetic and electric fields: Traces of classical orbits in the energy spectrum ## I Introduction In the last decade there has been a constant interest in the transport properties of the periodically modulated two dimensional electron gas (2DEG). In particular, in the presence of a lateral modulation of a one-dimensional character the resistivity may be strongly anisotropic, which essentially reflects the anisotropy of the electronic states. Two types of modulations can be achieved in the experimental devices: electrostatic potential modulations and, more recently, magnetic field modulations. Weak modulations of both types lead already to pronounced magnetoresistance effects in the presence of an average magnetic field $`B_0`$ applied perpendicular to the 2DEG. These effects occur at low and intermediate $`B_0`$ values, well below the magnetic quantum regime where Shubnikov-de Haas oscillations appear. At very small values of $`B_0`$ a pronounced positive magnetoresistance is observed, followed at intermediate $`B_0`$ values by the “Weiss oscillations” due to commensurability effects. Both effects are adequately understood within a classical transport calculation based on Boltzmann’s equation, and can be traced back to the predominance of different types of classical trajectories. The positive magnetoresistance is understood as caused by “channeled orbits” which exist if the modulation is sufficiently strong or, equivalently, the average magnetic field is sufficiently small. For electric modulation they occur near the minima of the modulation potential (“open” orbits ), and for magnetic modulation near the lines of vanishing total magnetic field (“snake” orbits ). They are always confined within a single period of the modulation, which we choose in $`x`$ direction. They are wavy trajectories allowing for fast motion of electrons with velocities within small angles around the direction of translational invariance ($`y`$ direction). These channeled orbits occur in addition to the “drifting orbits”, which are self-intersecting trajectories with loops (along each of which the direction of the velocity changes by $`2\pi `$), so that usually a low drift velocity in $`y`$ direction results. For sufficiently small $`B_0`$, drifting orbits may extend over many periods of the modulation. At sufficiently large $`B_0`$ (sufficiently small modulation amplitudes) only the drifting orbits survive. The “Weiss oscillations” manifest a commensurability effect depending on the ratio of the extent of drifting orbits (at the Fermi energy) and the modulation period. With increasing modulation strength, the positive magnetoresistance becomes more pronounced and extends to larger $`B_0`$ values, suppressing progressively the low-$`B_0`$ Weiss oscillations. This effect is well understood within the classical calculation , if both types of trajectories are adequately included, and it has recently also been obtained by a quantum calculation for a strong modulation . A qualitatively new type of magnetoresistance effect has recently been observed by Ye et al. on samples with an extremely strong magnetic modulation. Samples with a surface array of ferromagnetic micro-strips were measured in tilted magnetic fields, so that the applied magnetic field had a large component parallel to the surface, producing a large magnetization of the ferromagnetic strips, while only the small perpendicular component determined the average magnetic field $`B_0`$ in the 2DEG. In this way a huge positive magnetoresistance with superimposed Shubnikov-de Haas like oscillations was obtained at low values of the average magnetic field, at which no magnetic quantum effects should be expected for weak modulation. It rather seems that the quantum oscillations are induced by the large-amplitude periodic magnetic modulation field. Such conditions require a quantum transport theory and, as a first step, the understanding of the quantum electronic states of a 2DEG with a strong magnetic modulation. This is the motivation of the present work. Channeled and drifting quantum states in linearly varying magnetic field are already discussed by other authors. The Schrödinger equation for periodic magnetic fields alternating in sign, has been solved previously, but only for the case when the average field is zero . In the present paper we shall study the quantum electronic states in strong periodic magnetic fields with a non-vanishing average, and compare it with the case of a strong electric modulation. In both situations, rather complicated energy spectra are obtained, with striking qualitative similarities and quantitative differences. For the case of strong electric modulation, such a complicated energy spectrum has recently been published , but without an attempt of an explanation. We will demonstrate in this paper that a close comparison with classical motion leads to a detailed and intuitive understanding of these spectra and the corresponding eigenstates. In Sec. II we start with some general remarks on the relation between quantum and classical description of the 2D electron motion in 1D lateral superlattices, and we introduce suitable reduced units. In Sec. III we focus on the effect of a simple harmonic magnetic modulation of arbitrary strength. In Sec. IV we include an electric modulation, which requires a somewhat different analytical procedure. The inclusion of electric modulation seems also necessary from the experimental point of view, since the ferromagnetic strips on the sample surface introduce a periodic stress field in the sample, which acts as an electric modulation on the 2DEG. Finally, in Sec. V we summarize the essential features derived in the paper and extend the discussion beyond the model of simple harmonic modulations. Some of the present results have been recently published in a preliminary form. ## II General remarks We consider a (non-interacting) 2DES in the $`x`$-$`y`$ plane subjected to a magnetic field with $`z`$ component $`B_z(x)=B_0+B_m(x)`$ and an electrostatic field in $`x`$ direction leading to a potential energy $`U(x)`$. Our aim is a close comparison of the classical and the quantum description of the electron motion (in terms of orbits and wavefunctions, respectively) in such fields, especially in the case that $`U(x)`$ and $`B_m(x)`$ are periodic in $`x`$ with the same period $`a`$ and vanishing average values. To evidence the translation invariance in $`y`$ direction in the (either classical or quantum) Hamiltonian $$H=\frac{1}{2m}\left(𝐩+e𝐀\right)^2+U,$$ (1) we describe $`B_z(x)`$ by an $`x`$-dependent vector potential $`𝐀(x)=A(x)𝐞_𝐲`$ with $`A(x)=xB_0+A_m(x)`$ and $`A_m(x)=_0^x𝑑x^{}B_m(x^{})`$. Then $`y`$ is a cyclic variable and the canonical momentum $`p_y`$ is conserved, and one obtains an (one-dimensional) effective Hamiltonian $`H(X_0)=p_x^2/2m+V(x;X_0)`$. For $`B_00`$, the effective potential can be written $$V(x;X_0)=\frac{m}{2}\omega _0^2\left(xX_0+\frac{A_m(x)}{B_0}\right)^2+U(x),$$ (2) where $`X_0=p_y/eB_0`$ is the center coordinate of the effective potential and $`\omega _0=eB_0/m`$ is the cyclotron frequency, both in the absence of modulation. In the quantum description, the reduction to a one-dimensional problem is achieved by the product ansatz $`\mathrm{\Psi }_{n,X_0}(x,y)=L_y^{1/2}\mathrm{exp}(ip_yy/\mathrm{})\psi _{n,X_0}(x)`$ for the energy eigenfunctions, where $`L_y`$ is a normalization length, and the discrete quantum number $`n=0,1,2,\mathrm{}`$ counts the nodes of the reduced wavefunction $`\psi _{n,X_0}(x)`$. If $`U(x)`$ and $`A_m(x)`$ are bounded, the $`\psi _{n,X_0}(x)`$ drop Gaussian-like for $`|x|\mathrm{}`$, and for a fixed value of the quasi-continuous quantum number $`X_0`$ the energy spectrum $`E_n(X_0)`$ is discrete. In the classical description, we use the equation $`mv_y=p_y+eA(x)`$, which may also be derived directly from Newton’s equation, to eliminate the velocity $`v_y`$. The effective motion in $`x`$ direction is determined by $`H(X_0)=E`$. Similar to the wavefunctions, the orbits for given constants of motion, $`X_0`$ and $`E`$, are bounded in $`x`$ direction, however the energy $`E`$ is a continuous variable. For a given $`E=E_\mathrm{F}`$, each position $`x`$ (with $`U(x)<E_\mathrm{F}`$) is the turning point of two orbits which are characterized by the center coordinates $$X_0^\pm (x)=x+\frac{A_m(x)}{B_0}\pm R_0\sqrt{1\frac{U(x)}{E_\mathrm{F}}},$$ (3) obtained from $`H(X_0)=E_\mathrm{F}`$ for $`v_x=p_x/m=0`$. Here $`R_0=v_\mathrm{F}/\omega _0`$ is the cyclotron radius of electrons moving with energy $`E_\mathrm{F}=mv_\mathrm{F}^2/2`$ in the magnetic field $`B_0`$. For given $`E_\mathrm{F}`$ and $`X_0`$, orbits exist in intervals in which $`X_0^{}(x)X_0X_0^+(x)`$ holds. This allows a convenient classification of the possible orbits at fixed energy $`E_\mathrm{F}`$ and for varying $`X_0`$. Of course, the same classification can also be done by directly investigating the effective potential. This may be preferred if one is interested in orbits at different energies but the same $`X_0`$. The calculation of the orbits is a simple textbook problem, but must in general be done numerically. In accordance with the translational symmetry of the problem, we will in the following not distinguish orbits which differ only by a rigid shift in $`y`$ direction. If an electron is at time $`t_i`$ at position $`(x_i,y_i)`$ on an orbit characterized by the constants of motion $`E_\mathrm{F}`$ and $`X_0`$, with turning points $`x_l`$ and $`x_r`$ ($`x_l<x_i<x_r`$), it moves towards one of the turning points so that at time $$t(x;X_0,E_\mathrm{F})=t_i+_{x_i}^x\frac{dx^{}}{|v_x(x^{};X_0,E_\mathrm{F})|}$$ (4) it is at position $`(x,y(x;X_0,E_\mathrm{F}))`$, with $$y(x;X_0,E_\mathrm{F})=y_i+_{x_i}^x\frac{v_y(x^{};X_0)}{v_x(x^{};X_0,E_\mathrm{F})}𝑑x^{},$$ (5) where $`|v_x(x;X_0,E_\mathrm{F})|=v_\mathrm{F}\sqrt{1V(x;X_0)/E_\mathrm{F}}=\omega _0\sqrt{[X_0^+(x)X_0][X_0X_0^{}(x)]}`$ and $`v_y(x;X_0)=(\omega _0/2)[X_0^+(x)+X_0^{}(x)2X_0]`$. If at one of the turning points $`x_l`$ or $`x_r`$, where $`v_x(x;X_0,E_\mathrm{F})=0`$, the derivative $`V(x;X_0)/x`$ vanishes, we call this turning point and this orbit “critical”. At critical turning points the integrals (4) and (5) diverge, so that the critical orbits there asymptotically approach straight lines parallel to the $`y`$ axis. For non-critical orbits the integrals (4) and (5) converge as $`x`$ approaches the turning points, and the total orbit can be composed out of right-running ($`v_x>0`$) and left-running ($`v_x<0`$) pieces with finite traverse time $`T(X_0,E_\mathrm{F})=_{x_l}^{x_r}𝑑x/|v_x(x;X_0,E_\mathrm{F})|`$. The probability density of finding the electron at position $`x`$ is $`W(x;X_0,E_\mathrm{F})=1/[T(X_0,E_\mathrm{F})|v_x(x;X_0,E_\mathrm{F})|]`$. This is the classical analog to $`|\psi _{n,X_0}(x)|^2`$. If $`U(x)=U(x+a)`$ and $`B_m(x)=B_m(x+a)`$ are periodic with period $`a`$, as we will assume in the following, the effective potential, Eq. (2), has the symmetry $`V(x+a;X_0+a)=V(x;X_0)`$. As a consequence, the energy spectrum is also periodic, $`E_n(X_0+a)=E_n(X_0)`$, and can be restricted to the “first Brillouin zone” $`0X_0a`$. The eigenfunctions can be taken to satisfy $`\psi _{n,X_0+a}(x)=\psi _{n,X_0}(xa)`$. The corresponding classical symmetry is that an orbit characterized by $`E_\mathrm{F}`$ and $`X_0+a`$ differs from that characterized by $`E_\mathrm{F}`$ and $`X_0`$ only by a rigid shift of amount $`a`$ in $`x`$ direction. The dispersion of the energy bands $`E_n(X_0)`$ implies a group velocity in $`y`$ direction, $$n,X_0|v_y|n,X_0=\frac{1}{m\omega _0}\frac{dE_n(X_0)}{dX_0},$$ (6) which is the expectation value of the velocity operator in the energy eigenstate $`\psi _{n,X_0}`$. It is the quantum equivalent to the classical drift velocity, i.e. the average velocity (in $`y`$ direction) along the corresponding classical orbit. The drift velocity in $`x`$ direction vanishes, since the orbits are bounded in $`x`$ direction. ### A Suitable units For an economic comparison of classical and quantum aspects it is important to use suitable length and energy units, which are meaningful for both the quantum description and the classical limit. Doing so, we will see that the classical features depend on fewer scaled parameters than the quantum ones. To be specific but still rather general, we assume in the following periodic modulations of the form $`B_m(x)=B_m^0b(Kx)`$ and $`U(x)=V_0u(Kx)`$ for the magnetic and the electric modulation, respectively, where $`b(\xi )`$ and $`u(\xi )`$ are dimensionless periodic functions with period $`2\pi `$ and vanishing average values. Thus $`B_m(x)`$ and $`U(x)`$ have the same period $`a=2\pi /K`$, but may have different shapes and phases. In the numerical examples we will use for both $`b(\xi )`$ and $`u(\xi )`$ simple cosines, eventually with a phase shift. The average magnetic field $`B_0`$ sets, with the magnetic length $`l_0=\sqrt{\mathrm{}/(m\omega _0)}`$ and the cyclotron energy $`\mathrm{}\omega _0`$, both a length and an energy scale, which are useful for quantum calculations, but have no meaning for the classical motion. For the discussion of commensurability effects, such as the Weiss oscillations, the cyclotron orbits must be compared with the period $`a`$ of the modulation. Therefore $`a`$ is a natural choice for the lengths unit. The choice of a suitable energy unit is motivated as follows. Classically, $`B_0`$ determines only the cyclotron frequency $`\omega _0`$, and one needs an independent length $`l`$ to define an energy scale $`V_{\mathrm{mag}}=m\omega _0^2l^2/2`$. Using $`l`$ as length unit, we may define dimensionless variables $`\xi =x/l`$ and $`\xi _0=X_0/l`$. The effective potential, Eq. (2), then can be written as $`V(x;X_0)=V_{\mathrm{mag}}\stackrel{~}{v}(\xi ;\xi _0)`$ with $$\stackrel{~}{v}(\xi ;\xi _0)=[\xi \xi _0+sa(Kl\xi )/Kl]^2+wu(Kl\xi ),$$ (7) where $`s=B_m^0/B_0`$, $`a(\zeta )=_0^\zeta 𝑑\zeta ^{}b(\zeta ^{})`$, and $`w=V_0/V_{\mathrm{mag}}`$. In the quantum description, the kinetic energy operator $`(\mathrm{}^2/2m)d^2/dx^2=E_ld^2/d\xi ^2`$, introduces a new energy scale $`E_l=\mathrm{}^2/(2ml^2)`$, which has no classical analog. Introducing the energy ratio $`\alpha =E_l/V_{\mathrm{mag}}`$, we write the effective Schrödinger equation as $$\left[\alpha \frac{d^2}{d\xi ^2}+\stackrel{~}{v}(\xi ;\xi _0)\stackrel{~}{\epsilon }_n(\xi _0)\right]\stackrel{~}{\psi }_{n,\xi _0}(\xi )=0,$$ (8) with $`\stackrel{~}{\epsilon }_n(\xi _0)=E_n(X_0)/V_{\mathrm{mag}}`$ and $`\stackrel{~}{\psi }_{n,\xi _0}(\xi )=\sqrt{l}\psi _{n,X_0}(x)`$. If we would take $`l=l_0`$, we had $`V_{\mathrm{mag}}=E_l=\mathrm{}\omega _0/2`$ and thus simply $`\alpha =1`$. The effective potential Eq. (7) would then depend on the constant of motion $`\xi _0`$ and, in addition, on three dimensionless model parameters, $`s`$, $`w`$, and $`Kl_0`$. To specify an eigenstate or, in the classical description, a trajectory, one further needs an energy value $`\stackrel{~}{\epsilon }`$ as a second constant of motion. A description that, for fixed constants of motion, needs three parameters to specify the effective potential and, furthermore, relies on $`l_0`$ and $`\mathrm{}\omega _0`$, which have no meaning in classical mechanics, is rather clumsy and not acceptable. Instead we take $`l=1/K`$ and, therefore, $`V_{\mathrm{mag}}=V_{\mathrm{cyc}}`$, where $`V_{\mathrm{cyc}}=m\omega _0^2/(2K^2)`$ is the energy of a classical cyclotron orbit of radius $`1/K`$ in the homogeneous magnetic field $`B_0`$. Now the effective potential Eq. (7) depends only on the two dimensionless modulation strengths $`s`$ and $`w=V_0/V_{\mathrm{cyc}}`$, which both are well defined within the classical approach. Also the constants of motion, $`\xi _0=KX_0`$ and $`\stackrel{~}{\epsilon }=E/V_{\mathrm{cyc}}=(KR_0)^2`$, including the dimensionless version of Eq. (3), $$\xi _0^\pm (\xi )=\xi +sa(\xi )\pm \sqrt{\stackrel{~}{\epsilon }wu(\xi )},$$ (9) remain meaningful in the classical limit. This choice of units will also be very useful for a systematic discussion of the quantum mechanical energy spectra. Quantum mechanics enters the effective Schrödinger equation (8) only via the parameter $`\alpha =(l_0K)^4`$, which scales the kinetic energy. It determines the only true quantum aspect of the spectrum, namely the spacing of the energy levels $`\stackrel{~}{\epsilon }_n(\xi _0)`$. We will see in Sec. III B that all the essential structural features of the energy spectrum, e.g. the complicated back-folded structure due to the coexistence of “channeled” and “drifting” states, are determined solely by the “classical” parameters $`s`$ and $`w`$. The density of the quantized levels $`\stackrel{~}{\epsilon }_n(\xi _0)`$, on the other hand, increases with increasing ratio $`a/l_0`$. As a simple example one may consider the well known case of a weak electric or magnetic cosine modulation, which leads to modified Landau bands of oscillatory width. The band width assumes minima near the “flat band” energies $`E_\lambda ^\pm =m(\omega _0a)^2(\lambda \pm 1/4)/8`$, with “$`+`$” (“$``$”) for magnetic (electric) modulation and $`\lambda =1,2,\mathrm{}`$. These flat band energies are distinct multiples of our energy unit $`V_{\mathrm{cyc}}`$, and occur at $`\stackrel{~}{\epsilon }_\lambda ^\pm =\pi ^2(\lambda \pm 1/4)`$, independent of the special values of the model parameters $`B_0`$ and $`a`$. The level spacing, on the other hand, is of the order $`\mathrm{}\omega _0`$ and depends in our units on $`\mathrm{}\omega _0/V_{\mathrm{cyc}}=2l_0^2K^2=2\sqrt{\alpha }`$. ## III Magnetic cosine modulation We first consider a pure magnetic modulation, $`U(x)0`$, $`B_m(x)=B_m^0b(Kx)`$, so that the effective potential Eq. (7) becomes $$V(\xi ;\xi _0)=V_{\mathrm{cyc}}[\xi \xi _0+sa(\xi )]^2.$$ (10) For $`s=0`$ one obtains the well known Landau levels and the Landau oscillator wave functions, $`f_{nX_0}(x)`$. We use the set $`f_{nX_0}`$ as the basis of our Hilbert space in order to obtain numerical solutions for $`s0`$, by numerical diagonalization of $`H(X_0)`$. The electron effective mass is that of GaAs, $`m=0.067m_0`$. We further assume spin degeneracy. For the numerical parameters chosen here the size of the basis will vary between 150-300 Landau levels. Before discussing the numerical results we summarize some properties of the effective potential and of Eq. (9), which now reduces to $$\xi _0^\pm (\xi )=\xi +sa(\xi )\pm KR_0.$$ (11) For a fixed $`\xi _0`$ the local extrema of the effective potential, given by $`V(\xi ,\xi _0)/\xi =0`$, are the points where the total magnetic field is zero, i. e. the roots of $$1+sb(\xi )=0,$$ (12) and the points where the effective potential is zero, i. e. the roots of $$\xi \xi _0+sa(\xi )=0.$$ (13) An important aspect for the following discussion is that the roots of the first kind, Eq.(12), if existent, are independent of $`\xi _0`$, while those of the second kind, Eq.(13), do depend on $`\xi _0`$. We will see that orbits with $`\xi `$ values near roots of the first kind are channeled, while those with $`\xi `$ values near roots of the second kind are drifting orbits. The analytic dependence of the effective potential on the relevant position coordinate $`\xi `$ is determined by the modulation strength $`s`$. Therefore the number of its possible zeroes, the classification of orbits and the energy spectrum depend critically on the parameter $`s`$. To demonstrate this, we choose in the following examples $`b(\xi )=\mathrm{cos}\xi `$, and consequently $`a(\xi )=\mathrm{sin}\xi `$. ### A Weak modulation, $`s1`$ For $`s<1`$, the effective potential has exactly one minimum of the second kind for each value $`\xi _0`$, which is due to the confinement by the average magnetic field. The functions $`\xi _0^\pm (\xi )`$ in equation (11) have no extrema. For each value $`\xi _0`$ they determine exactly one orbit, which is a drifting cyclotron orbit. By this we mean a self-intersecting orbit consisting of loops along each of which the azimuth angle in velocity space, $`\phi =\mathrm{arctan}(v_y/v_x)`$ increases by $`2\pi `$. A typical example is illustrated in Fig. 1 for $`s=0.5`$, $`\xi _0=\pi /2`$ (i.e. $`X_0=a/4`$), and two energy values $`E_\mathrm{F}`$. Figure 1(a) shows the effective potential. For a given energy $`E=E_\mathrm{F}`$ (horizontal line) a classical orbit exists where $`V(\xi ;\xi _0)E_\mathrm{F}`$. Figure 1(b) shows the location of the turning points as the crossing points of the horizontal line $`\xi _0=\pi /2`$ with the functions $`\xi _0^\pm (\xi )`$. The corresponding drifting orbit exists in the interval with $`\xi _0^{}(\xi )\xi _0\xi _0^+(\xi )`$. The orbits in real space are illustrated in Fig. 1(c). In Fig. 2 we plot the corresponding quantities for $`s=0.5`$ and $`\xi _0=\pi `$ (i.e. $`X_0=a/2`$). In this case the effective potential is symmetric with respect to the center coordinate $`X_0`$. As a consequence, the orbits are closed and their drift velocity in $`y`$ direction is zero. For small energies, $`E_\mathrm{F}/V_{\mathrm{c}yc}=(KR_0)^2<\pi ^2`$ (i.e. $`2R_0<a`$), the extents of the orbits in $`x`$ direction are smaller than a modulation period and essentially determined by the local values of the total magnetic field. At high energies, $`E_\mathrm{F}/V_{\mathrm{c}yc}1`$, the orbits extend over several periods of the modulation and the extent of an orbit, i.e., the width of the effective potential valley at the corresponding energy, is determined by the cyclotron radius in the average magnetic field ($`x_rx_l2R_0`$). In Fig. 3(a) we display the first 50 energy bands $`E_{n\xi _0}`$ calculated from the (first 150) original, degenerated Landau levels, for s=0.5. The level spacing of the lowest energy bands is seen to follow the local value of the total magnetic field, Fig. 3(b). This is expected from the local approximation $`E_{n\xi _0}(n+1/2)\mathrm{}eB(\xi _0)/m`$, which is valid if the extent of the wavefunctions $`\psi _{n,X_0}(x)`$ is smaller than the modulation period. With our energy unit $`V_{\mathrm{c}yc}`$ the apparent level spacing of energies which are independent of the period $`a`$ becomes proportional to $`\sqrt{\alpha }`$. For example, if the local approximation $`E_{n\xi _0}(n+1/2)\mathrm{}\omega (\xi _0)`$ holds for $`E_{n\xi _0}<4V_{\mathrm{c}yc}`$, as in Fig. 3(a), this implies that it holds for $`n+1/2<4V_{\mathrm{c}yc}/[\mathrm{}\omega (\xi _0)]=2[\omega _0/\omega (\xi _0)]/\sqrt{\alpha }`$. Thus, the number of bands which are well described by the local approximation increases quadratically with increasing modulation period $`a`$. The local approximation fails at higher energies when the width of the wavefunctions becomes larger than the period of the modulation, and the structure of the energy spectrum changes. Indeed it is well known from the limit of very weak magnetic modulation, $`s1`$, that in contrast to this local approximation the bands become flat at the energies $`E_\lambda /V_{\mathrm{c}yc}=\pi ^2(\lambda +1/4)`$, for $`\lambda =1,2,\mathrm{}`$. These flat band conditions are the quantum equivalents to the classical commensurability conditions leading to the Weiss oscillations in magnetotransport, and do not change their positions in a plot like Fig. 3(a), even if we change the modulation period. A larger modulation period $`a`$ just leads to a higher density of the energy bands. In Fig. 3(c) we plot for $`\xi _0=\pi /2`$ the effective potential and the square of the energy eigenfunctions for the energy values considered in Fig. 1. Width and location of the wavefunctions in the effective potential is in close agreement with that of the corresponding classical orbits. In Fig. 3(d) we plot the corresponding quantities for the symmetric situation $`\xi _0=\pi `$, to be compared with Fig. 2. These wavefunctions belong to (relative) extrema of the energy bands, and thus have zero group velocity, in agreement with the zero drift velocity of the corresponding classical orbits. The wavefunctions in Fig. 3(c) belong to finite energy dispersion and describe motion in the positive ($`n`$=3) and the negative ($`n`$=43) $`y`$ direction, respectively, in agreement with the correponding classical orbits in Fig. 1. For large quantum numbers $`n`$ and weak modulation the group velocities can be shown to reduce quantitatively to the drift velocities of the corresponding classical orbits. In Fig. 4 we consider the “critical” situation $`s=1`$. The derivatives $`\xi _0^\pm (\xi _{ex})=0`$ and $`\xi _0^{\pm \prime \prime }(\xi _{ex})=0`$ vanish for $`\xi _{ex}=(2p+1)\pi `$ ($`p`$ integer), i.e. for the positions where the magnetic field vanishes, Eq.(12). For all values of $`\xi _0`$ the effective potential (7) becomes flat at these points $`\xi _{ex}`$ (see Fig. 4(c)). The classical situation is as for $`s<1`$ with the exception that for $`\xi _0=\xi _{ex}\pm KR_0`$ there are “critical” orbits which asymptotically approach straight lines parallel to the $`y`$ axis on their left (for $`+`$) or their right (for $``$) side, where $`B(x)=0`$. The dashed lines plotted over the energy spectrum of Fig. 4(a) show the evolution of the flat regions of the effective potential with $`\xi _0`$, i.e. the parabolas resulting from $`V(\xi _{ex};\xi _0)`$ with $`p=0`$ and $`p=\pm 1`$. In the first Brillouin zone these lines are seen as the back-folding of the lowest parabola centered on $`\xi _{ex}`$ with $`p=0`$, and they are an indication of a kind of a free electron motion along the lines where the magnetic field is zero. Close to these parabolas the energy bands have large dispersion near inflexion points, and the energy separation between adjacent bands is minimum. Similar features have been obtained in the energy spectra for single magnetic wells by Peeters and Matulis. In other words, such states experience a weak effective magnetic field, due to the constant effective potential over a substantial spatial region. The wavefunctions corresponding to states with large energy dispersion have large amplitudes at the positions of flat effective potential (vanishing total magnetic field). This is demonstrated for two selected states \[($`n`$=43, $`\xi _0`$=3.18) and ($`n`$=5, $`\xi _0`$=0.50)\] in Fig. 4(b), together with the probability distributions of the corresponding classical orbits. The effective potentials together with the corresponding classical orbits are plotted in Fig. 4(c). The trajectory corresponding to the state ($`n`$=5, $`\xi _0`$=0.50) is close to a critical orbit with a critical right turning point. This leads to an enhanced probability density near that point, which is also reflected in the quantum mechanical probability density. We will see that for slightly stronger modulation a new type of nearly free motion occurs with energies close to the parabolas $`V(\xi _{ex};\xi _0)`$ in the energy spectrum. ### B Strong modulation, $`s>1`$ For $`s>1`$, $`\xi _0^\pm (\xi )=0`$ at $`a(\xi )\mathrm{cos}\xi =1/s`$, and $`\xi _0^\pm (\xi )`$ has extrema at the following positions: $`\mathrm{minima}:\xi _{min}^{(p)}`$ $`=`$ $`(2p+1)\pi +\delta ,`$ (14) $`\mathrm{maxima}:\xi _{max}^{(p)}`$ $`=`$ $`(2p+1)\pi \delta ,`$ (15) where $`p`$ is an integer and $`\delta =\mathrm{arctan}\sqrt{s^21}`$. The values at these extrema are $`\xi _0^\pm (\xi _{min}^{(p)})`$ $`=`$ $`(2p+1)\pi g(s)\pm KR_0`$ (16) $`\xi _0^\pm (\xi _{max}^{(p)})`$ $`=`$ $`(2p+1)\pi +g(s)\pm KR_0,`$ (17) where $$g(s)=\sqrt{s^21}\mathrm{arctan}\sqrt{s^21}>0.$$ (18) The effective potential $`V(\xi ;\xi _0)`$ has extrema of the first kind, Eq. (12), at the same positions. The extrema with values $`V(\xi _{min}^{(p)};\xi _0)=V_{\mathrm{cyc}}[(2p+1)\pi g(s)\xi _0]^2`$ are minima if $`(2p+1)\pi >\xi _0`$, and maxima otherwise, and those with values $`V(\xi _{max}^{(p)};\xi _0)=V_{\mathrm{cyc}}[(2p+1)\pi +g(s)\xi _0]^2`$ are maxima if $`(2p+1)\pi >\xi _0`$, and minima otherwise. #### 1 Classical approach The number of zeroes of $`\xi _0^\pm (\xi )\xi _0`$ depends on both $`s`$ and $`\xi _0`$. If $`g(s)<\pi `$, $`\xi _0^\pm (\xi )\xi _0`$ has at most three zeroes. If $`\xi _0=\xi _0^\pm (\widehat{\xi })`$ for any $`\widehat{\xi }`$ satisfying $`\xi _{max}^{(p)}<\widehat{\xi }<\xi _{min}^{(p)}`$, i.e. if $`(2p+1)\pi g(s)<\xi _0KR_0<(2p+1)\pi +g(s)`$, $`\xi _0^\pm (\xi )\xi _0`$ has three zeroes. The same argument holds for Eq. (13), i.e. the effective potential has three zeroes. For $`(2p1)\pi +g(s)<\xi _0KR_0<(2p+1)\pi g(s)`$, on the other hand, there exists only a single zero. In Fig. 5 we show, for $`s=2`$ \[ i.e. $`g(s)`$=0.685\], an example where the effective potential has a single zero near $`\xi /2\pi =0.1`$, so that for sufficiently low energy only a single drifting orbit exists. The number and the type of the possible orbits depend on the energy. At the highest energy shown in Fig. 5(a) two orbits exist (solid lines in Fig. 5(d)). There is a drifting cyclotron orbit extending over more than two periods of the modulation, with the left turning point on $`\xi _0^+(\xi )`$ (uppermost curve in Fig. 5(c)) near $`\xi /2\pi =1.1`$, and the right turning point on $`\xi _0^{}(\xi )`$ (bottom curve in Fig. 5(c)) near $`\xi /2\pi =1.3`$. Near the relative minimum of $`\xi _0^{}(\xi )`$ close to $`\xi /2\pi =1.7`$, which corresponds to a relative minimum of the effective potential (thick dashed line in Fig. 5(a)), there exists a “channeled orbit” moving in positive $`y`$ direction. We define channeled orbits as trajectories which have both turning points either on $`\xi _0^+(\xi )`$ or on $`\xi _0^{}(\xi )`$, in contrast to the drifting orbits with one turning point on $`\xi _0^+(\xi )`$ and the other on $`\xi _0^{}(\xi )`$. In contrast to the self-intersecting drifting orbits, the channeled orbits are always confined to less than a single modulation period, and they move without self-intersections in a relatively narrow interval of angles around the positive or the negative $`y`$ direction \[see Fig. 5(d)\]. Note that the curvature of the trajectories changes sign at the positions where the total magnetic field vanishes, see Fig. 5(b). If we lower the energy to $`E/V_{\mathrm{cyc}}=40`$, we arrive in Fig. 5 at a situation where only a single drifting orbit exists (dashed lines). In general, the extent in $`\xi `$ direction of the drifting orbits decreases with decreasing energy. At the lowest energy indicated in Fig. 5 (lowest dotted line in (a) and innermost lines in (c)), we have again a drifting orbit near $`\xi /2\pi =0.1`$ and a channeled orbit around $`\xi /2\pi =0.6`$. At this low energy, the extent of the drifting orbit is considerably smaller than the modulation period. In Fig. 6 we show, for the same modulation strength, $`s=2`$, a situation, $`\xi _0=\pi `$, where the effective potential has three zeroes, as is emphasized in the inset of Fig. 6(a). These zeroes are separated by two shallow maxima. If a (positive) $`E`$ value below these maxima is chosen, one finds three narrow drifting cyclotron orbits located around the zeroes of the effective potential (solid lines). For higher energies one may find either one drifting and two channeled orbits (dotted lines) or a single drifting orbit (e.g. for $`0.5<E/V_{\mathrm{c}yc}<30`$, not indicated in the figure). Actually the “drifting” orbits located around $`\xi =\pi `$ have zero drift velocity due to symmetry reasons. In summary, for $`0<g(s)<\pi `$ we find for given values of the constants of motion, $`\xi _0`$ and $`E`$, at least one and at most three orbits. For larger values of $`s=B_m^0/B_0`$, more orbits may exist for a given pair of $`\xi _0`$ and $`E`$ values. A careful analysis of the extrema of the functions $`\xi _0^\pm (\xi )`$ shows, e.g., that for $`\pi <g(s)<2\pi `$ between three and five orbits belong to the same pair of $`\xi _0`$ and $`E`$. We will come back to this case below. Apparently the plots of the effective potential $`V(\xi ;\xi _0)`$ are very useful to see which orbits are possible for a fixed value of the center coordinate $`\xi _0`$ and different energies. Channeled orbits exist in side valleys near relative minima of $`V(\xi ;\xi _0)`$. If, on the other hand, the energy of the motion is given, the plots of the locations of turning points $`\xi _0^\pm (\xi )`$ is very useful to classify the possible orbits for different values of $`\xi _0`$. Channeled orbits exist near relative minima of $`\xi _0^{}(\xi )`$ and relative maxima of $`\xi _0^+(\xi )`$. #### 2 Quantum calculation The energy spectra become more complicated in the case $`s>1`$, Fig. 7 and 8. Regions of different character can be distinguished in these spectra. Areas, where the energy bands are nearly parallel lines with low dispersion (region I), alternate with regions, where steep bands with large dispersion seem to cross bands with weak dispersion (region II). In fact the energy bands never cross each other and the apparent intersections are anti-crossing points with exponetially small gaps. The boundaries of these regions are given by classical values only. If the energy is scaled by the classical cyclotron energy $`V_{\mathrm{c}yc}`$, for fixed modulation strength $`s`$ the regions II are surrounded by the parabolas $`E=V(\xi _{\mathrm{m}in}^{(p)};\xi _0)`$ and $`E=V(\xi _{\mathrm{m}ax}^{(p)};\xi _0)`$ which, for $`p2`$, are indicated by dashed lines in the spectra. For any fixed $`\xi _0`$ such a pair of parabolas gives the minimum and the maximum value of a certain side valley of the effective potential $`V(\xi ;\xi _0)`$ (extrema of the first kind, see Eq. (12)). The energy interval in between these values indicates the depth of that valley, i.e. an energy range in which classically channeled orbits exist, in addition to the drifting orbits. In Fig. 7(b) the effective potential is plotted for the symmetric case $`\xi _0=\pi `$ (dotted line), corresponding to the classical situation described in Fig. 6. Also shown are the states for $`n=0`$ (lower solid line) and for $`n=20`$ (upper solid line) and $`n=22`$ (upper dashed line). Apparently, state $`n=20`$ corresponds to a classical drifting orbit, whereas $`n=22`$ is the symmetric superposition of two states corresponding to channeled orbits in the side valleys. The latter has practically the same energy as the corresponding antisymmetric superposition ($`n=21`$), which is not shown. On the scale of Fig. 7(a), all states $`n=20`$, 21, and 22 seem to have the same energy, $`E/V_{\mathrm{cyc}}38`$. The states $`n=21`$ and 22 are hybridizations of states belonging to the branches with high energy dispersion and opposite sign of the group velocity $$v_y=\frac{K}{m\omega _0}\frac{dE_{n\xi _0}}{d\xi _0}.$$ (19) Figure 7(c) shows the effective potential for the asymmetric case $`\xi _0/2\pi =0.2`$, corresponding to the classical situation described in Fig. 5. Here we show six states, the ground state $`n=0`$ located near the zero of the effective potential, the two “channeled” states “bound” in the potential valley around $`\xi /2\pi =0.6`$, the extended “drifting” state $`n=25`$ near $`E=40V_{\mathrm{cyc}}`$, the extended state $`n=44`$, and the localized “channeled” state $`n=45`$. The energies of all these states are indicated in Fig.7(a). The states which extend over more than a period of the modulation belong to weakly dispersive energy bands and correspond to the classical drifting orbits. The states belonging to the energy branches with strong dispersion have large amplitudes in side valleys of the effective potential and vanish practically outside these valleys. They correspond to classical channeled orbits. The apparent number of nodes of the large-amplitude parts of these “channeled” states increases with energy as if they were truly bound states in these narrow valleys. Note, however, that the wave functions of channeled states still have $`n`$ nodes, but the corresponding oscillations are not observable at the scale of the figure. Outside the valleys, the maxima in between the nodes are a few orders of magnitude smaller than the main peaks inside the valleys. The number of quantized states within a given valley of the effective potential depends on the average magnetic field and the period of the magnetic modulation, even if the parameters $`V_{\mathrm{cyc}}`$ and $`s`$ are fixed. In Fig.7 the period $`a`$, or the field $`B_0`$, is too small to have states quantized in the low-energy triple minimum of the effective potential for the symmetric case of Fig.7(b) (see also Fig.6). To investigate this situation, we show in Fig.8(a) a denser spectrum for the same modulations strength $`s=2`$. In Fig.8(b) the spectrum near $`\xi _0=\pi `$ is enlarged. Five energy values are indicated, and in Fig.8(c) the corresponding (squares of the) wavefunctions are plotted for the states $`n=0`$, 2, 3, 4, and 5, together with the effective potential. The antisymmetric state $`n=1`$, which is nearly degenerate with $`n=2`$, is not shown. This demonstrates that all the classical features have their quantum analog, provided the model parameters (here $`\alpha `$) are suitably chosen. For the magnetic cosine modulation, the depth of the valleys of the effective potential, $$|V(\xi _{\mathrm{m}ax}^{(p)};\xi _0)V(\xi _{\mathrm{m}in}^{(p)};\xi _0)|/V_{\mathrm{c}yc}=4g(s)|(2p+1)\pi \xi _0|,$$ (20) increases with the energy (i.e. with $`|p|`$ for fixed $`\xi _0`$), and thus more and more channeled states appear at higher energies. For sufficiently high energies the strips with channeled states in the energy spectra may thus extend over the whole Brillouin zone. This will also happen for sufficiently large $`s`$. The energy dispersion of the channeled states depends strongly, nearly quadratically, on $`\xi _0`$ according to Eq.(10), which expresses the nearly free motion of the electrons on channeled orbits in $`y`$ direction. For $`s>1`$ and $`g(s)<\pi `$, the area of the regions II of the spectrum increases with increasing $`s`$, and the area of the regions I shrinks accordingly. For $`g(s)=\pi `$, one has $`V(\xi _{\mathrm{m}in}^{(p)};\xi _0)=V(\xi _{\mathrm{m}ax}^{(p1)};\xi _0)`$ and the corresponding parabolas coincide, leaving no room for regions I. If the modulation is so large that $`g(s)\pi `$, drifting and channeled states coexist everywhere in the spectrum. In Fig. 9 we have chosen $`s=5`$, corresponding to $`g(s)=3.53`$. Close to the edges of the Brillouin zone, e.g. for $`\xi _0/2\pi =0.016`$, Fig. 9(c), we can identify channeled, e.g. $`n=22`$ and $`n=16`$, and drifting states, e.g. $`n=19`$. But now these drifting states are relatively narrow, confined in local minima of the effective potential and not in the wide potential well centered around $`\xi _0`$, which is given by the confinement due to the average magnetic field. This case is already known from the discussion of Fig. 6. The local minimum of the effective potential at $`\xi =0`$ is a minimum of the second kind, with vanishing potential. Consequently, these drifting states are similar to weakly perturbed Landau levels with energy gaps $`\mathrm{}eB(\xi _0)/m`$, as can be observed by a careful look at Fig. 9(a). In the center of the Brillouin zone say for $`\xi _0/2\pi =0.493`$, Fig. 9(d), the effective potential has three zeroes of the second kind (see Eq.(13)) near $`\xi =\xi _0`$. We therefore can find similar narrow drifting states, like $`n=5`$ and $`n=6`$, but also wide drifting states at higher energies, like $`n=17`$ and channeled states in local minima of first kind, like $`n=7`$. As in the classical picture, the velocity of these narrow drifting states is in general lower than that of the channeled states. ## IV Mixed harmonic modulations For sufficiently strong mixed electric and magnetic modulations one expects a similar situation as for the strong magnetic modulation, with a coexistence of channeled and drifting orbits, and their quantum analogs. In the presence of an electric modulation, we have no explicit analytic expressions for the minima of the effective potential, Eq. (7), not even for simply harmonic modulations. Nevertheless, a qualitative understanding of the classical and the corresponding quantum mechanical motion is possible. For a given constant of motion $`\xi _0`$ the effective potential has side valleys with possible channeled orbits, if $`V(\xi ;\xi _0)/\xi =0`$ has more than one solution $`\xi `$. This is the case, if the function $$\xi _0(\xi )=\xi +sa(\xi )\frac{w}{2}\frac{u^{}(\xi )}{1+sb(\xi )},$$ (21) with $`w=V_0/V_{\mathrm{cyc}}`$, assumes the value $`\xi _0(\xi )=\xi _0`$ at more than one $`\xi `$ value. At such $`\xi `$ values the effective potential has extrema with the values $$V(\xi ;\xi _0(\xi ))/V_{\mathrm{cyc}}=wu(\xi )+\left[\frac{w}{2}\frac{u^{}(\xi )}{1+sb(\xi )}\right]^2.$$ (22) To be specific, we choose for the following $`b(\xi )=\mathrm{cos}\xi `$, $`a(\xi )=\mathrm{sin}\xi `$ and $`u(\xi )=\mathrm{cos}(\xi +\phi _s)`$. Apparently, Eqs. (21) and (22) provide a parametric representation of the possible relative extrema of the effective potential in the energy-versus-$`\xi _0`$ diagram, similar to the dashed lines in Figs. 7 and 8 which define the regions II where “channeled” states coexist with “drifting” ones. Due to the symmetries $`\xi _0(\xi +2\pi )=\xi _0(\xi )+2\pi `$ and $`V(\xi +2\pi ;\xi _0(\xi +2\pi ))=V(\xi ;\xi _0(\xi ))`$, it is sufficient to consider only one period $`0\xi 2\pi `$ of the parameter $`\xi `$, provided the $`\xi _0(\xi )`$ values are back-folded into the first Brillouin zone. For strong magnetic modulation ($`s>1`$) the denominators in Eqs. (21) and (22) lead to poles. Then the regions of type II extend to arbitrary high energies, similar to the case of pure magnetic modulation. A qualitatively different behavior is obtained for weak magnetic, but arbitrarily strong electric modulation, since then the denominators of Eqs. (21) and (22) remain positive (for $`s<1`$). Consequently, for given values of $`w`$, $`s`$, and $`\phi _s`$ the possible values of $`V(\xi ;\xi _0(\xi ))`$, Eq. (22), are bound and channeled orbits can exist only below a certain energy. ### A Pure electric modulation As a particularly simple example we consider a pure electric cosine modulation, $`s=0`$, $`\phi _s=0`$. For weak modulation ($`w<2`$) the function $`\xi _0(\xi )`$ of Eq. (21) has a unique inverse, i.e. the effective potential $`V(\xi ;\xi _0)`$ has for all values of $`\xi _0`$ only a single extremum, namely the absolute minimum, and no “channeled” states should be expected. The energy spectra for this weak-modulation limit are well known and will not be reproduced here. Apart from a phase shift, they look similar to Fig. 3(a) but with flat bands near $`E/V_{\mathrm{c}yc}=\pi ^2(\lambda 1/4)`$, for $`\lambda =1,2,\mathrm{}`$. The corresponding classical trajectories at sufficiently high energies are drifting cyclotron orbits. At very low energies, $`E<V_0=wV_{\mathrm{c}yc}`$, a peculiarity occurs, since then the classical trajectories are captured within a single valley of the electric potential, with turning points given by Eq. (9) in the interval $`\xi _c\xi 2\pi \xi _c`$ (modulo $`2\pi `$) with $`\xi _c=\mathrm{arccos}(E/V_0)>0`$. For $`\xi _0`$ values in the interval $`\xi _c<\xi _0<2\pi \xi _c`$ these trajectories are self-intersecting drifting orbits, whereas for $`\xi _0<\xi _c`$ and $`\xi _0>2\pi \xi _c`$ there exist channeled orbits with $`v_y>0`$ and $`v_y<0`$, respectively. The orbit with $`\xi _0=\xi _c`$ approaches the left turning point at $`\xi =\xi _c`$ with a tangent parallel to the $`x`$ axis, and that with $`\xi _0=2\pi \xi _c`$ does the same at the right turning point $`\xi =2\pi \xi _c`$. This peculiar low-energy behavior is, of course, not restricted to the weak modulation limit, but occurs always when the trajectories are captured in a minimum of the electric potential, i.e. for $`E<V_0`$. It is demonstrated in Fig. 10(b), and will not be discussed further. If the electric modulation is strong enough, $`w>2`$, the function $`\xi _0(\xi )=\xi (w/2)\mathrm{sin}\xi `$, Eq. (21), has extrema at $`\xi _+=\mathrm{arccos}(2/w)>0`$ and $`\xi _{}=\xi _+`$ (modulo $`2\pi `$) with values $`\xi _0(\xi _\pm )=g(w/2)`$, where $`g(s)`$ is defined by Eq. (18). Then, for $`|\xi _0|g(w/2)`$ the equation $`\xi _0(\xi )=\xi _0`$ has three solutions $`\xi `$ in the interval $`|\xi |<\pi `$, which are local extrema of the effective potential with values $$V(\xi ;\xi _0(\xi ))/V_{\mathrm{cyc}}=1+(w/2)^2[1(w/2)\mathrm{cos}\xi ]^2.$$ (23) In order to find in the energy spectra the regions II corresponding to side valleys of the effective potential, one may proceed as follows. One plots in the extended zone scheme $`V(\xi ;\xi _0(\xi ))`$ versus $`\xi _0(\xi )`$, starting at $`\xi =\pi `$, where $`\xi _0(\xi )=\pi `$ and $`V(\xi ;\xi _0(\xi ))=V_0`$. With increasing $`\xi `$, also $`\xi _0(\xi )`$ and $`V(\xi ;\xi _0(\xi ))`$ increase and reach at $`\xi =\xi _{}`$ their maximum values $`g(w/2)`$ and $`V_{\mathrm{cyc}}(1+w^2/4)`$, respectively. As $`\xi `$ increases from $`\xi =\xi _{}`$ to $`\xi =0`$, $`\xi _0(\xi )`$ and $`V(\xi ;\xi _0(\xi ))`$ decrease towards the values $`0`$ and $`V_0`$, respectively. Increasing $`\xi `$ from 0 to $`\pi `$ leads to the mirror image of the described trace with respect to $`\xi _0=0`$: $`\xi _0(\xi )=\xi _0(\xi )`$ and $`V(\xi ;\xi _0(\xi ))=V(\xi ;\xi _0(\xi ))`$. Finally these four line segments have to be folded back into the “first” Brillouin zone $`0\xi _02\pi `$ to obtain the absolute minimum of the effective potential as a function of $`\xi _0`$ and the boundaries of the regions II. In contrast to the strong magnetic modulation, these regions become narrower with increasing energy and end at $`\xi _0=\pm g(w/2)`$ (modulo $`2\pi `$) with energy $`E/V_{\mathrm{cyc}}=1+w^2/4`$. For $`g(w/2)>\pi `$ the back-folding leads to an overlap of different branches of the region II, that is the coexistence of back and forth running “channeled” states with “drifting” states in the same area of the $`E`$-$`\xi _0`$ diagram. Figure 10 shows for a typical example the quantum mechanical energy spectrum together with the boundaries of region II obtained in this manner. The “very complicated” energy spectrum obtained recently by Shi and Szeto for strong electric modulation is thus explained by the coexistence of channeled and drifting states. In previous work it was pointed out that, for given modulation period $`a`$ and strength $`V_0`$ and given energy $`E=E_\mathrm{F}`$, channeled orbits can exist only if the magnetic field $`B_0`$ is smaller than a critical field $`B_{\mathrm{crit}}`$. Solving $`E_\mathrm{F}/V_{\mathrm{cyc}}=1+w^2/4`$ for $`E_\mathrm{F}>V_0=wV_{\mathrm{cyc}}`$ and $`w>2`$ with respect to the magnetic field, one obtains the known result $$B_{\mathrm{crit}}=\frac{2\pi V_0}{eav_\mathrm{F}}\left[\frac{2}{1+\sqrt{1(V_0/E_\mathrm{F})^2}}\right]^{1/2}.$$ (24) ### B Weak magnetic modulation If a magnetic modulation is added to an electric one, very complicated interference effects may result. Only if the phase shift $`\phi _s`$ is zero or $`\pi `$, the resulting energy spectrum will be symmetric in $`\xi _0`$. Even in that case, the distribution of channeled states (regions II) in the $`E`$-$`\xi _0`$ diagram may become rather complicated, especially at low energies. For the mixed case channeled states may occur even if the modulation parameters $`w`$ and $`s`$ are not large enough to produce them for the pure electric and the pure magnetic modulation of these strengths. For weak magnetic modulation, $`0<s<1`$, and arbitrary strength of the electric modulation, channeled states can exist only below a certain energy, as in the pure electric modulation case. For arbitrary phase shift $`\phi _s`$ the energy spectrum may be so asymmetric that in a certain energy range only channeled orbits exist which carry current in one (say the positive $`y`$-) direction, but no channeled orbits carrying current in the opposite direction. Such a situation is presented in Fig. 11. The regions II, where channeled and drifting states coexist, is again calculated from Eqs. (21) and (22), i.e. from purely classical arguments. If a 2DEG is subjected to such an asymmetric mixed modulation, it may happen that in the thermal equilibrium the channeled states carry a finite current. Of course, this current must be compensated by a corresponding opposite current carried by the drifting states. For large magnetic modulation, $`s>1`$, and arbitrary electric modulation, the magnetic modulation dominates the energy spectra at large energies. The regions II with channeled states become more and more important, as can be seen from the pole structure of Eqs. (21) and (22). For very weak electric modulation, the results reduce to those of the pure magnetic modulation, apart from some peculiarities at very low energies, where additional regions of channeled orbits may exist. In magnetically modulated systems prepared by deposition of magnetic micro-strips there is always an induced electric modulation due to the interface stress between the ferromagnets and the semiconductor. The phase shift with respect to the magnetic modulation occurs when the external magnetic field is tilted. Nevertheless, in the known experimental situations the stress potential amplitude is presumably much weaker than our bare potential. ## V Summary and discussion We have discussed in detail the quantum electronic states and energy spectra $`E_n(X_0)`$ of a 2DEG in strong one-dimensional magnetic and electric superlattices, and in a non-vanishing average external magnetic field. By comparing the quantum results with the corresponding characteristics of the classical motion, we achieved a detailed and intuitive understanding of the energy spectra and eigenstates. We found that the complicated parts of the energy spectra (“regions II”), where branches with strong dispersion coexist with those of low dispersion, coincide with the areas in the $`EX_0`$ diagram in which classically “channeled” orbits exist. For a systematic investigation of the possible energy spectra and eigen states, and of the corresponding types of classical trajectories, it is useful to exploit the scaling properties of the Hamiltonian. Then it is not necessary to vary independently all the basic model paramaters, i.e. the strengths $`B_m^0`$ and $`V_0`$ of magnetic and electric modulation, the modulation period $`a`$, and the average magnetic field $`B_0`$. If one uses suitable units for energy and length, $`V_{\mathrm{cyc}}`$ and $`a/2\pi `$, respectively, one obtains the same classical results and the same gross features of the energy spectra (the same position of the regions II), if one changes the four parameters $`B_m^0`$, $`V_0`$, $`a`$, and $`B_0`$ in such a manner that the two reduced modulation strengths $`s=B_m^0/B_0`$ and $`w=V_0/V_{\mathrm{cyc}}`$ remain constant. For different parameter sets with the same values of $`s`$ and $`w`$, only the density of the energy bands is different in the plot of $`E_n(X_0)/V_{\mathrm{cyc}}`$ versus $`KX_0`$, not its overall appearance. This is illustrated by Figs. 7(a) and 8(a), for which the regions II coincide. The reason for this behaviour is that, in these energy and length units, the effective potential is invariant under the scaling transformation $`B_m^0\gamma B_m^0`$, $`B_0\gamma B_0`$, $`a\lambda a`$ and $`V_0\gamma ^2\lambda ^2V_0`$, for arbitrary positive $`\gamma `$ and $`\lambda `$. To leave the quantum result exactly unchanged under a change of the four model parameters, one has to keep $`\alpha =(l_0K)^4`$ also unchanged. This is because only with the restriction $`\lambda =1/\sqrt{\gamma }`$ the kinetic energy operator is also independent of the scaling parameter $`\gamma `$ \[see Eq. (8)\]. Thus, in the suitable units, the exact quantum result depends only on three independent parameters instead of four, and the characteristic classical features depend only on two. There is a close correspondence between the quantum states belonging to strong-dispersion branches of the energy spectrum and the classical channeled orbits. These orbits occur near lines of vanishing total magnetic field or near minima of the electric modulation potential and are restricted to individual side valleys of the effective potential. They are always restricted to a part of a single modulation period in $`x`$ direction and represent a fast motion along (wavy) lines without self-intersections in positive or negative $`y`$ direction. The corresponding quantum states are also essentially confined to the same space region and belong to energy branches with strong dispersion. At a given value of the constant of motion $`X_0`$, channeled orbits exist in energy intervals bounded by adjacent relative minima and maxima of the effective potential, defining bottom and top of the corresponding side valley. Plotting these classically defined extrema versus $`X_0`$, one obtains the boundaries of the regions II of the quantum energy spectrum. Classically, for each channeled orbit there exists a “drifting” orbit with the same constants of motion $`X_0`$ and $`E`$. These drifting orbits are self-intersecting trajectories which, for sufficiently large energy, extend over more than one modulation period in $`x`$ direction and drift slowly in $`y`$ direction. The corresponding quantum states belong to low-dispersion branches of the energy spectrum. Quantum mechanically, the channeled states do not appear at exactly the same energies as the drifting states, and they usually have a larger energy spacing than the latter, since they are confined to a narrower effective potential well. We have demonstrated these features by model calculations based on simple harmonic modulation fields. Qualitatively the obtained results and the methods to derive them can easily be extended to more general modulation fields, containing higher harmonics. This will be necessary, if the distance of the 2DEG from the sample surface is not much larger than the period of the surface structure creating the modulation. Anharmonic effective modulation potentials may also result from non-linear screening effects, even if the bare modulation potential is harmonic. We have also performed several additional calculations and consistency checks which are not documented in the main text. E.g., we have checked the equivalence of classical drift velocity and quantum group velocity beyond the analytically accessible case of very weak modulation fields. For some examples with strong modulation, we evaluated the quantum mechanical group velocity along several energy bands $`E_n(X_0)`$ and compared the result with the drift velocity of the corresponding classical trajectories with the same energy and $`X_0`$ values. For most parts of the bands the two velocities agreed perfectly. A systematic deviation was observed only in parameter regimes where the classical trajectories are close to critical orbits, which have no quantum analog. Near the critical orbits the modulus of the classical drift velocity increases rather rapidly, whereas the quantum mechanical group velocity shows no anomaly. We have also extended the band structure calculations to very strong magnetic modulation ($`B_m^0/B_0=20`$). While at high energies a complicated superpositon of bands with steep and with flat dispersions, similar to that in Fig. 9(a), was obtained, the bands at low energies tend to cluster into groups separated by relatively large gaps. The low-energy part of the spectrum was already reminiscent of the spectrum for vanishing average magnetic field, where one obtains a one-dimensional Bloch energy spectrum for each value of $`p_y=eB_0X_0`$. Concerning previous and forthcoming transport calculations, we conclude from the close correspondence of the quantum and the classical approach, that at weak average magnetic fields the classical calculations are appropriate, provided the modulation fields are not too strong. For the very strong magnetic modulation mentioned in the introduction, it may however happen, that the energy level spacing of “channeled orbits” exceeds the thermal energy $`k_BT`$ in a regime where $`\mathrm{}\omega _0k_BT`$. Then we would expect modulation induced quantum effects in the positive-magnetoresistance regime at low $`B_0`$. ###### Acknowledgements. We thank D. Pfannkuche for critical reading of the manuscript. This work was supported by the German Bundesministerium für Bildung und Forschung (BMBF), Grant No. 01BM622. One of us (A.M.) is grateful to the Max-Planck-Institut für Festkörperforschung, Stuttgart, for support and hospitality. Fig.1 Zwerschke Fig.2 Zwerschke Fig.3 Zwerschke Fig.4 Zwerschke Fig.5 Zwerschke Fig.6 Zwerschke Fig.7 Zwerschke Fig.8 Zwerschke Fig.9 Zwerschke Fig.10 Zwerschke Fig.11 Zwerschke
no-problem/9906/cond-mat9906417.html
ar5iv
text
# Topological Invariants in Microscopic Transport on Rough Landscapes: Morphology, Hierarchical Structure, and Horton Analysis of River-like Networks of Vortices \[ ## Abstract River basins as diverse as the Nile, the Amazon, and the Mississippi satisfy certain topological invariants known as Horton’s Laws. Do these macroscopic (up to $`10^3`$ km) laws extend to the micron scale? Through realistic simulations, we analyze the morphology and hierarchical properties of networks of vortex flow in flux-gradient-driven superconductors. We derive a phase diagram of the different network morphologies, including one in which Horton’s laws of length and stream number are obeyed—even though these networks are about $`10^9`$ times smaller than geophysical river basins. \] Introduction.— The nature of river basins, including their physical structure and evolution, has been a problem of major interest to civilized societies throughout history. Horton’s laws are perhaps one of the most intriguing properties of river networks . In order to apply them to a network, the individual streams composing the network must be identified and labeled with an order number, as in the top left corner of Fig. 1(a). The lowest order streams are the smallest outlying tributaries on the edges of the network, according to the Strahler ordering scheme. At each point where two tributary streams join, a new stream begins. Whenever two tributaries of the same order meet, the outgoing stream has an order number one higher than that of the tributaries. If two tributaries of different orders meet, the outgoing stream has the same order number as the higher ordered tributary. Eventually, all streams in the network combine to form the highest order (main) stream. The number of streams of order $`w`$ is $`N_w`$, while $`L_w`$ is the average length of streams of order $`w`$. Horton’s laws state that the bifurcation ratio $`R_B`$ and the length ratio $`R_L`$, given by $`R_B=N_w/N_{w+1}`$ and $`R_L=L_{w+1}/L_w`$, are constant, or independent of $`w`$. These ratios also provide the fractal dimension of the rivers $`D_FlogR_B/\mathrm{log}R_L`$. Geophysical river basins typically have values of $`R_B4`$ and $`R_L2`$. Do these (Horton’s) laws apply to microscopic landscapes? Here we present evidence that these macroscopic laws are obeyed at the microscopic scale by river-like networks of flowing quantized magnetic flux. Vortex River Basins.— Near the depinning transition, magnetic vortices in type-II superconductors move in intricate flow patterns that have been seen both in computer simulations and in experiments, including finger-like or dendritic shapes as well as the filamentary flow of vortices in river-like paths and networks (see, e.g., and references therein). Despite the ubiquity of the river-like pathways produced by the vortex motion, very little work has been done towards characterizing the morphology of these flow patterns. Moreover, concepts and ideas used for decades to characterize geophysical river basins have not been applied to the study of the microscopic flow through tree–shaped channel networks. This is surprising since the underlying physics of vortex and geological rivers offers striking similarities: driven non-equilibrium dissipative systems displaying branched (or ramified) transport among metastable states on a rough landscape . One is driven by the Lorentz force and the other by gravity. Like geophysical rivers, vortex flow basins exhibit sinuosity (i.e., tortuosity), anabranching, braiding, occasional sudden floods, and other features that make them remarkably similar to geophysical rivers . Indeed, some satellite photographs of river basins are strikingly similar to the channels produced by vortex motion. However, significant differences also exist, including: flow direction, quantized flux flow versus continuum water flow, compressible vortex lattice versus incompressible fluid, negligible inertia with overdamped vortex dynamics versus massive fluid, non-erosional versus erosional landscape, peripheral flux sources versus uniform rain, and correlated long-range versus short-range interactions (so the rapidly varying vortex–vortex repulsion landscape smoothes out the underlying static pinscape). This strongly-correlated vortex dynamics generates flux motion that can be either continuous-flow type, like water, or intermittent stick-slip-type motion—depending on the balance of forces. Also, vortices typically move over relatively flat landscapes with many divots, as opposed to the mountain-range-like very rough landscapes of some geophysical rivers. Moreover, vortex river basins occur inside materials at approximate scales between 1 to 100 $`\mu `$m, much smaller than geophysical river basins (of up to $`10^3`$ km)—and also spanning a smaller range of length scales. Thus, given these numerous similarities and differences, it is very unclear a priori which macroscopic results carry over to the microscopic domain. By conducting realistic simulations of slowly driven vortices moving over many samples, we have identified several distinct network phases. These vortex basins appear in the initial penetrating front of vortices. Remarkably, we find that: for a wide range of parameters networks of vortex channels obey Horton’s laws just as geophysical river networks do. This is remarkable, given the many physical differences between basins of flux quanta and geophysical rivers and that they move over very different types of potential-energy landscapes. Unlike previous work, here we first present a detailed list of analogies and differences between river basins and networks of vortex channels. Afterwards, we present the first morphological phase diagram for vortex motion. Finally, we analyze the hierarchical structure of the vortex channels. Simulation.— We model a transverse 2D slice (in the $`x`$$`y`$ plane) of an infinite zero-field-cooled $`T=0`$ superconducting slab containing flux-gradient-driven 3D rigid vortices that are parallel to the sample edge . Vortices are added at the surface at periodic time intervals, and enter the superconducting slab under the force of their own mutual repulsion . The slab is $`36\lambda \times 36\lambda `$ in size, where $`\lambda `$ is the penetration depth. The vortex-vortex repulsive interaction is correctly modeled by a modified Bessel function, $`K_1(r/\lambda )`$. The vortices also interact with 972 non-overlapping attractive parabolic wells of radius $`\xi _p=0.3\lambda `$. The density of pins $`n_p`$ is $`n_p=0.75/\lambda ^2`$. All pins in a given sample have the same maximum pinning force $`f_p`$, which ranged from $`f_p=0.3f_0`$ to $`f_p=6.0f_0`$ in thirteen different samples. For each sample type, we considered five realizations of disorder. Thus, the five points at each pinning force, delineating the broad crossover boundary between Hortonian and braided phases in Fig. 2, refer to these five realizations of disorder. A sixth point, indicating the average value from the five trials, is not visible when it overlaps with another point. We measure all forces in units of $`f_0=\mathrm{\Phi }_0^2/8\pi ^2\lambda ^3`$, magnetic fields in units of $`\mathrm{\Phi }_0/\lambda ^2`$, and lengths in units of the penetration depth $`\lambda `$. Here, $`\mathrm{\Phi }_0`$ is the flux quantum. The overdamped equation of vortex motion is $`𝐟_i=𝐟_i^{vv}+𝐟_i^{vp}=\eta 𝐯_i`$, where the total force $`𝐟_i`$ on vortex $`i`$ (due to other vortices $`𝐟_i^{vv}`$, and pinning sites $`𝐟_i^{vp}`$) is given by $`𝐟_i`$ $`=_{j=1}^{N_v}f_0K_1(|𝐫_i𝐫_j|/\lambda )\widehat{𝐫}_{ij}+`$ $`_{k=1}^{N_p}(f_p/\xi _p)|𝐫_i𝐫_k^{(p)}|`$ $`\mathrm{\Theta }\left[\xi _p|𝐫_i𝐫_k^{(p)}|\right]\widehat{𝐫}_{ik}.`$ Here, $`\mathrm{\Theta }`$ is the Heaviside step function, $`𝐫_i`$ ($`𝐯_i`$) is the location (velocity) of the $`i`$th vortex, $`𝐫_k^{(p)}`$ is the location of the $`k`$th pinning site, $`\xi _p`$ is the pinning site radius, $`N_p`$ ($`N_v`$) is the number of pinning sites (vortices), $`\widehat{𝐫}_{ij}=(𝐫_i𝐫_j)/|𝐫_i𝐫_j|`$, $`\widehat{𝐫}_{ik}=(𝐫_i𝐫_k^{(p)})/|𝐫_i𝐫_k^{(p)}|`$, and we take $`\eta =1`$. Morphological Characterization.— In order to identify and characterize the vortex river networks formed as the flux-gradient-driven front initially penetrates the sample, we divide our simulation area into a $`300\times 300`$ grid. Each time a vortex enters a grid element, the counter associated with that element is incremented. All grid elements that are visited at least once by a vortex are considered part of the network . The maximum number of vortices in the sample is approximately 1200. The pinning density $`n_p`$ and radius $`\xi _p`$ were kept constant at $`n_p=0.75/\lambda ^2`$ and $`\xi _p=0.3\lambda `$, while the pinning force $`f_p`$ varied from sample to sample. We also performed additional simulations in which $`f_p`$ was kept constant and $`n_p`$ varied from $`n_p=0.15/\lambda ^2`$ to $`n_p=2.15/\lambda ^2`$. We observed three distinct vortex river network morphologies, depending on the local magnetic field $`B`$ and the pinning force $`f_p`$, as indicated in one of our main results: the “morphological phase diagram” in Fig. 2. In Fig. 1 the vortex trajectories are presented for the three morphologies. In samples with low pinning force values, $`f_p0.75f_0`$ (see Fig. 2), vortices flow throughout the sample, producing dense vortex river basins. These become space-filling for large times—or large fields since the external field is slowly ramped up. An example of the vortex channels in this regime, as they appear after 160000 MD steps, is shown in Fig. 1(c). If the simulation is allowed to proceed for a larger number of MD steps, the channels eventually fill the entire region shown in Fig. 1(c). For stronger pinning, $`f_p1.0f_0`$, and low vortex densities $`B\mathrm{\Phi }_0/\lambda ^23B_\varphi /2`$, we observe branched “Hortonian” river networks that follow Horton’s laws of stream number and length \[see Fig. 1(a)\]. At higher magnetic fields $`B\mathrm{\Phi }_0/\lambda ^23B_\varphi /2`$, the vortex rivers become highly braided or interconnected and are no longer Hortonian in morphology \[see Fig. 1(b)\]. Unlike the dense networks of Fig 1(c), where preferred vortex paths are uncommon, in the braided regime vortices consistently move along certain pathways, while in some areas of the sample vortex motion rarely occurs. For low pinning forces in the dense network regime, vortex motion occurs both interstitially (with the vortices moving only in the areas between pinning sites) and by means of depinning. If vortex depinning is occurring in a landscape with traps of comparable strength, no favored paths for flux motion can form, leading to the observed dense pathways. The Hortonian and braided regimes arise once the pinning is strong enough that predominantly interstitial motion occurs. That is, pinned vortices almost never depin. Other vortices are prevented from moving close to a pinned vortex by the vortex-vortex repulsion, which has a longer (by near two orders of magnitude) range than the attraction of each pinning site. Since there are regions of the sample (i.e., at or near pinned vortices) where flux motion does not occur, the flow of the moving vortices must be concentrated in certain well-defined regions or rivers, leading to the formation of either Hortonian or braided rivers. The broad crossover between Hortonian and braided rivers occurs when the flux density has increased enough that a large fraction of the pinning sites are occupied. In Fig. 2, the crossover region increases from $`B0.9\mathrm{\Phi }_0/\lambda ^2`$ for $`f_p=1.0f_0`$, to $`B1.3\mathrm{\Phi }_0/\lambda ^2`$ for $`f_p=6.0f_0`$. In each case, the crossover occurs at vortex densities higher ($`3B_\varphi /2B<2B_\varphi `$) than the matching field $`B_\varphi =0.75\mathrm{\Phi }_0/\lambda ^2`$, when $`N_p=N_v`$ . This is in agreement with the results for the inset of Fig. 2, which shows that the transition from Hortonian rivers to braided rivers occurs at higher vortex densities as $`n_p`$ (and thereby the matching field) is increased. Additional support for this interpretation comes from examining the fraction $`R_{\mathrm{ups}}`$ of unoccupied pinning sites \[Fig. 3(a)\]. At the matching field, $`B_\varphi =0.75\mathrm{\Phi }_0/\lambda ^2`$, only about $`65\%`$ of the pins are occupied . The pins are not fully occupied until a field of $`B1.4\mathrm{\Phi }_0/\lambda ^22B_\varphi `$ is applied—when the potential energy landscape experienced by the moving vortices becomes much more uniform. Horton Analysis.— In order to determine whether the vortex river networks we observe obey Horton’s laws, we performed Hortonian analysis on five different realizations of disorder for each of eight different pinning forces $`f_p`$ falling within the Hortonian regime. In each trial, a branching river was identified for analysis. The numbers $`N_w`$ and lengths $`L_w`$ of streams of order $`w=1`$ to 4 were recorded. Representative plots of the type used to determine the length ratio, $`R_L=L_{w+1}/L_w`$, and the bifurcation ratio, $`R_B=N_w/N_{w+1}`$, are shown in Fig. 4. In the best fit exponential regressions used to extract $`R_L`$ and $`R_B`$, the average correlation coefficient was $`0.99`$, indicating a good fit to the Hortonian relationships. The average values for $`R_B`$ and $`R_L`$ throughout the Hortonian river region were $`R_B=3.99\pm 0.18`$ and $`R_L=2.04\pm 0.12`$, in excellent agreement with geophysical rivers. . The characteristics of the Hortonian river networks are dependent on the pinning force $`f_p`$. In Fig. 3(b) we plot the length ratios $`R_L`$, and fractal dimensions $`D_F`$, for each pinning force in the Hortonian region. The braching ratio (not shown) is roughly constant as $`f_p`$ is varied. Changing the pinning force alters the ease with which individual vortices can be depinned, and thereby changes $`R_L`$ and $`D_f/d_c`$. As the pinning force decreases, it is more likely that some vortices will be depined and form new pathways of vortex motion. This will decrease the length of the higher order rivers by cutting short how far the vortex channels propagate before bifurcating. Therefore $`R_L`$ will decrease with decreasing $`f_p`$. Since a larger number of paths are created the $`D_f`$ will increase with decreasing $`f_p`$, in agreement with Fig. 3(b). Concluding remarks.— We have analyzed the morphologies of flux-flow channels slowly driven to its marginally stable state, as a function of flux density and disorder strength. We have identified three distinct morphologies which include: a (large $`B`$) dense network regime, where flow can occur anywhere; a braided network regime, where flow is restricted to certain regions; and a (low $`B`$) Hortonian network regime, where Horton’s laws of length and branching ratio are obeyed in agreement with geophysical rivers. Indeed, it seems promising to analyze tree-shaped channel flow at the microscopic level adapting concepts that have already been successful in treating macroscopic river basins. These types of analysis are largely unexplored. The direction and success of such an approach constitutes an open and fascinating area. CJO (APM) acknowledges support from the GSRP of the microgravity division of NASA (NSF-REU). We thank the Maui Supercomputer Center, R. Riolo, and the UM-PSCS for providing computing resources. We thank F. Marchesoni, M. Bretz, E. Somfai, D. Tarboton, and S. Peckham for their comments.
no-problem/9906/astro-ph9906316.html
ar5iv
text
# A disk census for the nearest group of young stars: Mid-infrared observations of the TW Hydrae Association ## 1 Introduction Circumstellar disks appear to be a natural consequence of the star formation process. Observations of young pre-main-sequence (PMS) stars show that many of them are surrounded by optically thick disks of solar system dimension with masses comparable to or greater than the “minimum-mass solar nebula” of 0.01 $`M_{}`$ (see Beckwith 1999 for a review). Infrared emission in excess of stellar photospheric fluxes provides the most readily measurable signature of such disks. Excesses at $`\lambda `$ 10 $`\mu `$m are found in $``$50% of the low-mass stars in star-forming regions (Strom et al. 1993). It has been suggested that circumstellar disks evolve from optically thick to optically thin structures in about 10 Myr (Strom et al. 1993). That transition may mark the assembly of grains into planetesimals, or clearing of the disk by planets. Indeed, low-mass debris disks have now been imaged around several main sequence stars with ages ranging from 10 Myr to 1 Gyr (Jayawardhana et al. 1998; Holland et al. 1998; Greaves et al. 1998; Koerner et al. 1998; Trilling & Brown 1998). However, age estimates for early-type isolated main sequence stars are highly uncertain. Therefore, the timescale for disk evolution and planet formation is still poorly constrained, and may depend critically on the presence or absence of a close binary companion. The recent discovery of a group of young stars associated with TW Hydrae offers a unique laboratory to study disk evolution and planet formation. TW Hya itself was first identified as an “isolated” T Tauri star by Rucinski & Krautter (1983). Subsequent searches by de la Reza et al. (1989) and Gregorio-Hetem et al. (1992) found four other young stars in the vicinity of TW Hya and suggested that the stars may be kinematically associated. On the basis of strong X-ray emission from all five systems, Kastner et al. (1997) concluded that the group forms a physical association at a distance of $``$50 pc with an age of 20$`\pm `$10 Myr. (See Jensen, Cohen, & Neuhäuser 1998 for a different point of view.) Webb et al. (1999) have identified five more T Tauri star systems in the same region of the sky as candidate members of the “TW Hya Association” (TWA), based on the same signatures of youth –namely high X-ray flux, large Li abundance, and strong chromospheric activity– and the same proper motion as the original five members. Furthermore, they suggest that the wide binary HR 4796, which contains an A0V star, is also part of the Association, even though its Hipparcos parallactic distance of 67 pc places it further away than most other members of the group. The three other TWA stars with Hipparcos distances –TW Hya, HD 98800, and TWA 9– are at 56, 47 and 50 pc, respectively. Being the nearest group of young stars, the TW Hya Association is ideally suited for sensitive disk searches in the mid-infrared. Furthermore, its estimated age of $``$10 Myr provides a strong constraint on disk evolution timescales and fills a significant gap in the age sequence between $``$1-Myr-old T Tauri stars in molecular clouds like Taurus-Auriga and Chamaeleon and the $``$50-Myr-old open clusters such as IC 2602 and IC 2391. Over the past two years, we have conducted mid-infrared observations of the candidate TWA stars. Our discovery of a spatially-resolved disk around HR 4796A and our high-resolution observations of the close binary Hen 3-600 have already been reported (Jayawardhana et al. 1998, 1999). Here we present observations of the other members of the group, including the discovery of a possible 10$`\mu `$m excess in CD -337795, and discuss implications for the origin and age of the TW Hya Association as well as for disk evolution timescales. ## 2 Observations During three observing runs in 1998 and 1999, we have obtained mid-infrared images of candidate members of the TW Hya Association using the OSCIR instrument on the 4-meter Blanco telescope at Cerro Tololo Interamerican Observatory (CTIO) and the 10-meter Keck II telescope. The log of our observations is given in Table 1. OSCIR is a mid-infrared imager/spectrometer built at the University of Florida, using a 128$`\times `$128 Si:As Blocked Impurity Band (BIB) detector developed by Boeing. Additional information on OSCIR is available on the Internet at www.astro.ufl.edu/iag/. On the CTIO 4-m telescope, OSCIR has a plate scale of 0.183”/pixel, which gives a field of view of 23”$`\times `$23”. Our observations were made using the standard chop/nod technique with a chop frequency of 5 Hz and a throw of 23” in declination. On Keck II, its plate scale is 0.062”/pixel, providing a 7.9”$`\times `$7.9” field of view. Here we used a chop frequency of 4 Hz and a throw of 8”. Images were obtained in the N(10.8 $`\mu `$m) band for the entire sample, and in the IHW18(18.2 $`\mu `$m) band for a few bright targets. ## 3 Results In Table 2, we present the measured 10$`\mu `$m fluxes for the entire sample, and compare them to the expected photospheric fluxes, assuming K - N=0 for all late-type stars; For HR 4796A, we used K - N=-0.03, as given by Kenyon & Hartmann (1995) for an A0V star. Among the candidate TWA stars, only TW Hya, HD 98800, Hen 3-600A and HR 4796A – all of which were first detected by IRAS – show significant excess at mid-infrared wavelengths. We have detected a modest 10$`\mu `$m excess in CD -337795 for the first time (see below). None of the other late-type stars have excess, suggesting that they do not harbor dusty inner disks. Comments on individual objects TW Hydrae The spectral energy distribution (SED) of TW Hya from near- to far-infrared wavelengths, including our flux measurements at 10.8$`\mu `$m and 18.2$`\mu `$m, is shown in Figure 1a. The excess at $`\lambda `$ 20$`\mu `$m is unusually strong compared to the median of classical T Tauri stars in Taurus (solid line in Figure 1a). It is worth noting that TW Hya also has a large H$`\alpha `$ equivalent width of -220Å, consistent with an actively accreting disk. CD -337795 We measure a 10$`\mu `$m flux of 96$`\pm `$9 mJy for CD -337795, somewhat above its estimated photospheric emission of 70$`\pm `$5 mJy. This modest excess, at a level of 2.6$`\sigma `$, is well below what is expected for an optically thick inner disk (Figure 1b). One possibility is that our assumption K - N=0 is not correct; if K - N$``$0.3, there would be no 10$`\mu `$m excess. However, we note that using K - N=0 gives good agreement with measured fluxes for other stars of similar spectral type in the sample. If the excess is real, it could be due to an optically thin disk or a faint, as yet undetected stellar companion. We note that CD -337795 may be a spectroscopic binary according to Webb et al. (1999). Furthermore, Lowrance et al. (1999) have reported the discovery of a possible brown dwarf companion 2”, or $``$100 AU, from CD -337795. Our 10$`\mu `$m flux measurement is within an aperture of 1” radius (seeing$``$0.7”), and thus should not include a contribution from this brown dwarf. In any case, to account for the observed 10$`\mu `$m excess, the brown dwarf ($`K=11.5`$) would have to have K - N=3.6! HR 4796B We placed an upper limit of 23 mJy to the 10$`\mu `$m emission from HR 4796B from CTIO data and have now detected its photosphere at 16$`\pm `$2 mJy at Keck. This result is of particular interest because the age of HR 4796B is fairly well established. Using the Hipparcos distance of 67 pc to HR 4796A and D’Antona and Mazzitelli (1994) evolutionary tracks, it is possible to estimate an age of $`8\pm 3`$ Myr for B which is an M2.5 star (Jayawardhana et al. 1998). This age is consistent with the upper bound provided by the measurement of the strong Li absorption line at 6708 Å (Stauffer et al. 1995). The lack of 10$`\mu `$m excess in this object suggests that HR 4796B does not have a dusty inner disk. Unfortunately, we cannot determine at present whether it originally had an optically thick disk, which has since depleted, or whether it formed without a disk (like some 50% of T Tauri stars appear to be). If future far-infrared and sub-millimeter observations find evidence for an outer disk around HR 4796B, that would argue for rapid evolution of the inner disk, either through coagulation of dust or accretion on to the central star. ## 4 Discussion ### 4.1 Origin and age of the TW Hya Association Whether the TWA stars are physically related in origin is a matter of controversy. Kastner et al. (1997) and Webb et al. (1999) argue that this is an unusual grouping of relatively young (10 Myr old) stars unlikely to be a chance coincidence, while Jensen et al. (1998) argue that the proper motions of three primary members – TW Hya, HD 98800, and CD -367429 – are inconsistent with them having formed together 10 Myr ago. Confusing the issue further, TW Hya has all the characteristics of an actively accreting T Tauri star (Rucinski & Krautter 1983; de la Reza et al. 1989; Gregorio-Hetem et al. 1992), typical of much younger systems (ages $`<`$ 3 Myr), while most of the other members show weak or no H$`\alpha `$ emission (Kastner et al. 1997; Webb et al. 1999), and, as we have shown here, little or no infrared excess emission from dusty disks. Our data do not bear directly on the question of the physical connection between the stars in this association, but a few comments can be made about membership determinations. The first question to be addressed is whether the TWA members really constitute a special and unlikely concentration of objects. Webb et al. (1999) argue for this conclusion, based on three points: first, essentially that there are few or no classical T Tauri stars known outside of clouds; second, that X-ray surveys do not show similar numbers of 10 Myr old stars; and third, that there is no evidence for 10-100 Myr old populations in the solar neighborhood. The first argument is not very strong because there is only one bona fide Classical T Tauri star in the TW Hya group and so statistics are poor. The third argument appears to be inconsistent with the results of Briceño et al. (1997), who showed that many of the X-ray bright low-mass stars in the ROSAT All Sky Survey (RASS) have ages of 50-100 Myr. The second argument also has problems; as Martín & Magazzù (1999) showed from a study of Li equivalent widths in RASS-selected stars in the direction of Taurus, while most of the systems are likely to be 50-100 Myr old, a modest fraction ($`20`$%) of these objects may indeed be $`10`$ Myr old. Another way to address the question of the overdensity is to use estimates of the expected average birthrate in the solar neighborhood. Using the results of Miller & Scalo (1979) for a constant birthrate and the age parameter $`T_o=12\times 10^9`$ yr, and their form for the initial mass function, we predict that the number of stars formed per year between $`0.8M_{\mathrm{}}`$ and $`0.3M_{\mathrm{}}`$ in the solar neighborhood is $`2.1\times 10^9`$ stars yr<sup>-1</sup> pc<sup>-2</sup>. Thus, within a radius of 75 pc from the Sun, there should be approximately 370 stars with ages $`10`$ Myr. If these stars were distributed uniformly across the sky within a band of $`\pm 45^{}`$ from the galactic equator (i.e., an effective scale height of approximately 70 pc), the surface density of such objects would then be $`1.3\times 10^2`$ stars per square degree. The main body of the TW Hya Association identified by Webb et al. (1999) spans a range of 13 degress in declination by about 19 degrees in right ascension. Thus in these 245 square degrees one should expect on average 3.1 stars in the 1-10 Myr age range, whereas Webb et al. identify 11 such objects. If one includes HR 4796 and TWA 10, as suggested by Webb et al., the total number of observed objects increases to 14 but the number of predicted randomly-distributed objects in the larger 17 degree by 28 degree region also goes up to 6. Thus, while the above calculations suggest that the TW Hya Association is probably a significant enhancement in the local density of young stars, the possibility that this is a chance alignment cannot be completely ruled out. The predicted average density depends upon parameter choices that are not certain, such as the precise volume that is being sampled and the actual ages of the stars (changes of a factor of two in these properties strongly affect the apparent significance of the grouping). While it is probably not appropriate to use an averaged birthrate for such young stars, note that for typical space velocities of $`5\mathrm{km}\mathrm{s}^1`$ (Hoff et al. 1998), groups of age 10-20 Myr can overlap if they originated 50-100 pc apart. Regardless of the physical association of these objects (which really only is used to support the application of the Hipparcos distances of the main objects to all suggested members), the lack of Li depletion indicates that these stars cannot be much older than $``$10 Myr. Whatever molecular cloud(s) these stars formed in, the absence of relatively nearby molecular gas is most easily explained if the natal clouds disperse in $`10`$ Myr (Hoff et al. 1998), consistent with the general absence of 3-10 Myr old stars in molecular clouds such as Taurus-Auriga (e.g., Briceño et al. 1997), rather than by requiring very high space velocities to move stars from present-day clouds (Soderblom et al. 1998). ### 4.2 Implications for disk evolution timescales Our mid-infrared observations show that many of the stars in the TW Hya Association have little or no disk emission at 10$`\mu `$m. Even among the five stellar systems with 10$`\mu `$m excesses, most show some evidence of inner disk evolution. The disk around the A0 star HR 4796A has an $`r60`$ AU central hole in mid-infrared images (Jayawardhana et al. 1998; Koerner et al. 1998). The SEDs of HD 98800 and Hen 3-600A also suggest possible inner disk holes (Jayawardhana et al. 1999). The excess we report here for CD -337795 is modest, and could well be due to a faint companion. Only TW Hya appears to harbor an optically thick, actively accreting disk of the kind observed in $``$1-Myr-old classical T Tauri stars; it is the only one with a large H$`\alpha `$ equivalent width (-220 Å). It would be of great interest to further constrain TW Hya’s SED with flux measurements at wavelengths between 2-10 $`\mu `$m. If most TWA stars are $``$10 Myr old, the above results suggest that their inner disks have already depleted either through coagulation of dust or accretion on to the central star. The fact that only one (TW Hya) out of 16 entries in Table 2 shows classical T Tauri characteristics –compared to $``$50% of $``$1-Myr-old stars in star-forming regions– argues for rapid evolution of inner disks in pre-main-sequence stars. Observations at far-infrared and sub-millimeter wavelengths may reveal whether most TWA stars still retain their outer disks. These stars are also ideal targets for sensitive brown dwarf and planet searches with the Space Interferometry Mission (SIM) and the proposed Terrestrial Planet Finder (TPF). We wish to thank the staff of CTIO and Keck Observatory for their outstanding support. The research at CfA was supported by NASA grant NAG5-4282 and the Smithsonian Institution. The research at the University of Florida was supported by NASA, NSF, and the University of Florida. Figure Caption Figure 1. Composite spectral energy distribution (SED) of (a) TW Hydrae and (b) CD -337795. The solid line in each plot is the median SED for Taurus CTTS, normalized at H (from D’Alessio et al. 1999), and the dashed lines show the quartile fluxes to provide some idea of the range of observed CTTS fluxes.
no-problem/9906/quant-ph9906037.html
ar5iv
text
# Quantum Measured Information ## Abstract A framework for a quantum information theory is introduced that is based on the measure of quantum information associated with probability distribution predicted by quantum measuring of state. The entanglement between states of measured system and ”pointer” states of measuring apparatus, which is generated by dynamical process of quantum measurement, plays a dominant role in expressing quantum characteristics of information theory. The quantum mutual information of transmission and reception of quantum states along a noisy quantum channel is given by the change of quantum measured information. In our approach, it is not necessary to purify the transmitted state by means of the reference system. It is also clarified that there exist relations between the approach given in this letter and those given by other authors. Quantum information theory is a new field with potential implication for the conceptual foundations of quantum mechanics. It appears to be the base for a proper understanding of the emerging fields of quantum computation , quantum communication, and quantum cryptography. Recently, a correlated state in quantum systems, so-called quantum entangled or quantum entanglement, is utilized to study quantum information , in particular, quantum teleportation. A theory of quantum information has emerged which shows striking parallels with, but also fascinating differences from, classical information theory entirely based on the von Neumann entropy of quantum state. Although some useful fundamental results about quantum information theory, e.g., quantum noiseless coding theorem and the capacity of quantum noisy channels, have been obtained recently, quantum information is still mystery in many respects. To our knowledge, there exist several approaches for quantum information theory . However, the relation among these approaches is obscure. We argue in this letter that it is necessary to introduce quantum measured information which is regarded as the measure of information for quantum input and output. This leads us to propose a scheme for quantum information theory. Our approach can give rise to a unified description of classical correlation and quantum entanglement. Furthermore, we shall clarify the relations between our approach and those given by other authors for quantum information theory. In classical information theory, there is a set of mutually exclusive classical states. In quantum mechanics, a quantum state is represented by a vector in a Hilbert space, or a density operator on that space. Classically, the input system may retain its original state, while the no-cloning theorem implies that in the quantum case the input system cannot in general remain in its initial state. However, in many quantum applications, one is interested not only in transmitting a discrete set of states, but also in arbitrary superpositions of those states. That is, one wants to transmit entire subspace of states. It is well known that an arbitrary state can be represented as a mixture of pure states, i.e., by imposing classical randomness on pure states. In this sense pure states are ”noiseless”, i.e., they contain no classical source of randomness. In our point of view, this does not imply that the pure states contain no quantum source of randomness. Such randomness comes from probability distribution predicted by quantum measuring about quantum states. We now consider the question of measurement in quantum mechanics. From our point of view there is no fundamental distinction between measuring apparatus and other physical systems. Therefore, a measurement is simply a special case of interaction between physical systems, which has the property of correlating a quantity in one system with a quantity in the other. Nearly every interaction between systems produces some correlation however. Suppose that at some instant a pair of systems are independent, so that the composite system state function is a product of subsystem states. Then this condition obviously holds only instantaneously, since the systems are interacting, the independence will be immediately destroyed and the systems will become correlated. There is still one more requirement that we must impose on an interaction before we shall call it a measurement. If the interaction is to produce a measurement of mechanical quantity $`A`$ of subsystem $`S_1`$ by the quantity $`P`$ of another one $`S_2`$, we require that such interaction shall never decrease the information in the reduced distribution about $`A`$. Furthermore, we also expect that a knowledge of $`P`$ shall give us more information about $`A`$ than we had before the measurement took place, since otherwise the measurement would be useless. The restriction that the interaction shall not decrease the information of the reduced system $`S_1`$ has the interacting consequence that the eigenstates of $`A`$ will not be disturbed, since otherwise the information of $`A`$ would be decreased. The time evolution of the composed system made of the measured system and the apparatus system should exhibit the following properties. The initial states of systems $`S_1`$ and $`S_2`$ are the form of superposition $`_ia_i|\varphi _i>`$ and $`|P(0)>`$, respectively. Then, after a specified time of interaction the total state $`_ia_i|\varphi _i>|P(0)>`$ will be transformed into a a form of the superposition states $`_ia_i|\varphi _i>|P_i>`$, i.e., the initial independent state is evolved into entanglement state. Information concepts have been used in the context of quantum measurements long ago, and various quantities, all labeled entropies, have been introduced to characterize uncertainties about events or about states of a system. For a completed orthonormal set $`\{|\varphi _i>\}`$ and a pure state $`|\psi >=_ia_i|\varphi _i>`$, we have a square-amplitude distribution $`|a_i|^2`$ called the distribution of $`|\psi >`$ over $`\{|\varphi _i>\}`$. In the probabilistic interpretation this distribution represents the probability distribution over the results of a measurement with eigenstates $`\{|\varphi _i>\}`$ performed upon the measured system in the state $`|\psi >`$. An entropy of information depends not only on the actual set of probabilities for the considered events, but also on the measure associated with the a priori probabilities predicted. Such entropy of information is given by Shannon entropy $`S=_i|a_i|^2\mathrm{log}|a_i|^2`$. Considering the dynamical process of quantum measurement as mentioned above, we can arrange the distribution corresponding to the reduced density matrix of measured system into the a priori probabilistic distribution predicted by quantum measurement. While for the composite system made of the measured system and the apparatus system, its state is evolved into such a state $`|\psi >=_ia_i|\varphi _i>|P_i>`$. The reduced density matrix can be obtained by tracing the state out the degrees of freedom of the apparatus system, i.e., $`\rho _M=Tr_P(|\psi ><\psi |)=_i|a_i|^2|\varphi _i><\varphi _i|`$. The quantum measured information carried by the quantum state $`|\psi >`$ can be read as $$S_M(\rho )=S(\rho _M)=Tr\rho _M\mathrm{log}\rho _M.$$ (1) In order to investigate the quantum mutual information, we shall consider the mathematical description of model about a noisy quantum channel following Schumacher. Suppose a quantum system $`S`$ is subjected to a dynamical evolution, which may represent the transmission of $`S`$ along a noisy quantum channel. In general, the evolution of $`S`$ will be represented by a superoperator $``$ which gives the mapping from the initial states of the system $`\rho `$ to the final states after the evolution of the system $`\rho ^{}`$, i.e., $`\rho ^{}=(\rho )`$. The mapping represented by $``$ is a linear, trace-nonincreasing and completely positive map. The evolution of system will be unitary only if it is isolated from other systems. The input quantum state, after interaction with an environment, is lost, having become the output state. Any attempt at copying the quantum state before decoherence will result in a classical channel. Thus a joint probability for input and output symbols does not exist for quantum channels. However, this is not essential, as the quantity of interest is the quantum measured information associated with the probabilistic distribution predicted by quantum measurement, which is regarded as a measure of information carried by a quantum state. Let us recall the definition of the mutual information in classical information theory. The mutual information is a measure of an amount of information that one random variable contains about another random variable, and is the reduction in the uncertainty of one random variable due to the knowledge of the other. In the quantum case, the corresponding quantity is the entropy of information given by that the change of quantum measured distribution, which is resulted in by the evolution along a noisy quantum channel. After the interaction with the environment, the quantum measured distribution $`\rho _M`$ of quantum state becomes $`\rho _M^{}=(\rho _M)=_i|a_i|^2(|\varphi _i><\varphi _i|)`$. In fact, The $`\rho _M^{}`$ can be equivalently expressed as $`\rho _M^{}=Tr_P\mathrm{𝟏}_P(|\psi ><\psi |)`$. It should be noticed that the tracing process implies the determination of probabilistic distribution predicted by quantum measurement. In general, the state $`\rho _E^{}=\mathrm{𝟏}_P(|\psi ><\psi |)`$ is a quantum mixing state, which carries the quantum information given by the von Neumann entropy, i.e., $`S(\rho _E^{})=Tr\rho _E^{}\mathrm{log}\rho _E^{}`$. In physics, it is known that the entropy change in the quantum states represents the amount of the information obtained by the quantum measurement. The amount of information of gain is equal to subtracting the information $`S(\rho _E^{})`$ before quantum measurement from the quantum measured information $`S(\rho _M^{})`$. This leads to the quantum mutual information represented by $$I_{}=S(\rho _M^{})S(\rho _E^{}),$$ (2) based on the quantum measured information. Up to now, we have discussed how the quantum mutual information is measured when one transmits a pure quantum state along a noisy quantum channel by means of concept of quantum measured information. It is well known that in general case we should study the problem of transmission of some quantum mixed states because the states of a quantum system are fragile. Based on the following propositions , we can easily generalize the above idea to the case of quantum mixed states. The first proposition is that if $`\rho `$ is a pure state, there exists a composite system made of two subsystems of which $`\rho `$ is the state and the von Neumann entropies of the subsystems satisfy $`S(\rho _1)=S(\rho _2)`$, where $`\rho _1`$ and $`\rho _2`$ are the density matrixes of the subsystems. Moreover, the positive spectra of $`\rho _1`$ and $`\rho _2`$ coincide. Secondly, given $`\rho _1`$, one can always find a Hilbert space $`_2`$ and a pure density matrix $`\rho `$ in the Hilbert space $`_1_2`$ such that $`\rho _1=Tr_2\rho `$. Now, let us suppose that a transmitted state in the quantum mixed state $`\rho _1=_mp_m|S_m^1><S_m^1|`$. We can take an auxiliary Hilbert space $`_2`$ of which the dimensions are the same as those of $`_1`$. Thus, a pure state of the composite system can be constructed as $`\rho =|\chi ><\chi |`$, here $`|\chi >=_m\sqrt{p_m}|S_m^1>|S_m^2>`$. If one plans to measure the quantum mechanical quantity $`A_1`$ of the subsystem being in the state $`\rho _1`$, he should expand the state $`|\chi >`$ with the quantity $`A_1`$ corresponding to the complete and orthogonal eigenstates $`\{|\varphi _i^1>\}`$ $$|\chi >=\underset{m,i}{}\sqrt{p_m}c_{mi}|\varphi _i^1>|\varphi _i^2>=\underset{i}{}\stackrel{~}{c}_i|\varphi _i^1>|\varphi _i^2>.$$ (3) The dynamical evolution of quantum measurement is that of combining the measured system with the apparatus system of measuring, which leads to the state $`|\chi >|P(0)>`$ into $`|\stackrel{~}{\psi }>=_i\stackrel{~}{c}_i|\varphi _i^1>|\varphi _i^2>|P_i>`$. So the quantum measured information of the quantum state $`\rho _1`$ is read as $$\begin{array}{c}S(\stackrel{~}{\rho }_M)=_i|\stackrel{~}{c}_i|^2\mathrm{log}|\stackrel{~}{c}_i|^2=Tr\stackrel{~}{\rho }_M\mathrm{log}\stackrel{~}{\rho }_M\hfill \\ =Tr[Tr_P(|\stackrel{~}{\psi }><\stackrel{~}{\psi }|)\mathrm{log}Tr_P(|\stackrel{~}{\psi }><\stackrel{~}{\psi }|)],\hfill \end{array}$$ (4) where $`\stackrel{~}{\rho }_M=Tr_{P,2}(|\stackrel{~}{\psi }><\stackrel{~}{\psi }|)=Tr_{\stackrel{~}{P}}_{i,j}\stackrel{~}{c}_i\stackrel{~}{c}_j^{}|\varphi _i^1>|\stackrel{~}{P}_i><\stackrel{~}{P}_j|<\varphi _j^1|`$. The set of states $`\{|\stackrel{~}{P}_i>=|\varphi _i^2>|P_i>\}`$ exhibits that the states of the apparatus system are completely entangled with those of the auxiliary subsystem $`S_2`$. Hence, we can regard the states $`\stackrel{~}{P}_i`$ as the ”pointer” basis of quantum measurement. This implies that the subsystem $`S_2`$ is completely auxiliary, and can be absent in our formalism of quantum information theory. Consequently, it is not necessary in our approach of quantum information theory based on the quantum measured information to purify the initial transmitted state in the Schumacher’s approach. By means of the previous discussion about the quantum mutual information of transmission of quantum pure state and the corresponding reception, we can immediately write the expression of quantum mutual information of transmission of quantum mixed states in the noisy quantum channel mentioned above. The result is $$\stackrel{~}{I}_{}=S(\stackrel{~}{\rho }_M^{})S(\stackrel{~}{\rho }_E^{}).$$ (5) $`\stackrel{~}{\rho }_E`$ denotes the dynamically evolved state of the composite system made of the measured system and the apparatus system, i.e., $`\stackrel{~}{\rho }_E=|\stackrel{~}{\psi }><\stackrel{~}{\psi }|`$. Through the noisy quantum channel, the states $`\stackrel{~}{\rho }_E`$ and $`\stackrel{~}{\rho }_M`$ become $`\stackrel{~}{\rho }_E^{}=\mathrm{𝟏}_{\stackrel{~}{P}}(\stackrel{~}{\rho }_E)`$ and $`\stackrel{~}{\rho }_M^{}=Tr_{\stackrel{~}{P}}\stackrel{~}{\rho }_E^{}`$, respectively. It should be emphasized again that the $`\stackrel{~}{\rho }_M^{}`$ stands for the change of the probabilistic distribution predicted by quantum measurement after the transmitted state going through the noisy quantum channel. If a quantum channel is trivial, i.e., $`=`$ identity map, then the quantum mutual information equals to the quantum measured information of inputs. This is easily seen from the relation that $`\stackrel{~}{I}_{}_{=\mathrm{𝟏}}=S(\stackrel{~}{\rho }_M^{})_{=\mathrm{𝟏}}=S(\stackrel{~}{\rho }_M)`$. According to the Araki-Lieb triangle inequality about the entropy of information, we can lead to the quantum mutual information presented here satisfying $`\stackrel{~}{I}_{}S(\stackrel{~}{\rho }_M)`$ which is the first part of the data processing inequality. Now, we shall consider a more complicated quantum channel. Suppose the initial state of the measured system $`S`$ is $`\rho _S`$ and further suppose $`S`$ undergoes two successive dynamical evolutions described by superoperators $`_1`$ and $`_2`$. This scheme of the noisy quantum channel can be represented by the evolutions of quantum state $`\rho _S_1(\rho _S)_2_1(\rho _S)`$. Following Schumacher and Neilson and applying the strong subadditivity inequality, we can prove the second part of the data processing inequality $`\stackrel{~}{I}__1\stackrel{~}{I}_{_2_1}`$. On the other hand, by writing the superoperator $``$ as a unitary evolution $`U_{SE}`$ on an extended system $`SE`$ followed by a partial trace over an environment system $`E`$, we can investigate the reverse data processing inequality in the quantum information theory, which reflects the fact that any quantum channel used in a forward manner can be used in a backward manner. If we consider the case of transmission of quantum states in a backward manner and exchange the input state with the output state, the order of time in the unitary evolution $`U_{SE}`$ is reverse, which leads to the evolution operator to be changed into $`U_{SE}^{}`$. So, in the transmission of backward manner, we should substitute the evolution of state $`^{}(\rho _S)`$ for the $`(\rho _𝒮)`$. Since the ”pointer” state of apparatus system is completely entangled with the input states, the exchange between the input states and the output states is equivalent to the change of $`|\varphi _i^1>|\stackrel{~}{P}_i>`$ by $`|\stackrel{~}{P}_i>|\varphi _i^1>`$ in the states $`\stackrel{~}{\rho }_M^{}`$ and $`\stackrel{~}{\rho }_E^{}`$. Noticing that the reverse of the noisy quantum channel leads $``$ to $`^{}`$, so we should use the superoperator $`\mathrm{𝟏}^{}`$ as a substitute for $`\mathrm{𝟏}`$ in the expression of the quantum mutual information in the backward manner. Based on the fact that the spectra of general quantum states $`\rho `$ and $`\rho ^{}`$ coincide, we can obviously see that the quantum mutual information Eq.(5) is invariant under the transformations of $`|\varphi _i^1>|\stackrel{~}{P}_i>`$ into $`|\stackrel{~}{P}_i>|\varphi _i^1>`$, and $`\mathrm{𝟏}`$ into $`\mathrm{𝟏}^{}`$. From these, we obtain the reverse data processing inequality about the quantum mutual information, $`\stackrel{~}{I}_{_2_1}=\stackrel{~}{I}_{_1_2}\stackrel{~}{I}__2S(\stackrel{~}{\rho }_M)`$. The appearance of quantum characteristics in a quantum state is related to quantum non-separability. In fact, in general, the states of apparatus system and the states of measured system are not separable. If the general quantum state constructed in terms of these states is restricted within the separable case, i.e., $`\rho _c=_iw_i\rho _i^1\rho _i^{\stackrel{~}{P}}`$ where $`\rho _i^1`$ is from the Schatten decomposition of the state of the measured system $`\rho _1=_iw_i\rho _i^1`$, we find that the quantum mutual information Eq.(5) is reduced to $`\stackrel{~}{I}_{}=S(Tr_{\stackrel{~}{P}}\mathrm{𝟏}(\rho _c))[S(\mathrm{𝟏}(\rho _c))S(\rho _c)]=S(Tr_{\stackrel{~}{P}}\mathrm{𝟏}(\rho _c))_iw_iS((\rho _i^1))`$. The change of the unmeasured information is present in the square bracket of the above expression because the separable state $`\rho _c`$ is a quantum mixed state. The input state can be equivalently described by the ”pointer” states of input exploring the property of quantum measurement. Then, for the case of separability, our definition of quantum mutual information based on the quantum measured information deduces to that of Ohya’s approach which is established by means of the compound state describing the correlation between an input state $`\rho _1`$ and the output state $`(\rho _1)`$. However, we should emphasize that Ohya’s approach is not complete because the quantum coherence of the entanglement states, which plays the important roles in quantum computation, quantum teleportation and quantum cryptography, is restricted out in his approach. In our scheme, the quantum mutual information reduces to the classical one when the system is classical. When the input system is classical, an input state is given by a probability distribution or a probability measure. For the case of probability distribution, the input state can be expressed by $`\rho =_iw_i\delta _i`$, where $`\delta `$ is the delta measure, i.e., $`\delta _i(j)=\delta _{i,j}`$. As the special case of the separability, the mutual information for the channel $``$ becomes $`I_{,c}=S((\rho ))_iw_iS((\delta _i))`$, which has been taken as the definition of the mutual information for a classical-quantum-classical channel . Since the measure of quantum information is the quantum measured information in our formalism, we can communicate the same amount of quantum informations by using the transmissions of the pure state $`\rho =_{i,j}a_ia_j^{}|\varphi _i><\varphi _j|`$ or the mixed state $`\stackrel{~}{\rho }=_i|a_i|^2|\varphi _i><\varphi _i|`$ in a Hilbert space. However, in the process of transmissions, the fidelities for the two type of quantum states are different. The quantum states $`\rho `$ and $`\stackrel{~}{\rho }`$ of transmissions along the noisy quantum channel $``$ are evolved into $`\rho ^{}=(\rho )`$ and $`\stackrel{~}{\rho }^{}=(\stackrel{~}{\rho })`$, respectively. For the case of transmission of the pure state, its fidelity is read as $`F(\rho ,\rho ^{})=Tr(\rho \rho ^{})`$. But, the fidelity for the mixed quantum states, which is defined in terms of Uhlmann’s formula of transition probability, is given by $`F(\stackrel{~}{\rho },\stackrel{~}{\rho }^{})=\{Tr[(\sqrt{\stackrel{~}{\rho }}\stackrel{~}{\rho }^{}\sqrt{\stackrel{~}{\rho }})^{\frac{1}{2}}]\}^2`$. It is necessary to analyze the quantitative properties of these fidelities although we do not extensively discuss this topic here. It is hoped that such analyzing results may be applicable to developing a quantum analogue of Shannon’s channel capacity theorem. Summarizing, we have introduced the quantum measured information to measure the quantum information of a quantum state. Using this point of view we have consistently decided the quantum mutual information, which measures the amount of quantum information conveyed in the noisy quantum channel. We has proven that such quantum mutual information obeys the data processing inequality in both forward manner and backward manner. I thank Z.C. Lu for helpful discussions. The work is supported by the NNSF of China (Grant No.19875041), the Special NSF of Zhejiang Province (Grant No.RC98022) and Cao Guang-Biao Foundation in Zhejiang University.
no-problem/9906/nucl-th9906096.html
ar5iv
text
# Evidence for a missing nucleon resonance in kaon photoproduction ## Abstract New SAPHIR $`p(\gamma ,K^+)\mathrm{\Lambda }`$ total cross section data show a resonance structure at a total c.m. energy around 1900 MeV. We investigate this feature with an isobar model and find that the structure can be well explained by including a new $`D_{13}`$ resonance at 1895 MeV. Such a state has been predicted by a relativistic quark model at 1960 MeV with significant $`\gamma N`$ and $`K\mathrm{\Lambda }`$ branching ratios. We demonstrate how the measurement of the photon asymmetry can be used to further study this resonance. In addition, verification of the predicted large decay widths into the $`\eta N`$ and $`\eta ^{}N`$ channels would allow distinguishing between other nearby $`D_{13}`$ states. PACS number(s): 14.20.Gk, 25.20.Lj, 13.60.Le, 13.30.Eg The physics of nucleon resonance excitation continues to provide a major challenge to hadronic physics due to the nonperturbative nature of QCD at these energies. While methods like Chiral Perturbation Theory are not amenable to $`N^{}`$ physics, lattice QCD has only recently begun to contribute to this field. In a recent study the excitation energies of $`1/2^{}`$ and $`3/2^{}`$ baryon resonances are calculated for the first time on the lattice with improved actions. The results show a clear splitting of these states from the ground state nucleon, demonstrating the potential and the promise of extracting $`N^{}`$ structure from lattice QCD. However, most of the theoretical work on the nucleon excitation spectrum has been performed in the realm of quark models. Models that contain three constituent valence quarks predict a much richer resonance spectrum than has been observed in $`\pi N\pi N`$ scattering experiments. Quark model studies have suggested that those ”missing” resonances may couple strongly to other channels, such as the $`K\mathrm{\Lambda }`$ and $`K\mathrm{\Sigma }`$ channels or final states involving vector mesons. The newly established electron and photon facilities have made it possible to investigate the mechanism of nucleon resonance excitation with photons with much improved experimental accuracy. Experiments with kaon-hyperon final states have been performed at ELSA and are being analyzed at JLab. Much improved data are becoming available in the $`p(\gamma ,K^+)\mathrm{\Lambda }`$, $`p(\gamma ,K^+)\mathrm{\Sigma }^0`$ and $`p(\gamma ,K^0)\mathrm{\Sigma }^+`$ channels, from total cross section to polarization observables. The new SAPHIR total cross section data for the $`p(\gamma ,K^+)\mathrm{\Lambda }`$ channel, shown in Fig. 1, indicate for the first time a structure around $`W=1900`$ MeV. This structure could not be resolved before, due to the low quality of the old data. It is the purpose of this paper to investigate this structure in the framework of an isobar model. Pioneered by Thom , most studies over the last 30 years analyzed the $`N(\gamma ,K)\mathrm{\Lambda }(\mathrm{\Sigma })`$ in a tree-level isobar framework that included a number of resonances whose couplings were adjusted to reproduce the experimental data. Due to the poor data quality it was not possible to decide which resonances contributed, even the magnitude of the background terms was uncertain. Recently, two new developments have provided significant progress in this field. First, a coupled-channels calculation that included final-state interactions linked the photoproduction process $`p(\gamma ,K^+)\mathrm{\Lambda }`$ to the hadronic process $`p(\pi ^{},K^0)\mathrm{\Lambda }`$. Secondly, the recent work on including hadronic form factors in photoproduction reactions while maintaining gauge invariance has resulted in the proper description of the background terms, allowing the use of approximate SU(3) symmetry to fix the Born coupling constants $`g_{K\mathrm{\Lambda }N}`$ and $`g_{K\mathrm{\Sigma }N}`$. Due to their isospin structure the $`K\mathrm{\Sigma }`$ photoproduction channels can involve the excitation of $`N^{}`$ as well as $`\mathrm{\Delta }`$ states. On the other hand, $`K\mathrm{\Lambda }`$ photoproduction only involves intermediate isospin 1/2 resonances and is therefore easier to describe. Here, we use the tree-level isobar model described in Ref. to analyze the $`p(\gamma ,K^+)\mathrm{\Lambda }`$ process in more detail. Guided by a recent coupled-channels analysis , the low-energy resonance part of this model includes three states that have been found to have significant decay widths into the $`K^+\mathrm{\Lambda }`$ channel, the $`S_{11}`$(1650), $`P_{11}`$(1710), and $`P_{13}(1720)`$ resonances. In order to approximately account for unitarity corrections at tree-level we include energy-dependent widths along with partial branching fractions in the resonance propagators . The background part includes the standard Born terms along with the $`K^{}`$(892) and $`K_1`$(1270) vector meson poles in the $`t`$-channel. As in Ref. , we employ the gauge method of Haberzettl to include hadronic form factors. The fit to the data was significantly improved by allowing for separate cut-offs for the background and resonant sector. For the former, the fits produce a soft value around 800 MeV, leading to a strong suppression of the background terms while the resonant cut-off is determined to be 1900 MeV. As shown in Fig. 1, our model cannot reproduce the SAPHIR total cross section data without inclusion of a new resonance with a mass of around 1900 MeV. While there are no 3- or 4-star isospin 1/2 resonances around 1900 MeV in the Particle Data Table , several 2-star states are listed. Of those only the $`D_{13}(2080)`$ state has been identified in older $`p(\pi ^{},K^0)\mathrm{\Lambda }`$ analyses to have a noticeable branching ratio into the $`K\mathrm{\Lambda }`$ channel. On the theoretical side, the constituent quark model by Capstick and Roberts predicts many new states around 1900 MeV; however, only a few of them have been calculated to have a significant $`K\mathrm{\Lambda }`$ decay width . These are the $`[S_{11}]_3`$(1945), $`[P_{11}]_5`$(1975), $`[P_{13}]_4`$(1950), and $`[D_{13}]_3`$(1960) states, where the subscript refers to the particular band that the state is predicted in. We have performed fits for each of these possible states, allowing the fit to determine the mass, width and coupling constants of the resonance. We found that all four states can reproduce the structure at $`W`$ around 1900 MeV, reducing the $`\chi ^2/N`$ from around 4.5 to around 3 in each case. Table I compares our extracted resonance parameters with the quark model predictions of Ref. . While all four of the above resonances have large decay widths into the $`K\mathrm{\Lambda }`$ channel, only the $`D_{13}`$(1960) state is predicted to also have significant photocouplings. Table I presents the remarkable agreement, up to the sign, between the quark model prediction and our extracted results for the $`D_{13}`$(1960). The sign remains ambiguous, since at this stage we only extract the product of coupling constants. For the other three states the partial widths extracted from our fit overestimate the quark model results by up to a factor of 30. How reliable are the quark model predictions? Clearly, one test is to confront its predictions with the extracted couplings for the well-established resonances in the low-energy regime of the $`p(\gamma ,K^+)\mathrm{\Lambda }`$ reaction, the $`S_{11}(1650)`$, $`P_{11}(1710)`$ and $`P_{13}(1720)`$ excitations. Table II shows that the magnitudes of the extracted partial widths for the $`S_{11}(1650)`$, $`P_{11}(1710)`$, and $`P_{13}(1720)`$ are in good agreement with the quark model. Therefore, even though the remarkable quantitative agreement in the case of the $`D_{13}`$(1960) is probably fortuitous, we believe the structure in the SAPHIR data is in all likelihood produced by this particular resonance. Is this state identical to the 2-star resonance $`D_{13}`$(2080) listed in the Particle Data Table? Table III displays a list of $`D_{13}`$ states below 2.2 GeV predicted by Refs. , along with the Particle Data Table listings. A closer examination of the literature reveals that there is some evidence for two resonances in this wave between 1800 and 2200 MeV ; one with a mass centered around 1900 MeV and another with mass around 2080 MeV. It is the former which has been seen prominently in two separate $`p(\pi ^{},K^0)\mathrm{\Lambda }`$ analyses . Thus, we believe that the state appearing in the SAPHIR data is in fact identical to the one seen in hadronic $`K\mathrm{\Lambda }`$ production and corresponds to the $`D_{13}`$(1960) state predicted by the quark model. The $`D_{13}`$ excitation around 2080 MeV seen in Refs. may well correspond to the quark model state $`D_{13}`$(2055) in the $`N=4`$ band. In order to clearly separate these nearby $`D_{13}`$ states, measuring other channels will be helpful. For example, Ref. predicts the $`D_{13}(1960)`$ to have large decay widths into the $`\eta N`$ and $`\eta ^{}N`$ channels, in contrast to the $`D_{13}(2055)`$ whose branching ratios into these channels are negligible. Figure 1 compares our models with and without the $`D_{13}`$(1960) with the SAPHIR total cross section data. Our result without this resonance shows only one peak near threshold, while inclusion of the new resonance leads to a second peak at $`W`$ slightly below 1900 MeV, in accordance with the new SAPHIR data. The difference between the two calculations is much smaller for the differential cross sections, as displayed in Fig. 2. As expected, including the $`D_{13}`$(1960) does not affect the threshold and low-energy regime while it does improve the agreement at higher energies. Figure 3 compares the recoil polarization for the two calculations. Clearly, the differences are small for all angles, demonstrating that the recoil polarization is not the appropriate observable to further study this resonance. The target asymmetry of $`K^+\mathrm{\Lambda }`$ photoproduction is shown in Fig. 4. Here we find larger variations between the two calculations, especially for higher energies. The three data points seem to favor a model without the new $`D_{13}(1960)`$; however, more complete and accurate measurements are clearly needed over the whole angular range before any conclusion can be drawn. The largest effects are found in the photon asymmetry shown in Fig. 5. For $`W1800`$ MeV, including the new resonance leads to a sign change in the photon asymmetry whose magnitude is almost one at intermediate angles. Therefore, we would suggest that measuring this observable is well suited to shed more light on the contribution of this state in kaon photoproduction. In conclusion, we have investigated the structure around $`W=1900`$ MeV in the new SAPHIR total cross section data in the framework of an isobar model. We found that the data can be well reproduced by including a new $`D_{13}`$ resonance with a mass, width and coupling parameters in good agreement with the values predicted by a recent quark model calculation. To further elucidate the role and nature of this state we suggest measurements of the polarized photon asymmetry around $`W=1900`$ MeV for the $`p(\gamma ,K^+)\mathrm{\Lambda }`$ reaction. With the arrival of new, high-precision cross section and polarization data the kaon photoproduction process will be able to unfold its full potential in the search and study of nucleon resonances. TM thanks the member of the Center for Nuclear Studies for the hospitality extended to him during his stay in Washington, D.C. This work was supported by the University Research for Graduate Education (URGE) grant (TM), and US DOE grant DE-FG02-95ER-40907 (CB).
no-problem/9906/astro-ph9906069.html
ar5iv
text
# 1 Young stellar kinematic groups Late-type stars members of young stellar kinematic groups D. Montes<sup>1</sup>, A. Latorre<sup>1</sup>, M.J. Fernández-Figueroa<sup>1</sup> <sup>1</sup> Departamento de Astrofísica, Facultad de Físicas, Universidad Complutense de Madrid, E-28040 Madrid, Spain (dmg@astrax.fis.ucm.es) To be published in ASP Conf. Ser., Stellar clusters and associations: convection, rotation, and dynamos (Second ”Three-Islands” Euroconference) (May 25 - 28, 1999, Mondello, Palermo, Sicily, Italy), R. Pallavicini, G. Micela and S. Sciortino eds. Abstract We have compiled a catalog of late-type stars (F5-M) member of representative young disk stellar kinematic groups: the Local Association (Pleiades moving group, 20 - 150 Myr), Ursa Mayor group (Sirius supercluster, 300 Myr), and Hyades supercluster (600 Myr). Other moving groups as IC 2391 supercluster (35 Myr) and Castor Moving Group (200 Myr) have been also included. Stars have been selected from previously established member of stellar kinematic groups based in photometric and kinematic properties as well as from candidates based in other criteria as their level of chromospheric activity, rotation rate, lithium abundance. Precise measurements of proper motions and parallaxes taken from Hipparcos Catalogue, and published radial velocity measurements are used to calculate Galactic space motions (U, V, W) in order to determine the membership of the selected stars to the different stellar kinematic groups. In addition to kinematic properties we also give for each star photometric, spectroscopic and physical properties as well as information about activity indicators and Li abundance. Some chromospherically active binaries results to be also members of some of these stellar kinematic groups. Introduction Stellar kinematic groups, moving groups or superclusters, are kinematically coherent groups of stars that should have the same basic properties as a genuine open cluster (except for spatial compactness). The origin of these groups can be the evaporation of a open cluster of the remnants of a star formation region. It has long been known that in the solar vicinity there are several kinematic groups of stars that share the same space motions that well know open clusters. The best documented groups are the Hyades supercluster (Eggen 1992b) associated with the Hyades cluster (600 Myr), the Ursa Mayor group (Sirius supercluster) (Eggen 1984a, 1992a, 1998b, Soderblom & Mayor 1993a, b) associated with the UMa cluster of stars (300 Myr). A younger kinematic group called the Local Association or Pleiades moving group seems to consists of a reasonably coherent kinematic stream of young stars with embedded clusters and associations such as the Pleiades, $`\alpha `$ Per, NGC 2516, IC 2602 and the Scorpious-Centaurus cluster (Eggen 1983, 1992c). The age of the star of this association range from about 20 to 150 Myr. Evidences have been found that X-ray and EUV selected active stars and lithium-rich stars (Favata et al. 1993, 1995, 1998; Jeffries & Jewell 1993; Mullis & Bopp 1994; Jeffries 1995) are member of this association. Other two young moving groups are the IC 2391 supercluster (35-55 Myr) (Eggen 1991, 1995) and the Castor Moving Group (200 Myr) (Barrado y Navacués 1998). Recently, several studies using extended samples of star with known radial velocities and astrometric data taken from Hipparcos (Chereul et al. 1998, 1999, Dehnen 1998, Asiain et al. 1999) not only confirm the existence of classical moving groups, but also detect finer structures that in several cases can be related to kinematic properties of nearby open clusters or associations. More complex structures characterized by several longer branches (Sirius, middle, and Pleiades branches) running almost parallel to each other across the UV-plane have been found by Skuljan et al. (1999) in their study of the velocity distribution of star in the solar neighborhood. A large fraction of the well known members to the different moving groups are early type star, however few studies are centered in late-type stars. Identification of significant numbers of the late-type population of these young moving groups would be extremely important to the study of the chromospheric and coronal activity and their age evolution. In this contribution we compile a sample of late-type stars, previously established members or possible new candidates to different young stellar kinematic groups (see Table 1). We examine their kinematic properties (distribution of stars in the velocity space, UV amd WV planes), using the more recent radial velocities and astrometric data available, in order to determine their membership to the different moving groups. Selection of the Sample Stars included in this work have been selected from previously established member of stellar kinematic groups (see references given in Table 1) based in photometric and kinematic properties as well as from candidates based in other criteria as their level of chromospheric activity, rotation rate, lithium abundance. We have selected star from different sources: * The study of Agekyan & Orlov (1984) which searched for kinematic groups in the solar neighborhood. * The study of ages of spotted late-type stars by Chugainov (1991). * X-ray and EUV selected active stars and lithium-rich stars (Favata et al. 1993, 1995, 1998; Jeffries & Jewell 1993; Tagliaferri et al. 1994, Mullis & Bopp 1994; Jeffries 1995, Schschter et al. 1996). * Single rapidly rotating stars as AB Dor, PZ Tel, HD 197890, RE J1816+541, BD+22 4409 (LO Peg), HK Aqr, V838 Cen, V343 Nor, LQ Hya, previously assigned membership of the Local Association. * Chromospherically active late-type stars dwarfs in the solar neighborhood with studied kinematic properties (Young et al. 1987; Upgren 1988; Soderblom 1990; Ambruster et al. 1998). * Flare stars with studied kinematic properties (Poveda et al. 1996). * Other chromospherically active single and binary stars (Strassmeier et al. 1993, Henry et al. 1995, 1996; Soderblom et al. 1998) * The study of nearby young solar analogs by Gaidos (1998). Membership to the moving groups In order to determine the membership of this sample to the different stellar kinematic groups we have studied the distribution of stars in the velocity space by calculating the Galactic space-velocity components (U, V , W) in a right-handed coordinated system (positive in the directions of the Galactic center, Galactic rotation, and the North Galactic Pole, respectively). The procedures in Johnson & Soderblom (1987) were used to calculate U, V, W, and their associated errors. \- Parallaxes and proper motions are taken from Hipparcos Catalogue (ESA, 1997); PPM (Positions and Proper Motions) Catalogue (Röser & Bastian 1991; Bastian et al, 1993; Röser et al, 1994); ACT Reference Catalog (Urban et al. 1997); and TCR (Tycho Reference Catalogue) (Hog et al. 1998). \- Radial velocities are taken primarily from the compilation WEB (Wilson Evans Batten) Catalogue (Duflot et al. 1995), the Catalogue of radial velocities of Nearby Stars (Tokovinin, 1992), and from other references given in SIMBAD, and in the CNS3, Catalogue of Nearby Stars, Preliminary 3rd Version (Gliese & Jahreiss 1991). – In Fig 1. we represent the (U, V) and (W, V) planes (Boettlinger Diagram) for our star sample. The distribution of the stars in this figure shows concentrations around the (U, V, W) position corresponding to the five moving groups listed in Table 1. In base of these concentrations we have classified the stars of our sample as member of one of these moving groups or as other young disk star if their classification is not clear but it is inside or near the boundaries (dashed line in Fig. 1) that determine the young disk population as defined by Eggen (1984b, 1989). In Fig. 2 we plot the (U, V) for the Local Association, with some star identified. – In Tables 2 to 6 <sup>1</sup><sup>1</sup>1Tables 2 to 6 available at http://www.ucm.es/info/Astrof/ltyskg.html we list the candidate stars for each moving group. We give the name, spectral type, coordinates (FK5 1950.0), radial velocity (V<sub>r</sub>) and the error in km/s, parallax ($`\pi `$) and the error in milli arc second (mas), proper motions $`\mu `$<sub>α</sub> and $`\mu `$<sub>δ</sub> and their errors in mas per year (mas/yr), and the U, V, W, calculated components with their associated errors in km/s. In the last column we mark with Y previously established members of the stellar kinematic group and Y? possible new members in base of their position in the (U, V) plane.
no-problem/9906/hep-ph9906382.html
ar5iv
text
# 1 Introduction ## 1 Introduction An effective charge encodes the entire perturbative correction of a QCD observable; for example, the ratio of $`e^+e^{}\gamma ^{}\mathrm{hadrons}`$ annihilation to muon pair cross sections can be written $$R_{e^+e^{}}(s)\frac{\sigma (e^+e^{}\mathrm{hadrons})}{\sigma (e^+e^{}\mu ^+\mu ^{})}=R_{e^+e^{}}^0(s)\left(1+\frac{\alpha _R(\sqrt{s})}{\pi }\right),$$ (1) where $`R_{e^+e^{}}^0`$ is the prediction at Born level. More generally, the effective charge $`\alpha _A(Q)`$ is defined as the entire QCD radiative contribution to an observable $`𝒪_A(Q)`$ : $$𝒪_A(\mathrm{\Lambda })=𝒪_A^0\left(\delta _A+\frac{\alpha _A(\mathrm{\Lambda })}{\pi }\right),$$ (2) where $`\delta _A`$ is the zeroth order QCD prediction (i.e., the parton model), and $`\alpha _A(\mathrm{\Lambda })/\pi `$ is the entire QCD correction. Note that $`\delta _A=0`$ or $`1`$ depending on whether the observable A exists at zeroth order. Important examples with $`\delta _A=1`$ are the $`e^+e^{}`$ annihilation cross-section ratio and the $`\tau `$ lepton’s hadronic decay ratio, $$R_\tau \frac{\mathrm{\Gamma }(\tau ^{}\nu _\tau +\mathrm{hadrons})}{\mathrm{\Gamma }(\tau ^{}\nu _\tau e^{}\overline{\nu _e})}=R_\tau ^0\left(1+\frac{\alpha _\tau (m_\tau )}{\pi }\right).$$ (3) In contrast, the effective charge $`\alpha _V(Q)`$ defined from the static heavy quark potential and the effective charge $`\alpha _{>2\text{jets}}`$ defined from $`e^+e^{}`$ annihilation into more than two jets, $`\sigma _{>2\text{jets}}`$, have $`\delta _A=0`$. One can define effective charges for virtually any quantity calculable in perturbative QCD; e.g., moments of structure functions, ratios of form factors, jet observables, and the effective potential between massive quarks. In the case of decay constants of the $`Z`$ or the $`\tau `$, the mass of the decaying system serves as the physical scale in the effective charge. In the case of multi-scale observables, such as the two-jet fraction in $`e^+e^{}`$ annihilation, the arguments of the effective coupling $`\alpha _{2jet}(s,y)`$ correspond to the overall available energy and characteristic kinematical jet mass fraction. Effective charges are defined in terms of observables and, as such, are renormalization-scheme and renormalization-scale independent. The scale $`Q`$ which enters a given effective charge corresponds to its physical momentum scale. The total derivative of each effective charge $`\alpha _A(Q)`$ with respect to the logarithm of its physical scale is given by the Gell Mann-Low function: $$\mathrm{\Psi }_A[\alpha _A(Q,m),Q/m]\frac{d\alpha _A(Q,m)}{d\mathrm{log}Q},$$ (4) where the functional dependence of $`\mathrm{\Psi }_A`$ is specific to the effective charge $`\alpha _A`$. Here $`m`$ refers to the quark’s pole mass. The pole mass is universal in that it does not depend on the choice of effective charge. It should be emphasized that the Gell Mann-Low $`\mathrm{\Psi }`$ function is a property of a physical quantity, and it is thus independent of conventions such as the renormalization procedure and the choice of renormalization scale. A central feature of quantum chromodynamics is asymptotic freedom; i.e., the monotonic decrease of the QCD coupling $`\alpha _A(Q^2)`$ at large spacelike scales. The empirical test of asymptotic freedom is the verification of the negative sign of the Gell Mann-Low function at large momentum transfer, a feature which must in fact be true for any effective charge. In perturbation theory, $$\mathrm{\Psi }_A=\mathrm{\Psi }_A^{\{0\}}\frac{\alpha _A^2}{\pi }\mathrm{\Psi }_A^{\{1\}}\frac{\alpha _A^3}{\pi ^2}\mathrm{\Psi }_A^{\{2\}}\frac{\alpha _A^4}{\pi ^3}+\mathrm{}$$ (5) At large scales $`Q^2>>m^2`$, where the quarks can be treated as massless, the first two terms are universal and basically given by the first two terms of the usual QCD $`\beta `$ function for $`N_C=3`$ $`\mathrm{\Psi }_A^{\{0\}}`$ $`=`$ $`{\displaystyle \frac{\beta _0}{2}}={\displaystyle \frac{11}{2}}{\displaystyle \frac{1}{3}}N_{F,A}^{\{0\}},`$ $`\mathrm{\Psi }_A^{\{1\}}`$ $`=`$ $`{\displaystyle \frac{\beta _1}{8}}={\displaystyle \frac{51}{4}}+{\displaystyle \frac{19}{12}}N_{F,A}^{\{1\}}.`$ (6) Unlike the $`\beta `$-function which controls the renormalization scale dependence of bare couplings such as $`\alpha _{\overline{M}S}(\mu )`$, the $`\psi `$ function is analytic in $`Q^2/m^2`$. In the case of the $`\alpha _V`$ scheme, the effective charge defined from the heavy quark potential, the functional dependence of $`N_{F,V}(Q^2/m^2)`$ is known to two loops . The purpose of this paper is to develop an accurate method for extracting the Gell Mann-Low function from measurements of an effective charge in a manner which avoids the biases and uncertainties present either in a standard fit or in numerical differentiation of the data. We will show that one can indeed obtain strong constraints on $`\mathrm{\Psi }_A^{\{0\}}`$ and $`\mathrm{\Psi }_A^{\{1\}}`$ from generalized moments of the measured quantities which define the effective charge. We find that the weight function $`f(\xi )`$ which defines the effective charge $`\alpha _{Af}(\mathrm{\Lambda })`$ from an integral of the effective charge $`\alpha _A(Q)`$ can be chosen to produce maximum sensitivity to the Gell-Mann Low function. As an example we will apply the method to the $`e^+e^{}`$ annihilation into more than two jets. Clearly one could also extract the Gell Mann-Low function directly from a fit to the data, but the fact that we are dealing with a logarithmic derivative introduces large uncertainties . Our results minimize some of these uncertainties. In addition, our analysis provides a new class of commensurate relations between observables which are devoid of renormalization scheme and scale artifacts. One can define generalized effective charges from moments of the observables. The classic example is $`\alpha _\tau (\mathrm{\Lambda })`$ where $`\mathrm{\Lambda }`$ is the generalization of the lepton mass. The relevant point is that $`R_\tau `$ can be written as an integral of $`R_{e^+e^{}}`$ , as follows: $$R_\tau (\mathrm{\Lambda }^2)=\frac{2}{_fq_f^2}_0^{\mathrm{\Lambda }^2}\frac{ds}{\mathrm{\Lambda }^2}\left(1\frac{s}{\mathrm{\Lambda }^2}\right)^2\left(1+\frac{2s}{\mathrm{\Lambda }^2}\right)R_{e^+e^{}}(s),$$ (7) where $`q_f`$ are the quark charges. As a consequence of the mean value theorem, the associated effective charges are related by a scale shift $$\alpha _\tau (\mathrm{\Lambda })=\alpha _R(\sqrt{s}=\mathrm{\Lambda }_\tau ),$$ (8) The ratio of scales $`\mathrm{\Lambda }_\tau /\mathrm{\Lambda }`$ in principle is predicted by QCD : The prediction at NLO is $$\frac{\mathrm{\Lambda }_\tau }{\mathrm{\Lambda }}=\mathrm{exp}\left[\frac{19}{24}\frac{169}{128}\frac{\alpha _R(\mathrm{\Lambda }_\tau )}{\pi }+\mathrm{}\right].$$ (9) Such relations between observables are called commensurate scale relations (CSR) . The relation between $`R_\tau `$ and $`R_{e^+e^{}}`$ suggests that we can obtain additional useful effective charges by changing the functional weight appearing in the integrand. Indeed it has been shown that, starting from any given observable $`𝒪_A`$ we can obtain new effective charges $`\alpha _{Af}`$ by constructing the following quantity $$𝒪_{Af}(\mathrm{\Lambda })=C_{\mathrm{\Lambda }_1^2(\mathrm{\Lambda })}^{\mathrm{\Lambda }_2^2(\mathrm{\Lambda })}\frac{ds}{\mathrm{\Lambda }^2}f\left(\frac{\sqrt{s}}{\mathrm{\Lambda }}\right)𝒪_A(\sqrt{s}),$$ (10) where $`C`$ is a constant and $`f(\xi )`$ is a positive arbitrary integrable function. In order for $`𝒪_{Af}`$ to define an effective charge $`\alpha _{Af}`$ through $$𝒪_{Af}(\mathrm{\Lambda })=𝒪_{Af}^0\left(\delta _A+\frac{\alpha _{Af}(\mathrm{\Lambda })}{\pi }\right),$$ (11) it is necessary that $`\mathrm{\Lambda }_1(\mathrm{\Lambda })=\lambda _1\mathrm{\Lambda }`$ and $`\mathrm{\Lambda }_2(\mathrm{\Lambda })=\lambda _2\mathrm{\Lambda }`$, with both $`\lambda _1`$ and $`\lambda _2`$ constant. Then, by the mean value theorem, $`\alpha _{Af}`$ is related again to $`\alpha _A`$ by a scale shift $$\alpha _{Af}(\mathrm{\Lambda })=\alpha _A(\mathrm{\Lambda }_{Af}),$$ (12) with $`\mathrm{\Lambda }_1<\mathrm{\Lambda }_{Af}<\mathrm{\Lambda }_2`$. An important observation is that PQCD predicts $`\lambda _{Af}=\mathrm{\Lambda }_{Af}/\mathrm{\Lambda }`$ to leading twist. If we ignore quark masses so that the two first coefficients of the Gell-Mann Low function are constant, one has $`{\displaystyle \frac{\alpha _A(\sqrt{s})}{\pi }}={\displaystyle \frac{\alpha _A(\mathrm{\Lambda })}{\pi }}{\displaystyle \frac{\mathrm{\Psi }_0}{2}}\mathrm{ln}\left({\displaystyle \frac{s}{\mathrm{\Lambda }^2}}\right)\left({\displaystyle \frac{\alpha _A(\mathrm{\Lambda })}{\pi }}\right)^2+`$ (13) $`+{\displaystyle \frac{1}{4}}\left[\mathrm{\Psi }_0^2\mathrm{ln}^2\left({\displaystyle \frac{s}{\mathrm{\Lambda }^2}}\right)2\mathrm{\Psi }_1\mathrm{ln}\left({\displaystyle \frac{s}{\mathrm{\Lambda }^2}}\right)\right]\left({\displaystyle \frac{\alpha _A(\mathrm{\Lambda })}{\pi }}\right)^3\mathrm{}`$ If we now use eqs.(10) and (11), we find $`{\displaystyle \frac{\alpha _{Af}(\mathrm{\Lambda })}{\pi }}`$ $`=`$ $`{\displaystyle \frac{\alpha _A(\mathrm{\Lambda })}{\pi }}{\displaystyle \frac{\mathrm{\Psi }_0}{2}}{\displaystyle \frac{I_{1f}}{I_{0f}}}\left({\displaystyle \frac{\alpha _A(\mathrm{\Lambda })}{\pi }}\right)^2`$ (14) $`+`$ $`{\displaystyle \frac{1}{4}}\left[\mathrm{\Psi }_0^2{\displaystyle \frac{I_{2f}}{I_{0f}}}2\mathrm{\Psi }_1{\displaystyle \frac{I_{1f}}{I_{0f}}}\right]\left({\displaystyle \frac{\alpha _A(\mathrm{\Lambda })}{\pi }}\right)^3\mathrm{},`$ where $`I_{lf}=_{\lambda _1^2}^{\lambda _2^2}f(\xi )(\mathrm{ln}\xi ^2)^l𝑑\xi ^2`$ is independent of the choices of observable $`A`$ and scale $`\mathrm{\Lambda }`$, but only provided that $`\mathrm{\Lambda }_1(\mathrm{\Lambda })=\lambda _1\mathrm{\Lambda }`$ and $`\mathrm{\Lambda }_2(\mathrm{\Lambda })=\lambda _2\mathrm{\Lambda }`$. Replacing $`s`$ by $`\mathrm{\Lambda }_{Af}^2`$ in eq. (13) and comparing with eq. (14), we find $$\lambda _{Af}=\mathrm{exp}\left\{\frac{I_{1f}}{2I_{0f}}+\frac{\mathrm{\Psi }_0}{4}\left[\left(\frac{I_{1f}}{I_{0f}}\right)^2\frac{I_{2f}}{I_{0f}}\right]\frac{\alpha _A(\mathrm{\Lambda })}{\pi }\mathrm{}\right\}.$$ (15) In general the commensurate scale relation will have the following expansion $$\mathrm{ln}\lambda _{Af}(\mathrm{\Lambda })=\underset{n=0}{\overset{\mathrm{}}{}}a_f^{(n)}\left(\frac{\alpha _A(\mathrm{\Lambda })}{\pi }\right)^n,$$ (16) where the first three coefficients are independent of $`A`$. Note that the above formulae are only valid inside regions of constant $`N_F`$ and sufficiently apart from quark thresholds. If we include the mass dependence, the effective charges, by the mean value theorem, are still related by a scale shift, although it cannot be written in the simple form of eq. (15). Indeed, even the lowest order of $`\lambda _{Af}`$ would have a small dependence on the energy and the effective number of flavors appearing in $`\mathrm{\Psi }_0`$. ## 2 Obtaining the Gell Mann-Low function directly from observables The main practical obstacle in determining the Gell Mann-Low function from experiment is that it is a logarithmic derivative. One can try to obtain the value of the parameters of the $`\mathrm{\Psi }`$ function from a direct fit to the data using the QCD forms, but any approximation to the derivative of the experimental results implicitly requires extrapolation or interpolation of the data. In order to observe a significant variation of the effective charge $`\alpha _A`$ one needs to compare two vastly separated scales. This is illustrated in Fig.1. However, to approximate $`\mathrm{\Psi }(\sqrt{s})\mathrm{\Delta }\alpha _A(\sqrt{s})/(\mathrm{\Delta }\mathrm{ln}\sqrt{s})`$ with a huge separation between $`\sqrt{s}`$ and $`\sqrt{s^{}}`$ is not very accurate since then the value for $`\mathrm{\Delta }\alpha _A/\mathrm{\Delta }\mathrm{ln}\sqrt{s}`$ is the slope of the $`Q`$ straight line in Fig. 1 instead of that of $`P`$, which gives an $`𝒪(\mathrm{\Delta }\mathrm{ln}\sqrt{s})^2`$ error. If we want to obtain $`\mathrm{\Psi }`$ from a finite difference approximation, we need to interpolate $`\mathrm{\Delta }\mathrm{ln}\sqrt{s}0`$, but in this case the experimental errors will most likely be much larger than the required precision. Such an interpolation procedure has already been applied in ref. near the $`\tau `$ region to test the running of $`\alpha _s`$ (including appropriate corrections to the leading twist formalism). In this energy region the value of the QCD coupling is rather large, and the interpolation yields evidence for some running. However, it has also been pointed out in , that the value of the coupling extrapolated from the $`\tau `$ region to high energies appears small compared to direct determinations. In the next section we shall use the effective charge formalism to derive several expressions within leading twist QCD which relate the intrinsic $`\mathrm{\Psi }_A`$ function of $`\alpha _A`$ directly to the observables $`𝒪_A`$. We shall show that with just three data points we can obtain good sensitivity to the value of $`\mathrm{\Psi }_0`$ without any numerical differentiation or fit. #### 2.0.1 Differential Commensurate Scale Relations Let us formally differentiate eq. (10) with respect to $`\mathrm{\Lambda }`$ $`{\displaystyle \frac{d𝒪_{Af}(\mathrm{\Lambda })}{d\mathrm{\Lambda }}}`$ $`=`$ $`{\displaystyle \frac{2C}{\mathrm{\Lambda }}}\left[\lambda _2^2f(\lambda _2)𝒪_A(\mathrm{\Lambda }_2)\lambda _1^2f(\lambda _1)𝒪_A(\mathrm{\Lambda }_1)\right]`$ (17) $``$ $`{\displaystyle \frac{2𝒪_{Af}(\mathrm{\Lambda })}{\mathrm{\Lambda }}}{\displaystyle \frac{C}{\mathrm{\Lambda }}}{\displaystyle _{(\lambda _1\mathrm{\Lambda })^2}^{(\lambda _2\mathrm{\Lambda })^2}}{\displaystyle \frac{ds}{\mathrm{\Lambda }^2}}𝒪_A(\sqrt{s}){\displaystyle \frac{\sqrt{s}}{\mathrm{\Lambda }}}{\displaystyle \frac{df(\sqrt{s}/\mathrm{\Lambda })}{d(\sqrt{s}/\mathrm{\Lambda })}}.`$ The first term in the right-hand side can be obtained directly from the data on $`𝒪_A`$. This is also the case for the second term, after using eqs. (2) and (12), since $$𝒪_{Af}(\mathrm{\Lambda })=𝒪_{Af}^0\left(\delta _A+\frac{\alpha _{Af}(\mathrm{\Lambda })}{\pi }\right)=𝒪_{Af}^0\left(\delta _A+\frac{\alpha _A(\mathrm{\Lambda }_f)}{\pi }\right)=\frac{𝒪_{Af}^0}{𝒪_A^0}𝒪_A(\mathrm{\Lambda }_f),$$ (18) Note that $`𝒪_{Af}^0`$ and $`𝒪_A^0`$ are known constants. Finally, there is a choice of $`f(\xi )`$ which allows us to recast the third term in the right-hand side of eq. (17) and provide a direct relation between the data and the effective charge. Namely, we choose $$\xi \frac{df(\xi )}{d\xi }=\rho f(\xi ),$$ (19) with $`\rho `$ any real number. That is, up to an irrelevant multiplicative constant, we take $$f(\xi )=\xi ^\rho .$$ (20) With this choice eq. (17) can be simply written as $$\frac{d𝒪_{A\rho }(\mathrm{\Lambda })}{d\mathrm{\Lambda }}=\frac{2C}{\mathrm{\Lambda }}\left[\lambda _2^{\rho +2}𝒪_A(\mathrm{\Lambda }_2)\lambda _1^{\rho +2}𝒪_A(\mathrm{\Lambda }_1)\right]\frac{\rho +2}{\mathrm{\Lambda }}𝒪_{A\rho }(\mathrm{\Lambda }).$$ (21) Note that, to simplify the notation, we have substituted the $`f`$ subscript by $`\rho `$. In terms of $`\mathrm{\Psi }_A`$ this means $$\mathrm{\Psi }_{A\rho }(\mathrm{\Lambda })=\mathrm{\Lambda }\frac{d\alpha _{A\rho }(\mathrm{\Lambda })}{d\mathrm{\Lambda }}=\frac{\pi \mathrm{\Lambda }}{𝒪_{A\rho }}\frac{d𝒪_{A\rho }(\mathrm{\Lambda })}{d\mathrm{\Lambda }}.$$ (22) But using its definition, we can easily see that $$𝒪_{A\rho }^0=\frac{2C𝒪_A^0}{(\rho +2)}(\lambda _2^{\rho +2}\lambda _1^{\rho +2}),$$ (23) so that, using eq. (18), we arrive at $$\mathrm{\Psi }_{A\rho }(\mathrm{\Lambda })=\pi \frac{\rho +2}{𝒪_A^0}\left[\frac{\lambda _2^{\rho +2}𝒪_A(\mathrm{\Lambda }_2)\lambda _1^{\rho +2}𝒪_A(\mathrm{\Lambda }_1)}{\lambda _2^{\rho +2}\lambda _1^{\rho +2}}𝒪_A(\mathrm{\Lambda }_\rho )\right],$$ (24) Note that we have just written $`\mathrm{\Psi }_{A\rho }(\mathrm{\Lambda })`$ directly in terms of observables. Therefore, we have related the universal $`\mathrm{\Psi }_0`$ and $`\mathrm{\Psi }_1`$ coefficients directly to observables, without any dependence on the renormalization scheme or scale. Up to this point $`\mathrm{\Lambda }_1`$ and $`\mathrm{\Lambda }_2`$ are arbitrary. In order to illustrate the meaning of eq.(24), we now choose $`\lambda _1=0`$ and $`\lambda _2=1`$, so that eq.(24) becomes: $$\mathrm{\Psi }_{A\rho }(\mathrm{\Lambda })=\pi \frac{\rho +2}{𝒪_A^0}\left[𝒪_A(\mathrm{\Lambda })𝒪_A(\mathrm{\Lambda }_\rho )\right],$$ (25) Let us remark that, although it may look similar, the above equation is not the finite difference approximation $$\mathrm{\Psi }_A(\mathrm{\Lambda })\frac{\pi \mathrm{\Lambda }}{𝒪_A^0}\frac{𝒪_A(\mathrm{\Lambda })𝒪_A(\mathrm{\Lambda }\mathrm{\Delta }\mathrm{\Lambda })}{\mathrm{\Delta }\mathrm{\Lambda }}+O(\mathrm{\Delta }\mathrm{\Lambda }^2)$$ (26) which is a good numerical approximation to $`\mathrm{\Psi }_A(\mathrm{\Lambda })`$ when $`\mathrm{\Delta }\mathrm{\Lambda }`$ is very small. In contrast, eq. (24), is exact (at leading twist) no matter whether $`\mathrm{\Lambda }\mathrm{\Lambda }_\rho `$ is big or small. However, we do not want to set $`\lambda _1=0`$, since then the integrated effective charges defined in eq.(10), contain higher twist contributions which are unsuppressed at low energies, and our leading twist formulae would be invalid in practice. In addition, some observables like the number of jets produced in $`e^+e^{}`$ annihilation are only well defined above some energy, which becomes a lower cutoff in the integral of eq.(10). Nevertheless, by choosing $`\mathrm{\Lambda }`$ and $`\lambda _2`$ appropriately, we can obtain any value of $`\mathrm{\Lambda }_10`$ and $`\mathrm{\Lambda }_20`$, even if we set $`\lambda _1=1`$, and so we will do so in the following. That is: $$\mathrm{\Psi }_{A\rho }(\mathrm{\Lambda })=\pi \frac{\rho +2}{𝒪_A^0}\left[\frac{\lambda _2^{\rho +2}𝒪_A(\mathrm{\Lambda }_2)𝒪_A(\mathrm{\Lambda })}{\lambda _2^{\rho +2}1}𝒪_A(\mathrm{\Lambda }_\rho )\right],$$ (27) which is an exact formula relating $`\mathrm{\Psi }_A`$ with the observable $`𝒪_A`$ at three scales $`\mathrm{\Lambda }<\mathrm{\Lambda }_\rho <\mathrm{\Lambda }_2`$. It happens, however, that we are interested in measuring not the $`\mathrm{\Psi }_{A\rho }`$ intrinsic function but $`\mathrm{\Psi }_A`$ itself. We thus arrive at our final result: $$\mathrm{\Psi }_A(\mathrm{\Lambda }\lambda _\rho (\mathrm{\Lambda }))\left[1+\frac{\lambda _\rho ^{}}{\lambda _\rho }\right]=\pi \frac{\rho +2}{𝒪_A^0}\left[\frac{\lambda _2^{\rho +2}𝒪_A(\mathrm{\Lambda }_2)𝒪_A(\mathrm{\Lambda })}{\lambda _2^{\rho +2}1}𝒪_A(\mathrm{\Lambda }\lambda _\rho )\right].$$ (28) where we have also defined $`\lambda _\rho =\mathrm{\Lambda }_\rho /\mathrm{\Lambda }`$. Note that $`\mathrm{\Psi }_A`$ appears in the above equation both at $`\mathrm{\Lambda }_\rho `$ and $`\mathrm{\Lambda }`$ through the $`\lambda _\rho ^{}`$ coefficient, defined as $`d\lambda /dLog\mathrm{\Lambda }`$, which only vanishes at leading order. Therefore, if we include higher order contributions the above equation is not enough to determine $`\mathrm{\Psi }_A`$ at one given scale. Let us work out first the implications of eq.(28) at leading order, since it contains all the relevant features of our approach. ### 2.1 Leading order Suppose then that we had three experimental data points at $`s_a<s_b<s_c`$. In order to apply eq. (27), we first identify $`\mathrm{\Lambda }_2=\sqrt{s_c/s_a}`$ and then we obtain the $`\rho `$ such that $`\sqrt{s_a}=\sqrt{s_b}/\lambda _\rho `$. The $`I_{k\rho }`$ integrals are given by $$I_{k\rho }=\frac{k!}{\rho /2+1}\underset{j=1}{\overset{2}{}}\left[(1)^j\lambda _2^{\rho +2}\underset{l=0}{\overset{k}{}}\left(\frac{(\mathrm{ln}\lambda _2^2)^{(kl)}(1)^l}{(\rho /2+1)^l(kl)!}\right)\frac{(1)^k}{(\rho /2+1)^k}\right].$$ (29) Thus, at leading order we have to obtain $`\rho `$ from $$\mathrm{ln}\frac{s_b}{s_a}=2\mathrm{ln}\lambda _\rho =\frac{I_{1\rho }}{I_{0\rho }}=\frac{s_c^{\rho /2+1}\mathrm{ln}(s_c/s_a)}{s_c^{\rho /2+1}s_a^{\rho /2+1}}\frac{1}{\rho /2+1},$$ (30) which can be evaluated numerically. As we have already commented, at leading order $`\lambda ^{}=0`$, and therefore $$\mathrm{\Psi }_A(\sqrt{s_b})=\pi \frac{\rho +2}{𝒪_A^0}\left[\frac{s_c^{\rho /2+1}𝒪_A(\sqrt{s_c})s_a^{\rho /2+1}𝒪_A(\sqrt{s_a})}{s_c^{\rho /2+1}s_a^{\rho /2+1}}𝒪_A(\sqrt{s_b})\right].$$ (31) Let us remark once more that these are leading-twist formulae, and $`s_a,s_b,s_c`$ should lie in a range where higher twist effects are negligible. ### 2.2 Beyond leading order As we have already seen, if we go beyond the leading order contributions, we have to use eq.(28), which does not completely determine the value of $`\mathrm{\Psi }_A`$ at a single scale. In principle, we need an additional equation. In fact, the $`\lambda ^{}`$ term can be neglected. Intuitively, this is due to the very slow evolution of $`\alpha _A`$. Let us give some numerical values; first, we will write $$\frac{\lambda _\rho ^{}}{\lambda _\rho }=\mathrm{\Psi }_A(\mathrm{\Lambda })\mathrm{\Omega }_\rho (\mathrm{\Lambda }),$$ (32) with $$\mathrm{\Omega }_\rho (\mathrm{\Lambda })\frac{d\mathrm{ln}\lambda _\rho }{d(\alpha _A(\mathrm{\Lambda }))}=2\underset{n=1}{\overset{\mathrm{}}{}}na_\rho ^{(n)}\left(\frac{𝒪_A(\mathrm{\Lambda })}{𝒪_A^0}\delta _A\right)^{n1}.$$ (33) From PQCD we know that the expansion of $`\mathrm{\Psi }_A`$ starts with $`\alpha _A^2`$. Thus, the $`\lambda ^{}`$ term in eq.(28) is an $`O(\alpha _A^4)`$ effect. It should only be taken into account if we are interested in $`\mathrm{\Psi }`$ up to that order. Numerically, the expected value of $`\mathrm{\Psi }_A(\mathrm{\Lambda })`$ at the energies we will be using, ranges from $`10^2`$ to $`2\times 10^2`$ at most. In addition, $`\mathrm{\Omega }`$ ranges from $`3\times 10^2`$ to 0.5. Thus, even in the worst case, the $`\lambda ^{}`$ term contribution would be slightly smaller than $`1\%`$ of $`\mathrm{\Psi }`$. If that term is to be kept, then we need and additional equation involving a fourth data point. We have found that the final error estimate increases since it is much harder to accommodate four points sufficiently separated within a given energy range. It seems that $`1\%`$ accuracy is the lower limit for this method. If additional higher twist corrections are included, it could be possible to extend the energy range to separate the points and improve the precision. Therefore, in what follows we will use eq. (31). However, the NLO $`\rho `$ parameter is now obtained by solving numerically the equation $`\mathrm{ln}{\displaystyle \frac{s_b}{s_a}}`$ $`=`$ $`{\displaystyle \frac{s_c^{\rho /2+1}\mathrm{ln}^2(s_b/s_a)}{s_c^{\rho /2+1}s_a^{\rho /2+1}}}{\displaystyle \frac{1}{\rho /2+1}}`$ $`+`$ $`{\displaystyle \frac{\mathrm{\Psi }_0}{2}}\left[{\displaystyle \frac{(s_as_c)^{\rho /2+1}\mathrm{ln}^2(s_b/s_a)}{(s_c^{\rho /2+1}s_a^{\rho /2+1})^2}}{\displaystyle \frac{1}{(\rho /2+1)^2}}\right]\left({\displaystyle \frac{𝒪_A(\sqrt{s_a})}{𝒪_A^0}}\delta _A\right),`$ where $`s_a<s_b<s_c`$ and $`\sqrt{s_b}=\lambda _\rho \sqrt{s_a}`$ and $`\sqrt{s_c}=\lambda _2\sqrt{s_a}`$. Note that now $`\mathrm{\Psi }_0`$ is an input, but the output is the NLO $`\mathrm{\Psi }`$ function. ## 3 Error estimates Although they have inspired our approach, observables with $`\delta _A0`$ are not well suited for our method, because the relative error in $`𝒪_A(E)`$ becomes at least one order of magnitude larger for the effective charge $`\alpha _A(E)`$. For example, using the $`e^+e^{}`$ hadronic ratio defined in Sect.1, if we introduce a $`1\%`$ error in $`R_{e^+e^{}}`$, the error in $`\alpha _R`$ is $`O(20\%)`$ and we have to separate the data points over five orders of magnitude to obtain $`\mathrm{\Psi }_R`$ with a $`10\%`$ precision. In practice, that renders the method useless. The problem we have described is avoided if we use an observable with $`\delta _A=0`$. That is the case, for instance, of the $`e^+e^{}`$ annihilation in more than two jets, $`\sigma _{>2\text{jets}}(s,y)=\sigma _{\text{tot}}\sigma _{2\text{jets}}`$, where $`y`$ is used to define when two partons are unresolved (i.e. their invariant mass squared is less than $`ys`$). This process does not occur in the parton model since it requires, at least, one gluon. Note that $`\mathrm{\Psi }_0`$ and $`\mathrm{\Psi }_1`$ are independent of $`y`$. At LO we can work with exact results, but as soon as we introduce higher orders, there is some degree of truncation in the formulae. We have therefore first constructed simulated data following a model that corresponds to the exact LO equations. Let us remark that these are models, not QCD. They are obtained by the truncation of $`\alpha _A`$ at a given order. Thus, in principle, they will have some different features from QCD, as for instance, some residual scale dependence. In the real world this will not occur. However, we have worked out these examples for illustrative purposes to obtain a rough estimate of the errors. #### 3.0.1 Leading order What we call the LO model is to use $$\frac{\alpha _A(Q)}{\pi }=\frac{\alpha (M_Z)}{\pi }\frac{\mathrm{\Psi }_0}{2}\mathrm{ln}\left(\frac{Q^2}{M_Z^2}\right)\left(\frac{\alpha _A(M_Z)}{\pi }\right)^2,$$ (35) exactly. We have taken $`\alpha _A(M_Z)`$ as the reference value for simplicity. Note, however, that the derivative of the above expression is $$\mathrm{\Psi }_A=\frac{\mathrm{\Psi }_0}{2}\left(\frac{\alpha _A(M_Z)}{\pi }\right)^2,$$ (36) which is a constant which differs by $`O(\alpha /\pi )^3`$ terms from the LO PQCD result $$\mathrm{\Psi }_A(Q)=\frac{\mathrm{\Psi }_0}{2}\left(\frac{\alpha _A(Q)}{\pi }\right)^2.$$ (37) In Table 1 we can see the estimates of the relative errors in our determination of $`\mathrm{\Psi }_A`$, which depend on the different position of the data points, as well as in their errors $`\mathrm{\Delta }𝒪_A`$. Since the observable vanishes in the parton model, the relative error in $`\alpha _A`$ is exactly that of $`𝒪_A`$. The results in the table deserve some comments. * First, the values of $`\sqrt{s_a}`$ and $`\sqrt{s_c}`$ have to be chosen to maximize their distance, within a region of constant $`N_F`$. Thinking in terms of $`\sigma _{>2\text{jets}}`$, they correspond either to the region where both energies are sufficiently above the b-quark pair threshold but still below $`t\overline{t}`$ production, or both are above the $`t\overline{t}`$ pair threshold, in regions accessible at NLC. * Second, we have chosen the same relative error for the measurements at the three points. The intermediate energy $`\sqrt{s_b}`$ is then tuned to minimize the error, which is obtained assuming the three $`𝒪_A`$ measurements are independent. Let us remark once again that we have not used at any moment the value of $`\mathrm{\Psi }_0`$, which is obtained from the data using this method. If we want to use higher order contributions, using the value of $`\mathrm{\Psi }_0`$ as an input, we would obtain information about higher order coefficients, like $`\mathrm{\Psi }_1`$ if we were to work at NLO. #### 3.0.2 Beyond leading order The NLO model is now given by: $`{\displaystyle \frac{\alpha _A(Q)}{\pi }}`$ $`=`$ $`{\displaystyle \frac{\alpha (M_Z)}{\pi }}{\displaystyle \frac{\mathrm{\Psi }_0}{2}}\mathrm{ln}\left({\displaystyle \frac{Q^2}{M_Z^2}}\right)\left({\displaystyle \frac{\alpha _A(M_Z)}{\pi }}\right)^2`$ (38) $`+`$ $`{\displaystyle \frac{1}{4}}\left[\mathrm{\Psi }_0^2\mathrm{ln}^2\left({\displaystyle \frac{Q^2}{\mathrm{\Lambda }}}\right)2\mathrm{\Psi }_1\mathrm{ln}\left({\displaystyle \frac{Q^2}{M_Z^2}}\right)\right]\left({\displaystyle \frac{\alpha _A(M_Z)}{\pi }}\right)^3.`$ and therefore, we obtain $$\mathrm{\Psi }_A(Q)=\frac{\mathrm{\Psi }_0}{2}\left(\frac{\alpha _A(Q)}{\pi }\right)^2\frac{\mathrm{\Psi }_1}{2}\left(\frac{\alpha _A(Q)}{\pi }\right)^3,$$ (39) which is the QCD NLO $`\mathrm{\Psi }_A`$ result up to $`O(\alpha /\pi )^4`$ terms. In contrast with the LO case, obtaining $`\rho `$ now requires some truncation of the formulae when passing from eqs. (13) and (10) to eq. (14). This is very interesting since we can thus obtain an estimate of the theoretical error due to truncation, which will be present in the real case too. It can be seen in Table 2 in the rows where $`\mathrm{\Delta }𝒪_A=0`$, and it is usually $`O(1\%)`$. Again we have also considered the experimental $`\mathrm{\Delta }𝒪_A(E_i)`$ uncertainties. The final error given in the last column is estimated assuming that the four experimental errors and the one due to truncation are all independent. Note that when passing from a $`1\%`$ experimental error to a $`3\%`$, the total error is not multiplied by 3, since the truncation error does not scale. The fact that we obtain larger errors in the NLO case may seem surprising, but it is not. The reason is that the LO is a very crude approximation of the $`\mathrm{\Psi }_A`$ QCD scaling behavior. In the LO model, the $`\mathrm{\Psi }`$ function was a constant, but in the NLO it changes with the energy scale, as it occurs in the realistic case. Indeed, the evolution of $`\alpha _A`$ at high energies becomes much slower so that the difference between $`\alpha _A`$ at two given points is smaller at NLO than at LO. Hence, for the same relative errors, the relative uncertainties in the NLO $`\mathrm{\Psi }`$ function are much bigger. Of course, we expect the real data to show a behavior much closer to the NLO model. ### 3.1 Using more than three points The advantage of fitting the data is that we can reduce the errors by larger statistics. But that is also true for our method. Up to now we have only used three points of data, but in the realistic case we expect to have several points at each energy range. It is then possible to form many triplets of data points, one at low energies ($`\sqrt{s_a}`$), another at intermediate energies ($`\sqrt{s_b}`$), and a last one in the highest range ($`\sqrt{s_b}`$). Each one of these triplets will yield different values and errors for $`\mathrm{\Psi }`$, which can later be treated statistically, thus decreasing the error estimates given in Table 2. ## 4 Conclusions We have obtained an exact and very simple relation between the Gell Mann-Low $`\mathrm{\Psi }`$ function of an effective charge of an observable and its integrals. These results are renormalization-scheme and renormalization-scale independent. By choosing specific weight functions, these relations can provide an experimental determination of the PQCD $`\mathrm{\Psi }`$ function, thus testing the theory and setting bounds on the properties of new particles that would modify the expected QCD behavior. We have shown that a good candidate for this study is the $`e^+e^{}`$ annihilation to more than two jets, since it is a pure QCD process. Even within the simple leading-twist formalism, which limits the applicability range, we have found that with just three precise measurements in present or presently planned accelerators, it could be possible to determine the $`\mathrm{\Psi }`$ function without making a QCD fit or any interpolation and numerical differentiation of the data, eliminating the specific uncertainties of these methods. Thus we can obtain a determination of $`\mathrm{\Psi }`$ with different systematics. It also seems possible to extend the method and ideas, to include higher twist effects which will allow the use of a wider range of energies. This could result in an even more powerful set of tests of perturbative QCD. ## Acknowledgments C.M. and J.R.P. thank the Theory Group at SLAC for their kind hospitality. J.R.P. acknowledges the Spanish Ministerio de Educación y Cultura for financial support, as well as the Departamento de Física de Partículas of the Universidade de Santiago de Compostela for its hospitality. We also thank M. Melles, J. Rathsman and N. Toumbas for helpful conversations.
no-problem/9906/quant-ph9906102.html
ar5iv
text
# 1 INTRODUCTION ## 1 INTRODUCTION Quantum interference of individual systems has recently been found capable of detecting objects without transferring energy to them. The effect has been named interaction-free detection<sup>3</sup><sup>3</sup>3Niels Bohr would most likely argue against the name in the following way: “It is true that in the measurements under consideration any direct mechanical interaction of the system and the measuring agencies is excluded, but …the procedure of measurements has an essential influence on the conditions on which the very definition of the physical quantities in question rests…\[T\]hese conditions must be considered as an inherent element of any phenomenon to which the term “\[interaction\]” can be unambiguously applied.” However, the name has been rather unanimously accepted in the quantum parlance and it is likely to stay there. and was based on the void detections which destroy path indistinguishability. In 1986 Pavičić formulated this in the following way. “Consider a photon experiment shown in Fig. 6 which results in an interference in the region D provided we do not know whether it arrived to the region by path $`s_1`$ or by path $`s_2`$. As it is well-known, experimental facts are: If we, after a photon passed the beam splitter B and before it could reach the point C, suddenly introduce a detector in the path $`s_2`$ in the point C and do not detect anything, then it follows that the photon must have taken the path $`s_1`$—and, really, one can detect it in the region D but it does not produce interference there. Quantum mechanically, if we registered the interference in the region D, we could not find an experimental procedure to directly either prove or disprove that the photon uses both paths simultaneously. However, the fact that by detecting nothing in point C we destroy the interference implies that the photon somehow knows of the other path when it takes the first one.” (Ref. 2, pp. 31, 32) Photon’s “knowledge” about the other path one can employ to detect an object (at point C) without transferring even a single quantum of energy to it. The efficiency of such an application with symmetrical Mach-Zehnder interferometer (shown in Fig. 6) is ideally only 25% for single detections and 33% in the long run as shown in Elitzur and Vaidman’s detailed formulation of the void detections in interference experiments in 1993. They also showed that one could increase the ideal efficiency to 50% if an asymmetrical beam splitter were used. In 1995 Kwiat et al. carried out Elitzur and Vaidman’s proposal with an asymmetrical beam splitter using photons obtained in a parametric down conversion. In this way an efficiency close to 50% has been achieved for correlated photons. However, the realization was concerned only with the confirmation of the effect and the 50% efficiency referred to the detected photons which supported the confirmation. For, in the experiment it was necessary to select, with irises, a very small fraction of the photons originally produced in downconversion, which resulted in a net detection efficiency of only 2%. The latter efficiency can be significantly improved but the downconversion can hardly be used for a straightforward realistic interaction-free device. A proposal put forward by Kwiat et al. in 1995 which aims at realistic efficiencies of not hitting tested objects is shown in Fig. 6. The device consists of two coupled resonators (cavities) separated by a highly reflective beam splitter and assumes inserting single photons into one of them. If an object were in the other cavity the probability of it being hit would remain comparatively low. In the absence of an object the photon should, after a certain number of cycles $`N`$, be in the right cavity with certainty. Inserting of a detector in the left cavity should verify the cases. Such an experiment would be very hard to carry out in realistic conditions even if the problem of inserting single photons and the detector were solved. In particular because no firing of the detector should mean the absence of the object and because of the high losses at the mirrors. In 1996 Kwiat et al. put forward another proposal, shown in Fig. 6, which is based on a previous elaboration of the optical Zeno effect. A horizontally polarized photon enters the resonator through the switchable mirror SM which keeps it in for $`N`$ cycles. After each cycle the polarization rotator PR turns the initial photon polarization by an angle $`\alpha `$. When there is no object in the resonator the wave function recombines at the polarizing beam splitter PBS within each cycle so that after $`N=\pi /(2\alpha )`$ cycles it exits the resonator vertically polarized. When there is an object in the resonator within each cycle we have got the Malus probability $`p=\mathrm{cos}^2\alpha `$ of photon passing straight through the horizontally polarizing beam splitter. After $`N`$ cycles the photon—horizontally polarized—exits through SM with the probability $`P=p^N`$. The probability of the object being hit by the photon is therefore $`Q=1P`$. For $`\alpha =1^{}`$, $`Q=3`$%. In this proposal, as opposed to the previous one, we do have different detectable outcomes for presence and absence of objects. Nevertheless, one has to start again from photon pairs generated in a parametric downconversion in order to be able to determine the photon’s entrance time and thus fix the number of cycles, i.e., the moment in which one should let the photon out of the resonator through the switchable mirror $`SM`$. Moreover, the mirror losses will have a detrimental effect on the experiment. Actually, this effect grows with the number of cycles: the larger the latter is, the lower is the ideal theoretical value of $`Q`$, and the bigger are the losses. In 1997 we conceived a different approach using a single monolithic total-internal-reflection resonator (MOTIRR) coupled by two frustrated-total-internal-reflection (FTIR) prisms. The physical principle of the device was essentially the same as for the scheme in Fig. 6 we are going to present in this paper with the only difference that the central loops were confined within a monolithic crystal. The presence of the object causes firing of detector D<sub>r</sub> and the absence causes firing of D<sub>t</sub>. The losses in a MOTIRR are extremely low and go down to 0.3%. Thus a realistic application of interaction-free measurements to the suitable small objects has been enabled. In this paper we are proposing a general purpose interaction-free device for all possible applications foreseen so far, calculate the losses that can be expected in realistic conditions, and show that the present setup evades limitations of the previous proposals. Suggested applications of the interaction-free measurements are numerous and range from the foundational physical experiments and experiments to medicine. Let us just cite some of them. It has been used to show that under a plausible condition Lorentz-invariant realistic interpretations of quantum mechanics are not possible. A preparation of a very well localized atom beam by means of a Mach-Zehnder interferometer for neutrons without physical interaction has been proposed and the first interaction-free experiment with neutrons has already been carried out. A possible interaction-free experiment in quantum dot systems has been discussed. An optical device for erasing fringes of atom interference without disturbing either the spatial wave function or its phase has been proposed thus strengthening the result of Scully et al. Cf. also Ref. 16. Testing of Bose-Einstein condensates (which can be blown apart by even a single photon) has been recently seen as the most immediate possible application. Also a preparation of a superposition at a macroscopic scale. And in the end several more distant possible applications such as selecting particular bacteria without killing them, safe X-ray photography, quantum computer application, etc. In any case we share the feeling that “the situation resembles that of the early years of the laser when scientists knew it would be an ideal solution to many unknown problems.” ## 2 EXPERIMENT Fig. 6 shows an outline of the proposed experiment. When there is no object in the device, an incoming laser beam is almost totally transmitted (up to 98%) into detector D<sub>t</sub> and when there is an object, an incoming laser beam is being (ideally) totally reflected into detector D<sub>r</sub>. The device consists of four prisms forming a resonator. The prisms are designed so that their entrance and exit faces are at right angles to the beam making rectangular loops and are covered with multilayer antireflection coating to minimize reflection losses. The entrance prism is coupled to the adjacent loop prism by the frustrated total reflection, which is an optical version of quantum mechanical tunnelling. Depending on the dimension of the gap between the prisms one can well define reflectivity $`R`$ within the range from $`10^5`$ to 0.99995. The uniqueness of the reflectivity at the gaps and at the same time no reflectivity at the entrance and exit faces of the prisms for each photon is assured by choosing the orientation of the polarization of the incoming laser beam perpendicular to the plane of incidence. As a source of the incoming beam a continuous wave laser (e.g., Nd:YAG) should be used because of its coherence length (up to 300 km) and of its very narrow linewidth (down to 10 kHz in the visible range). Let us now determine the intensity of the beam arriving at detector D<sub>r</sub> when there is no object in the path. Our detailed calculations show that a rigorous description of the device is formally equivalent to a Fabry-Perrot-type of a resonator with standard mirrors up to the phase shifts at the FTR’s which we take into account so as to include it into the phase which is being added by each round-trip. The portion of the incoming beam of amplitude $`A(\omega )`$ reflected at the FTR inner face of the incoming prism is described by the amplitude $`B_0(\omega )=A(\omega )\sqrt{R_1}`$, where $`R`$ is reflectivity. The remaining part of the beam tunnels into the resonator and travels around the resonator guided by one frustrated total reflection (with reflectivity $`\sqrt{R_2}`$ at the face next to the right prism where a part of the beam tunnels out into D<sub>t</sub>) and by two proper total reflections. The losses for such a set-up—as opposed to standard mirror Fabry-Perrot resonators—are very low as calculations and recent experiments show: below 2% for the type presented here and even below 0.3% for the set-up with a monolithic resonator we presented in Ref. 8. The losses in the present set-up are mostly due to absorption and scatter in the multilayer antireflection coatings and the crystals and to a much smaller extent due to imperfect total reflections. After a full round-trip the following portion of the beam joins the directly reflected portion of the beam by tunnelling into the left prism: $`B_1(\omega )=A(\omega )\sqrt{1R_1}\sqrt{R_2}\sqrt{R_3}\sqrt{R_4}\sqrt{1R_1}e^{i\psi }`$, where $`\psi =(\omega \omega _{res})T`$ is the phase added by each round-trip which also includes phase shifts at the gaps; here $`\omega `$ is the frequency of the incoming beam, $`T`$ is the round-trip time, $`\omega _{res}`$ is the resonator frequency, and $`\sqrt{R_3}`$, $`\sqrt{R_4}`$ are the two (realistic, and therefore not equal 1) total reflectivities in which we also include the afore mentioned absorption and scatter (which can be treated as trasmitivities); here we introduce $`\rho =\sqrt{R_3R_4}`$ as a measure of all the losses; $`\rho =1`$ corresponds to an ideal case with no losses. Each subsequent round-trip contributes to the geometric progression: $`B(\omega )`$ $`=`$ $`A(\omega )\{\sqrt{R_1}+(1R_1)\rho \sqrt{R_2}e^{i\psi }[1+\rho \sqrt{R_1R_2}e^{i\psi }+\mathrm{}]\}`$ (1) $`=`$ $`A(\omega )\{\sqrt{R_1}+{\displaystyle \frac{(1R_1)\rho \sqrt{R_2}e^{i\psi }}{1\rho \sqrt{R_1R_2}e^{i\psi }}}\},`$ so as to yield the following probability of the beam being reflected into $`D_r`$ $`B(\omega )B(\omega )^{}=A(\omega )A(\omega )^{}[1{\displaystyle \frac{(1R_1)(1\rho ^2R_2)}{12\rho \sqrt{R_1R_2}\mathrm{cos}\psi +\rho ^2R_1R_2}}].`$ (2) In an analogous way we obtain the probability of the beam being transmitted into $`D_t`$ $`C(\omega )C(\omega )^{}=A(\omega )A(\omega )^{}{\displaystyle \frac{(1R_1)(1R_2)}{12\rho \sqrt{R_1R_2}\mathrm{cos}\psi +\rho ^2R_1R_2}}.`$ (3) Since the frequency of the input laser beam can never precisely match the resonance frequency we make use of a Gaussian wave packet $`A(\omega )=A\mathrm{exp}[𝒯^2(\omega \omega _{res})^2/2]`$, where $`𝒯`$ is the coherence time which obviously must be significantly longer than the round trip time $`T`$. Thus we describe the incident wave by $`E_i^{(+)}(z,t)={\displaystyle _0^{\mathrm{}}}A(\omega )e^{i(kz\omega t)}𝑑\omega ,`$ (4) the reflected wave by: $`E_r^{(+)}(z^{},t)={\displaystyle _0^{\mathrm{}}}B(\omega )e^{i(kz^{}\omega t)}𝑑\omega ,`$ (5) and the transmitted wave by: $`E_t^{(+)}(z^{},t)={\displaystyle _0^{\mathrm{}}}C(\omega )e^{i(kz^{}\omega t)}𝑑\omega ,`$ (6) The energy of the incoming beam is the energy flow integrated over time: $`I_i={\displaystyle _{\mathrm{}}^{\mathrm{}}}E_i^{(+)}(z,t)E_i^{()}(z,t)𝑑t={\displaystyle _0^{\mathrm{}}}A(\omega )A^{}(\omega )𝑑\omega .`$ (7) The energies of the reflected and transmitted beams are given analogously by $`I_r=_0^{\mathrm{}}B(\omega )B^{}(\omega )𝑑\omega `$ and $`I_t=_0^{\mathrm{}}C(\omega )C^{}(\omega )𝑑\omega `$, respectively. The efficiency of the suppression of the reflection into $`D_r`$ is given by $`\eta =1{\displaystyle \frac{I_r}{I_i}}=(1R_1)(1\rho ^2R_2)\mathrm{\Phi },`$ (8) and the efficiency of the throughput into $`D_t`$ by: $`\tau ={\displaystyle \frac{I_t}{I_i}}=(1R_1)(1R_2)\mathrm{\Phi },`$ (9) where $`\mathrm{\Phi }={\displaystyle \frac{{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{\mathrm{exp}[𝒯^2(\omega \omega _{res})^2/2]d\omega }{12\rho \sqrt{R_1R_2}\mathrm{cos}[(\omega \omega _{res})𝒯/a]+\rho ^2R_1R_2}}}{{\displaystyle _0^{\mathrm{}}}\mathrm{exp}[𝒯^2(\omega \omega _{res})^2]𝑑\omega }},`$ (10) where $`a𝒯/T`$ is a ratio of the coherence time $`𝒯`$ and the round-trip time $`T`$. The coherence length should always be long enough ($`a>200`$) to allow sufficiently many round trips (at least 200). $`\mathrm{\Phi }`$ turns out to be very susceptible to the small changes of $`\rho `$ so as to yield rather different outputs of $`\tau `$ in opposition to $`\eta `$. (Cf. Figures 6 and 6.) Obviously both $`\eta `$ and $`\tau `$ should be as close to 1 as possible. A computer optimization shows that this can best be achieved by taking $`R_1=R_2`$. In Figures 6 and 6 we give the values of $`\eta `$ and $`\tau `$, respectively, for $`\rho `$’s which correspond to the throughput $`\tau `$ of about 98% which is considered achievable. The total reflectivities with losses below $`10^6`$ are achievable so that the given values for $`\rho `$ are not the problem so far as the total reflection is considered. As for the throughput $`\tau `$ the given values for $`\rho `$ are also apparently achievable. If however the absorption of antireflection coating turns out to be too high one can always substitute Pellin-Broca prisms with entrance and exit faces at Brewster’s angles (i.e., no reflection losses) for the present prisms with the multilayer antireflection coatings.. In order to carry out the experiment we have to lower the intensity of the beam until it is likely that only one photon would appear within an appropriate time window (1 ms – 1$`\mu `$s $`<`$ coherence time) what allows the intensity in the cavity to build up. The values for $`1\eta `$ are probabilities of detector $`D_r`$ reacting when there is no object in the system. The values for $`\tau `$ are probabilities of detector $`D_t`$ reacting when there is no object in the system. For example, for $`R=0.98`$ and $`\rho =0.9999`$ one obtains $`\eta =0.99`$ and $`\tau =0.98`$. $`\eta `$ and $`\tau `$ in Figs. 6 and 6 are calculated for $`a=500`$, i.e., for 500 round-trips which are multiply assured by continuous wave laser coherence length. Since we did take possible background counts into account by using the Gaussians for the calculation, we can equally rely on $`D_r`$ and on $`D_t`$ firing; also, we can use this fact for tuning the device. A response from $`D_r`$ means that there is an object in the system. In the latter case the probability of the $`D_r`$ response is ideally $`R`$, the probability of a photon hitting the object is $`R(1R)`$, and the probability of photon exiting into $`D_t`$ detector is $`(1R)^2`$. We start each testing by opening a gate for the incident beam and after either D<sub>r</sub> or D<sub>t</sub> fires, the testing is over. The cases when detectors fail to react either because of their inefficiency are not problematic because single photon detectors with 85% efficiency are already available. Such a failure would result in a slightly bigger time window, so that a chance of a photon hitting an object would remain practically unchanged. Thus, a possible 300 km coherence length of cw lasers does not leave any doubt that a real experiment of detecting objects (with an efficiency of over 98%) and without transferring a single quantum of energy to them (with the same efficiency) can be carried out successfully. ## 3 CONCLUSION We have shown that with our resonator based on total reflections and frustrated total reflections, interaction-free measurements can be carried out with a realistically achievable efficiency of 98%. The proposed design makes the device not only very suitable for the foundational experiments reviewed in Sec. 1 but also a good candidate for a more general application, e.g., in medicine, for X-raying patients practically without exposing them to radiation. The latter application was not possible with previous setups because they were all based on mirrors and the losses at X-ray mirrors could be too high for building a realistic interaction-free device. Total reflections we use are however applicable to X-rays and often used for constructing X-ray lasers. On the other hand our setup with two outputs is easily applicable to an interaction-free detection of gray objects where one concludes on the level of grayness by means of the statistics of repeated testings. ACKNOWLEDGMENTS M. P. acknowledges supports of the Alexander von Humboldt Foundation, Bonn, Germany, the Technical University of Vienna, Vienna, Austria, and the Ministry of Science of Croatia. He would also like to thank Johann Summhammer, Atominstitut der Österreichischen Universitäten, Vienna for many valuable discussions.
no-problem/9906/astro-ph9906185.html
ar5iv
text
# 1 Introduction ## 1 Introduction The observation of GeV photons by EGRET during few intense gamma-ray bursts suggested the idea that a large part of events could have a high energy component, not observed so far due to the low fluxes (Catelli 1997a). Gamma-ray emission in the GeV-TeV energy range is predicted by some fireball models (see Baring 1997, for a review). The study of the high energy part of the spectrum would be of great importance to investigate the physical conditions of the emitting region, restricting the range of fundamental parameters as the magnetic field, the density and the bulk Lorentz factor. Unfortunately, due to the cosmological distances of the GRBs sources, the high energy gamma-rays would be absorbed by pair production on starlight photons during their travel towards the Earth. According to Salomon and Stecker (1998), the flux of gamma rays of energy $`E>500`$ GeV is strongly reduced if the distance of the source is $`z>0.1`$. Since the minimum redshift so far measured among six host galaxies is $`z=0.695`$ (Djorgovski 1999), even assuming that GRBs emits gamma rays in the TeV region, most of the spectra observed at Earth would cutoff at energies less than few hundreds GeVs. Observation of high energy GRBs can be performed by ground based experiments, as air shower arrays, detecting the secondary particles produced in the atmosphere by primary gamma-rays. Their large field of view ($`\mathrm{\Omega }\pi `$ sr) and their duty cycle of almost 100$`\%`$ make them suitable to observe unpredictable events as GRBs. Due to the absorption of the high energy part of the spectrum, the detectors must be sensitive to gamma-rays of energy as low as 10-100 GeV. This can be achieved with two basic conditions: a) by operating at very high mountain altitude, in order to increase the number of detectable particles (as an example, the mean number of charged particles produced by a 100 GeV gamma-ray reaching the altitude of 2000 m is $`n_c1.3`$, while at 5000 m $`n_c25`$). b) by disposing of a very large and ”full-coverage” detection surface, in order to detect the largest number of shower particles. These conditions are fully satisfied by the ARGO-YBJ air shower detector. In the following we present the sensitivity of ARGO-YBJ to observe gamma-ray bursts as a function of the GRB spectral characteristics. ## 2 The ARGO-YBJ detector The ARGO-YBJ experiment has been conceived with the aim of detecting small size atmospheric air showers. It is under construction at the Yangbajing High Altitude Cosmic Ray Laboratory (Tibet, China), at an altitude of 4300 m above the sea level. It consists of a central core made by a single layer of Resistive Plate Chambers (RPCs) covering an area of $``$71 $`\times `$ 74 m<sup>2</sup>, sorrounded by an outer ring of 28 clusters of RPCs of area 42 m<sup>2</sup> each, for a total sensitive area $`A_d`$ 6100 m<sup>2</sup>. The detector is uniformely covered by a layer of lead 0.5 cm thick, in order to increase the number of charged particles by converting a fraction of the secondary photons, and to reduce the time spread of the shower front. One of the main field of research of ARGO-YBJ is gamma-ray astronomy in the energy range $`E>100`$ GeV and gamma-ray bursts physics above 10 GeV. A detailed description of the experiment, its capabilities and its physics goals are given by Abbrescia (1996) and Bacci (1998). ## 3 Detection of gamma-ray bursts The detection of high energy gamma-ray bursts can be performed by the ARGO-YBJ detector by using two different modes of operation: a) the ”single particle” technique; b) the ”low multiplicity” technique. In the following we discuss both methods and make a comparison of their sensitivity in the energy range 10 GeV$`<E<`$1 TeV. ### 3.1 The ”single particle” technique (SP). An air shower array can be sensitive to primary energies as low as 10-100 GeV operating in ”single particle mode”, i.e. recording all single secondary particles hitting the detector with energy larger than the detection energy threshold $`E_{th}`$; in this detection mode most of the events are due to solitary muons and electrons of air showers generated by low energy cosmic rays. A gamma-ray burst is detectable if the secondary particles due to the gamma-rays interactions in the atmosphere give a short time excess in the single particle counting rate, of amplitude larger than the statistical fluctuations of the all-sky cosmic rays background. The directions and energies of gamma-rays are not measurable; however this technique could provide a measurement of the total high energy flux and the temporal behaviour of the high energy emission (Vernetto 1999, Aglietta 1999, Cabrera 1999). The effective area to detect a primary gamma-ray of energy $`E`$ and zenith angle $`\theta `$, can be expressed as $`A_{eff}(E,\theta )A_df_gn_e(E,\theta )`$, where $`A_d`$=6100 m<sup>2</sup> is the sensitive area, $`n_e(E,\theta )`$ is the mean number of particles reaching the detector level (with an energy larger than the detection threshold) and $`f_g`$ is the gain factor due to the photons coversion in the lead layer ($`f_g1.1`$). The curve A in Fig.1 shows the ARGO-YBJ effective area as a function of the gamma-ray primary energy $`E`$, for a zenith angle $`\theta `$=20. Given a GRB with an energy spectrum $`dN_\gamma /dE`$ (photons per unit area per unit energy) and zenith angle $`\theta `$ the number of events detected is $`N_{SP}=A_df_gcos\theta 𝑑N_\gamma /𝑑En_e(E,\theta )𝑑E`$. The signal is observable if the number of detected particles $`N_{SP}`$ is significantly larger than the background statistical fluctuations $`N_b=\sqrt{A_dB\mathrm{\Delta }t}`$, where $`B`$ is background rate (events per unit area and unit time) and $`\mathrm{\Delta }t`$ is the GRB duration. The measured single particle background rate at the Yangbajing site is $`B1500`$ events m<sup>-2</sup> s<sup>-1</sup>. Requiring for the GRB signal a minimum statistical significance of 4 standard deviations, the number $`N_{SP}`$ of events in ARGO-YBJ from a GRB of time duration $`\mathrm{\Delta }t`$ = 1 s must be larger than $``$1.2 10<sup>4</sup>. ### 3.2 The low multiplicity technique (LM). The low multiplicity technique (LM) consists in the detection of very small air showers, by requiring at least 6 fired pads per shower (a pad is a detection unit of 56 $`\times `$ 56 cm<sup>2</sup>, see Abbrescia 1996). The effective area $`A_{eff}`$ of ARGO-YBJ to detect primary gamma-rays and protons using this technique has been obtained by simulations and it is shown in Fig.1 for primaries with zenith angle 20 (curve B for gamma-rays and curve C for protons). The LM effective area for gamma-rays is 2-3 orders of magnitude smaller than the ”single particle” one (curve A), due to the higher number of particles required to satisfy the trigger condition ($``$6 instead of 1). However the possibility to measure the primary arrival directions using the standard reconstruction technique of the shower front, reduces significantly the background, limited to cosmic rays with directions inside the angular error box. The angular resolution for primaries with energy $`E`$ 10 GeV (defined as the opening angle around the source containing the 70$`\%`$ of the signal showers) is $`r5^{}`$ (Abbrescia 1996). Using the cosmic ray primary proton spectrum measured by Honda (1995), the number of background events with arrival directions in a cone of radius $`r=5^{}`$ and zenith angle $`\theta `$=20 are expected to be $`B_{LM}`$ 160 s<sup>-1</sup>. The number of events in ARGO-YBJ due to the burst is: $`N_{LM}=0.7A_{eff}𝑑N_\gamma /𝑑E𝑑E`$. Requiring for the GRB signal a minimum statistical significance of 4 standard deviations, the number $`N_{LM}`$ of events in ARGO-YBJ due to a GRB of time duration $`\mathrm{\Delta }t`$ = 1 s and zenith angle $`\theta `$=20 must be larger than $``$50. ### 3.3 Sensitivity to detect GRBs For simplicity we assume a burst giving a gamma-ray flux at the top of the atmosphere as $`dN/dE=KE^\alpha `$ photons m<sup>-2</sup> and a power law energy spectrum extending with unchanged slope up to a maximum energy $`E_{max}`$, with $`E_{max}>`$10 GeV. This assumption is supported by EGRET observations, which report power law spectra extending with no visible cutoff up to the maximum energy determined by the instrument sensitivity (in some cases above 1 GeV). The average spectral slope observed in the 30 MeV-10 GeV region is $`\alpha =1.95\pm 0.25`$ (Dingus 1997). Obviously a sharp cutoff at $`E=E_{max}`$ is unrealistic, but for our purposes this simple parametrization can be adopted. The energy cutoff can be due to an intrinsic cutoff at the source or/and to the absorption of gamma-rays in the intergalactic space, as previously mentioned. The latter effect could affect the spectra at relatively low energy: according to Salomon and Stecker (1998), gamma-rays of energy larger than $``$ 40 (100) GeV would be strongly absorbed if the GRB distance is $`z`$=1.0 (0.5). In order to evaluate the ARGO-YBJ sensitivity, it is convenient to work in terms of $`F_{min}`$, defined as the minumum energy fluence in the energy range 1 GeV$`÷E_{max}`$ necessary to make a GRB observable by ARGO-YBJ, assuming that the spectrum extends with unchanged slope up to $`E_{max}`$. Fig. 2 shows $`F_{min}`$ as a function of $`E_{max}`$, with $`E_{max}`$ in the range 10 GeV$`÷`$1 TeV, using the SP and the LM techniques. The curves are given for a GRB duration $`\mathrm{\Delta }t`$=1 s and a spectral slope: $`\alpha `$ = 2.0. The minimum fluence for a different duration $`\mathrm{\Delta }t`$ scales as $`\sqrt{\mathrm{\Delta }t}`$. The minimum required statistical significance of the signal is $`\sigma =4`$ standard deviations. Obviously the sensitivity increases with $`E_{max}`$. The dependence on $`E_{max}`$ is stronger for LM than for SP, given the different behaviour of the effective areas as a function of the gamma-ray energy (moreover, the angular resolution improves with the energy, an effect which is not accounted for in the present calculations). This makes the SP technique more profitable when the energy spectrum has a relatively low energy cutoff, in this case, for $`E_{max}<`$50 GeV. This value ranges between 35 and 70 GeV if the slope $`\alpha `$ varies from 1.5 to 2.5. To compare the ARGO-YBJ sensitivity with the fluxes that can be reasonably expected at high energy, in the same figure we report the fluences in the 1$`÷`$100 GeV energy range obtained extrapolating (with the observed slopes) the spectra measured by EGRET during the 15 events detected by the TASC instrument (Catelli 1997b)(one of the events, showing an unusal steep spectrum with $`\alpha `$=3.67, is not shown, since the extrapolated fluence $`F=10^{10}`$ erg cm<sup>-2</sup> falls well out of the plot). As can be seen in the figure, most of the events have an energy fluence larger than the ARGO-YBJ limits. ## 4 Conclusions The ARGO-YBJ detector could observe GRBs in the energy range $`E>10`$ GeV using the ”single particle” technique (SP) and the ”low multiplicity” technique (LM). The SP method is more suitable for gamma-ray bursts with energy spectra not extending more than $``$ 50 GeV, while the LM method is preferable for more energetic spectra. Adopting both techniques, ARGO-YBJ could detect GRBs with energy fluence in the range 1$`÷`$100 GeV as low as $`F10^6÷10^5`$ erg cm<sup>-2</sup>, if the spectral slope is $`\alpha `$2. References Abbrescia M. et al., Proposal of the ARGO experiment, 1996 (can be downloaded at the URL: http://www1.na.infn.it/wsubnucl/cosm/argo/argo.html) Aglietta M. et al., 1999, Proc. Conf. GRBs in the afterglow era, in press Bacci C. et al., Addendum to the ARGO Proposal, 1998 (can be downloaded at the URL: http://www1.na.infn.it/wsubnucl/cosm/argo/argo.html) R.Cabrera et al., 1999, Proc. Conf. GRBs in the afterglow era, in press Catelli J.R., Dingus B.L. and Schneid E.J., 1997, 25<sup>th</sup> ICRC Proc, 3, 33 Catelli J.R., Dingus B.L. and Schneid E.J., 1997, AIP Conf.Proc. 428, 309 Dingus B.L., Catelli J.R. and Schneid E.J., 1997, 25<sup>th</sup> ICRC Proc, 3, 30 Djorgovski S.G. et al 1999, GCN Circular 289 Honda M et al., 1995, Phys.Rev.D 52, 4985 Salomon M.H. and Stecker F.W., 1998, ApJ 493, 547 Vernetto S., submitted to Astroparticle Phys, 1999, Astro-ph 9904324
no-problem/9906/hep-ph9906377.html
ar5iv
text
# The monopoles in the structure of the electron ## 1 Introduction The monopole idea as derived by Dirac brought a natural expalnation to the charge quantization in Electrodynamics. The existence of a single monopole in Nature could explain the discretization of all particles electric charges. After the initial idea the monopole concept faced some difficulties to be described through an action principle, . As the action principle for particles (where the monopole is considered as a source of magnetic field) and fields could not be properly defined, the quantization of charge in Electrodynamics as proposed by Dirac turned out to be problematic. In fact even in the Fiber Bundle formalism, , the problem persists because the definition of a monopole in Quantum Electrodynamics suffers of two theoretical problems: The lack of a general variational principle and the proper care of the singular Dirac strings. The work of Wu and Yang, , attempts to solve the second only. The objective of the present work is to propose an electromagnetic structure for the electron (or any other spin particle) as composed by an electric charge and two Dirac monopoles (these last two being of opposite magnetic charge). It is possible to show that an action principle can be assigned for this system, in which both the particles and fields can be derived. As a classical system, it can be shown also that, contrary to the primer classical investigations, the size of the electron’s structure must not to be of the order of the electron’s Compton wave lenght (in fact it may have a size as small as desidered). Another interesting consequence is that, being possible to describe a particle’s spin by monopoles in its structure, the quantization of charges proposed by Dirac holds in Quantum Electrodynamics. In view of recent contributions on the Dirac string problem, , it is interesting to consider the present work in this context, which it is addressed at the conclusion section. ## 2 The electron’s structure The point - electron is now formally defined as composed of an electric charge $`e`$ ($`e\mathrm{Re}`$) and two monopoles of opposite magnetic charge, $`g`$ and $`g`$ ($`g\mathrm{Re}`$). The $`g>0`$ monopole is defined as a source of magnetic field while the $`g<0`$ monopole as a sink for this field. By this definition the covariant equations for the electric and magnetic electron’s components are: $`_\mu F^{\mu \nu }`$ $`=`$ $`j_e^\nu `$ (1) $`_\mu F_D^{\mu \nu }`$ $`=`$ $`j_g^\nu `$ where $`F^{\mu \nu }`$ is the electromagnetic field tensor connected to the four - vector $`j_e^\nu `$ formed by the electric charge density and current, and $`F_D^{\mu \nu }`$ is the dual of $`F^{\mu \nu }`$ (which is connected to the four -vector $`j_g^\nu `$ formed by the magnetic charge density and current): $$F_D^{\mu \nu }=\frac{1}{2}\epsilon ^{\mu \nu \sigma \delta }F_{\sigma \delta }$$ (2) if $`\epsilon ^{\mu \nu \sigma \delta }`$ is the four - dimensional Levi - Civita tensor ($`\epsilon _{0123}=\epsilon ^{0123}=1`$, completely antisymmetric). Each charge and monopole can be described by a corresponding Lorentz’s equation, $`m_e{\displaystyle \frac{d^2z^\mu }{ds^2}}`$ $`=`$ $`e{\displaystyle \frac{dz^\nu }{ds}}F_{\mu \nu }(z)`$ (3) $`m_g{\displaystyle \frac{d^2x^\mu }{ds^2}}`$ $`=`$ $`g{\displaystyle \frac{dx^\nu }{ds}}F_{D\mu \nu }(x)`$ where $`m_{e,g}`$ is the (electric, magnetic) charge’s inertial mass, $`z_\mu (s)`$ is the charge’s world - line as function of the proper time $`s`$ and $`x_\mu (s)`$ the corresponding four - coordinates which account for the monopole’s position. It is important to define a valid local action principle for an electron which is described by charges and monopoles as well as for a system of two distinct electrons of the kind. As the electron has a half - integer spin (to be demonstrated later) and no net magnetic charge, all the problems related to the definition of an action principle are automatically solved . The fundamental question is the formal definition of the structure, i.e., of having a charge and monopoles at the same spacetime point by the present definition of the structure of the electron. The question about the possibility for a local action principle is based on the fundamental interaction between an electric charge and a magnetic monopole . Considering a system composed of just one charge and one monopole, the fields that appear in the equations of motion (3) represent radiative reaction terms and the fields due to their mutual interaction. In the case of the charge, the tensor field has a term due to the monopole influence, $`(1)\{\frac{1}{4}\epsilon ^{\mu \nu \rho \lambda }\epsilon _{\rho \lambda \sigma \delta }F^{\sigma \delta }\}`$, which is the dual of the dual field tensor times $`(1)`$, i.e., $`(1)(F^{\mu \nu })`$. The $`(1)`$ factor is to be multiplied by the dual of $`V^{\mu \nu }`$ (if now $`_\mu V^{\mu \nu }=j_g^\nu `$) when considering the action upon the electric charge. Omitting the self - reacting free field of the charge on itself, the equation of motion can be derived from the action principle on: $$m_e\left(\frac{z^\mu }{t}\frac{z_\mu }{t}\right)^{1/2}𝑑te\frac{z_\mu }{t}V_D^\mu (z)𝑑t$$ (4) where $`V_D^\mu `$ is the dual of a four - vector potential related to the magnetic density of charge and current, $$_\nu _\nu V^\mu =j_g^\mu .$$ (5) Once there are no longer homogeneous field equations, potentials can no longer be introduced by expressing the field tensor $`V^{\mu \nu }`$ as the curl $`V^{\mu \nu }=^\mu V^\nu ^\nu V^\mu `$. The only way it can be done is when $`V_D^\mu (z)`$ is never to be considered at the points where $`j_g^\mu 0`$ i.e., it must be assured the charge and the monopole never met at the same spacetime point. This requirement is not physical for the case of free monopoles, and therefore it is important to show this can be physically achieved in the case of the electron. In order to prove the second term in the action of equation (4) is physical in the case of the electron, it is necessary to consider first the corresponding nonlocal action as a consequence of $`j_g^\mu 0`$: $$V_D^\mu (x)=_{\mathrm{}}^0V_D^{\alpha \beta }(\xi )\frac{\xi _\beta }{x^\mu }\frac{\xi _\alpha }{\tau }𝑑\tau $$ (6) where the four- vector $`\xi ^\mu (x,\tau )`$ is characterized by $`\xi ^\mu (x,0)`$ $`=`$ $`x^\mu `$ (7) $`\underset{\tau \mathrm{}}{lim}\xi (x,\tau )`$ $`=`$ $`\mathrm{space}\mathrm{like}\mathrm{infinity}`$ The path from space-like infinity to $`x`$ is traversed by $`\xi `$ as $`\tau `$ varies from $`\mathrm{}`$ to $`0`$ and then, from equation (6) one finds: $$^\mu V_D^\nu ^\nu V_D^\mu =V_D^{\mu \nu }+_{\mathrm{}}^0\frac{\xi _\beta }{x^\nu }\frac{\xi _\gamma }{x^\mu }\frac{\xi _\alpha }{\tau }(^\gamma V_D^{\alpha \beta }+^\alpha V_D^{\beta \gamma }+^\beta V_D^{\gamma \alpha })𝑑\tau $$ (8) with $`(^\gamma V_D^{\alpha \beta }+^\alpha V_D^{\beta \gamma }+^\beta V_D^{\gamma \alpha })=\epsilon ^{\alpha \beta \gamma \sigma }\left[j_g\right]_\sigma (\xi )`$. The action will be nonlocal as long as, in order to get $`^\mu V_D^\nu ^\nu V_D^\mu =V_D^{\mu \nu }`$ from equation (8), the paths of the charge and monopole are previously set to not cross, so that a charge never meets a monopole at the same spacetime point. Besides the magnetic charge in the point - electron case is null (as it is composed of two monopoles of opposite sign plus an electric charge) it is important to define the conditions a monopole and a charge can be considered at the same spacetime point , i.e., it must be verified to be valid for the particular case of the electron’s structure. Considering the general case of a charge and a monopole that meet at some spacetime point, the contribution for $`F^{\mu \nu }`$ in equation (3) as given by the second term on the right of equation (8) will be : $$\underset{\tau 0}{lim}\left[\frac{\xi _\alpha }{\tau }\right]_{\tau =0}\epsilon ^{\alpha \mu \nu \sigma }_\tau ^0\left[j_g\right]_\sigma (\xi )e\frac{dz_\nu }{ds}𝑑\tau .$$ (9) This contribution will vanish provided $$\frac{dz_\nu }{ds}=\frac{dx_\nu }{ds}$$ (10) (according to the definitions in equation (3) $`j_g`$ is expressed in terms of $`dx_\nu /ds`$), as in this case $`\epsilon ^{\alpha \nu \mu \sigma }(dz_\sigma /ds)(dx_\nu /ds)=0`$. Therefore the local character of the action principle is assured as long as, when at the same spacetime point, the charge and the monopole have the same four - velocities. This of course, is the case for the electron’s structure, then the concept of having the electric and magnetic charges at the same spacetime point is well defined and leads to a valid local action principle for theparticles in the electron’s structure. In order to have a general variational principle from which both particle as well as field equations can be derived simultaneously for a system of distinct electrons, it is necessary to have $`j_g^\mu =0`$ everywhere . This is true for the case of the electron’s structure as it is composed of two monopoles of opposite sign at the same point with the charge: No net magnetic charge is defined and no free monopole is considered. As there is also no crossing of the world - lines of two distinct spin one - half electrons, a variational principle exists for the complete theory. In order to conclude the electron’s structure definition it is necessary to determine its spin. In classical or quantum approach the electron is supposed to have a nonzero radius in order the self - energy part due to the electromagnetic fields be finite. The stability and self - energy problems are not the concern in this work. The important issues are the facts the electron has a defined discrete electric charge and a quantized intrinsic magnetic flux. If some radius is to be assigned to the electron, where by the definition all the fields are set to be zero within the defined sphere of that radius, the charge and monopole densities inside the sphere are unknown. The important information is that these densities exist for some electron’s radius but no net magnetic flux on closed surfaces can be found since any closed surface must entirely contains the whole structure (the elementary particle has now a minimum radius). In principle, when $`g0`$ is assigned on some point, a variational principle cannot be defined in a way to derive the equations of motion for particles and fields simultaneously. In order to calculate the electron’s spin one considers the monopoles are not at the same position together with the charge, but arbitrarily close, then a general variational principle holds in this limit. The electron’s angular momentum can be easily calculated. It is expected by the Dirac equations that an irregular circulatory movement of the electron’s electric charge (Zitterbewegung) is the current responsable for the intrinsic magnetic moment. The two monopoles are disposed within the electron’s structure in a way that the intrinsic magnetic dipole is the combination of a positive and a negative monopole. This combination will be indistinguishable from that produced by the suitable current due to Zitterbewegung . The calculation now proceeds for the monopoles and charge system: In a Cartesian coordinate system, consider a positive monopole $`g`$ at $`(0,0,z_0)`$ with $`(z_0>0,z_0\mathrm{Re})`$, a negative monopole $`g`$ at $`(0,0,z_0)`$ and an electric charge $`e`$ at the origin. It is possible to calculate the angular momentum of the electromagnetic field, $`𝐋_{em}`$: $$𝐋_{em}=\frac{1}{4\pi c}𝐫\times \left(𝐄\times 𝐁\right)𝑑v$$ (11) where $`𝐫`$ is the three - vector distance, $`𝐄`$ and $`𝐁`$ the electric and magnetic fields respectivelly, and $`dv`$ the volume integration element. It results: $$𝐋_{em}=\frac{\left(2g\right)e}{c}\frac{𝐳}{𝐳}$$ (12) with $`𝐳/|𝐳`$ the versor on the positive direction on $`z`$ axis. It is independent of $`z_0`$, so the monopoles can be defined arbitrarily close to the charge for the same resulting angular momentum. Once the monopoles are defined inside the electron’s radius, a magnetic dipole field is generated and no isolated monopole is considered at all, as the combination of the two monopoles must account for the Zitterbewegung current. The magnetic flux is expected to be quantized for a discrete electron’s charge, due to Dirac strings or due to simple quantum - arguments ( the quantization of the angular momentum). The Dirac quantization condition , $`\mu e/c=nh/2`$ ($`n`$: integer), with $`\mu `$ the total magnetic flux, gives the expected quantized magnetic flux. Since the resulting angular momentum of the electric charge $`e`$ and the monopole $`g`$ is equal to that of the same electric charge $`e`$ and the $`g`$ monopole (as this last monopole is opposit directioned related to the first, regarding the central electric charge) the total angular momentum is as shown by equation (12). By Dirac quantization condition one then gets $`\left(2g\right)e/c=nh/2`$ for the system. The same result can be achieved if the electromagnetic field is quantized by half - integer numbers , so that the same relation can be derived directly from equation (12) with this assumption. ## 3 Comments and conclusions In the view of Dirac’s work the charge quantization can be now properly stated since there is in fact at least one monopole in Nature (each electron has at least two). It was shown that this structure has no problems to be described through a variational principle . It is interesting to mention the fact that if Dirac’s quantization condition $`\mu e/c=nh/2`$ is valid, the electron’s magnetic moment, $`e\mathrm{}/(2mc)`$ is $`2gz_0`$. As the angular momentum is $`\left(2g\right)e/c=nh/2`$, one gets: $$z_0=\frac{e^2}{2\pi mc^2}\left(\frac{1}{n}\right)$$ (13) where $`m`$ is the electron’s inertial mass and $`n`$ an integer number (which must be different from zero in the case of the spin electron). We get roughly, $$z_0\frac{10^{14}}{n}$$ (14) centimeters. As $`n`$ may have any integer value different from zero, the size of the electron as composed of one electric charge and two monopoles of opposite charge is arbitrarily small ($`n\mathrm{}`$) for the same value of the magnetic moment. Once it is possible to set $`z_00`$ the variational principle for particles and fields is well defined as for any action on the monopole $`+g`$, there is a $`g`$ monopole arbitrarily close. It is then interesting that in this semi - classical description the electron as composed of classical particles may have any small radius for some fixed known magnetic moment (in the usual classical approaches electric currents required some finite spacial dimension to be defined as source for the magnetic moment). It is important to mention the Dirac strings associated to the monopoles had an special attention in the past years as a separeted issue. Wu and Yang get rid of the singular string influences via a Fiber Bundle formalism in which it is not observable, , being considered only as a gauge artifact. He, Qiu and Tze argumented it is impossible to have $`g0`$ in pure Quantum Electrodynamics however. It was showed that the Dirac quantization condition is related to arbitrary gauge couplings in pure Quantum Electrodynamics. In view of the present work the conclusions in reference are very instructive, because, if the electron can be described as composed of monopoles in its structure, some physical significance to the strings must be supplied in Physics. ## 4 Acknowledgements I would like to thank professor H. -J. He (University of Michigan) for usefull comments. I am in debt to professors C. O. Escobar (Unicamp), R. Tschirhart (Fermilab), Y. W. Wah (University of Chicago) and R. Ray (Fermilab) for the opportunity and hospitality at Fermilab where this work was developed. I am also gratefull to Centro Universitário Salesiano de São Paulo Unisal for the finantial support. ## 5 References . P. A. M. Dirac, Proc. Roy. Soc. A133 (1931) 60 P. A. M. Dirac, Phys. Rev. 74 (1948) 817 F. Rohrlich, Phys. Rev. 150 (1966) 1104 D.Rosenbaum, Phys. Rev. 147 (1966) 891 M. N. Saha, Phys. Rev. 75 (1949) 1968 T. T. Wu and C. N. Yang, Phys. Rev. D12 (1975) 3845 H. -J. He, Z. Qiu and C. -H. Tze, Zeits. Phys. C65 (1995), hep-ph/9402293
no-problem/9906/astro-ph9906392.html
ar5iv
text
# X-ray luminous star-forming galaxies ## 1 INTRODUCTION A large number of faint (B$`<23`$) galaxies has been found in deep ROSAT surveys (eg Boyle et al. 1995, Georgantopoulos et al. 1996). These galaxies have high X-ray luminosities ($`L_x>10^{42}`$ $`\mathrm{erg}\mathrm{s}^1`$) and present narrow emission lines in their optical spectra. Despite the high probability of chance coincidences (ie field galaxies lying accidentally in the error box of the x-ray source), cross-correlations of the X-ray positions with deep optical images have proved the existence of this population at a high level of significance (eg Roche et al. 1995, Almaini et al. 1997). Schmidt et al. (1998) obtained Keck spectra for 50 X-ray sources in the Lockman hole. They find that the large majority of these host an AGN although some fraction of galaxies cannot be excluded. Now, we have the opportunity to study the properties of such luminous galaxies in the local Universe, using the ROSAT all-sky survey (RASS). Boller et al. (1992) cross-correlated the IRAS Point Source Catalogue with the RASS. Preferential spectroscopic follow-up observations of the highest X-ray luminosity objects (Moran et al. 1997) show that the vast majority of galaxies with $`L_x>10^{42}`$ $`\mathrm{erg}\mathrm{s}^1`$ are AGN. Here instead, we have carried out a cross-correlation of the sample of Ho et al. (1996, 1997) and the ROSAT All Sky Survey Bright Source Catalogue (RASS-BSC). The spectroscopic sample of Ho et al. contains moderate resolution, high-signal-to-noise spectra of all northern galaxies with B$`<`$12.5 providing a complete sample of the galactic activity in the nearby universe. The important advantage of the Ho et al. sample is the pre-existing, very good quality spectra which give us unambiguous classifications for all the galaxies. The RASS-BSC contains almost 18000 sources found in the all-sky survey carried out during the first years of the ROSAT mission. Our aim is to understand the X-ray emission mechanisms in the nearby galaxies and to test whether a class of X-ray luminous ($`L_x10^{4142}\mathrm{erg}\mathrm{s}^1`$) star-forming galaxies exists. ## 2 THE CROSS-CORRELATION The results of the cross-correlation are presented in table 1. There are 45 coincidences within 1 arcmin distance from the optical galaxy. On the basis of the sky density of the RASS-BSC sources, we expect less than 1 to be by chance. Columns 1, 2 and 3 contain the names and the coordinates of the objects; the X-ray luminosities calculated using the RASS-BSC count rates and a power-law of $`\mathrm{\Gamma }=2`$ are listed in column 4 (we use an $`H_o=65\mathrm{k}\mathrm{m}\mathrm{s}^1\mathrm{Mpc}^1`$); finally, in column 5 we give the spectroscopic classifications from Ho et al. (1997). We note that the sample is by no means statistically complete, due to the non-uniform coverage of the sky in the RASS. From table 1 we see that the large majority of the X-ray galaxies are AGN (Seyferts but also some LINERS). However, there are seven star-forming galaxies while there is also a large fraction (6 galaxies) of late type or normal galaxies. From the seven starforming galaxies of our sample two are well studied dwarf starforming galaxies ( NGC5204, NGC4449) with X-ray luminosities of $`10^{38}10^{39}\mathrm{ergs}^1`$ (Della Ceca, Griffiths & Heckman 1996). The archetypal star-forming galaxy M82 ($`7.5\times 10^{40}\mathrm{erg}\mathrm{s}^1`$) is also in our sample. An interesting finding is the presence of two star-forming galaxies with luminosities above $`10^{41}\mathrm{ergs}^1`$ reaching the luminosities of low luminosity AGN. A peculiar object is NGC5905 which although has a very high luminosity of $`1.5\times 10^{42}\mathrm{erg}\mathrm{s}^1`$ in the RASS, has shown a significant decline in X-ray flux in subsequent observations (Bade et al. 1996). However, its optical spectrum has not any signatures of AGN activity. ## 3 NGC3690 AND NGC3310 From the starburst galaxies in this sample we present here results on the two most luminous ones, NGC3310 and NGC3690. Both are nearby galaxies at distances of 19.6 Mpc and 63.2 Mpc respectively. Both galaxies appear to be in interacting pairs/mergers. NGC3690 forms an interacting pair with IC694. Their separation is 21” which translates to $`6`$ kpc at the assumed distance. For NGC3310 there is also strong evidence that it is the remnant of a recent merger, according to anomalies found in its rotation curve (Mulder et al. 1985), and its disturbed morphology (Ballick and Heckman, 1981). Evolutionary synthesis modelling of the optical and infrared spectra of these galaxies has shown that the age of the starburst is about 10Myr (Pastoriza et al. , 1993 and Nakagawa et al. , 1989 for NGC3310 and NGC3690 respectively). Especially in the case of the latter Nakagawa et al. find that the two bursts have different properties, implying either different ages or different Initial Mass Functions (IMF). ### 3.1 X-ray data analysis. We have obtained the ROSAT (PSPC and HRI) and ASCA observations of these galaxies from the archive. After following the standard screening procedure, we extracted PSPC and ASCA SIS and GIS spectra along with HRI images. The HRI images show that the soft X-ray emission is extended in NGC3310. In NGC3690 the emission comes from three distinct components (the two correspond to the nuclei of NGC3690 and IC694) which again appear to be spatially resolved. We fitted the ROSAT and ASCA spectra together. The spectral fitting results are presented in table 2. We found that they are fitted with a optically thin thermal plasma of temperature $`0.8`$keV and either a hot thermal plasma (kT $`1015`$keV), or a power-law with $`\mathrm{\Gamma }1.51.6`$. The spectral fits for the soft band are suggestive for a thermal origin of the X-ray emission, arising from diffuse hot gas, probably associated with a galactic super-wind (Heckman et al. 1996). However, the origin of the hard X-rays is still unclear as there are more than one possible mechanisms which can produce the observed spectrum. A power-law spectrum can be produced either by X-ray binaries or Inverse Compton scattering of the starburst infrared photons by the supernova generated relativistic electrons, while hot gas (and X-ray binaries) give a thermal plasma spectrum. These results are similar to the results found for the prototypical starburst galaxies M82 and NGC253 (Ptak et al. 1997, Moran and Lehnert 1997), suggesting a common X-ray emission mechanism in star-forming galaxies spanning a wide range of luminosities. Better signal-to-noise ratio and higher energy X-ray spectra but mainly high resolution X-ray imaging are needed in order to draw any conclusions on the origin of the hard X-ray emission in these galaxies. ## 4 CONCLUSIONS We have presented our results on the cross-correlation of the sample of Ho et al. (1995) with the RASS-BSC. Our intent was to search for X-ray luminous star-forming galaxies and to probe the X-ray content of the nearby universe. The cross-correlation gives 45 objects within a radius of 1 arcmin. Although this sample is dominated by AGN (mainly Seyferts), there are seven star-forming galaxies spanning a large range of X-ray luminosities ($`10^{38}`$-$`10^{42}\mathrm{ergs}^1`$) rivaling the luminosities of low luminosity AGN. We have analyzed archival observations of the two most luminous star-forming galaxies in our sample namely NGC3690 and NGC3310. We find that their soft spectra are fitted by an optically thin thermal plasma of temperature $`0.8\mathrm{keV}`$. The hard X-ray emission can be fit either with a high temperature thermal plasma ($`\mathrm{kT}1015`$ keV) or a flat power-law ($`\mathrm{\Gamma }1.6`$). These results are similar to those found for the prototypical starburst galaxies M82 and NGC253, suggesting a common X-ray emission mechanism in star-forming galaxies over a large range of luminosities. The combination of AXAF and XMM observations will shed more light on the origin of the hard X-ray emission which currently remains unknown in all star-forming galaxies. ## References Almaini, O., Shanks, T., Griffiths, R.E., Boyle, B.J., Roche, N., Georgantopoulos, I., Stewart, G.C., 1997, MNRAS, 291, 372 Bade N. Komossa S. and Dahlem M., 1996, A&A, 309,35L Ballick B. and Heckman T., 1981, A&A, 96, 271 Boller et al. , 1992, A&A, 261, 57 Boyle, B.J., McMahon, R., Wilkes, B.J., Elvis. M., 1995, MNRAS, 276, 315 Della Ceca R., Griffiths R.E., Heckman T.M., 1997, ApJ, 485, 581 Heckman T.M. et al. , 1996, in The enviroment and Evolution of Galaxies, edited by J.M. Shull and H.A. Thronson, Jr. (Kluwer, Dordrecht) Ho L.C., Filippenko A.V. and Sargent W.L., 1995, ApJS, 98, 477 Ho L.C., Filippenko A.V., and Sargent W.L., 1997, ApJS, 112, 315 Georgantopoulos, I. et al. , 1996, MNRAS, 280, 276 Moran E.C. and Lehnert L.D., 1997, ApJ, 478, 172 Moran E.C., Halpern L.P., Helphand D.J., 1996, ApJS, 106,341 Mulder P.S. and Van Driel W., 1996, 1986, A&A, 309,403 Nakagawa T., et al. ., 1989, ApJ, 340, 729 Pastoriza et al. , 1993, MNRAS, 260, 177 Ptak et al. , 1997, AJ, 113, 1286 Roche N. et al. , 1995, MNRAS, 273L, 15 Schmidt, M. et al. 1998, A&A, 329, 495
no-problem/9906/astro-ph9906261.html
ar5iv
text
# May 1999 SINP/TNP/99-19 Particle dark matter : an overviewInvited talk at the conference “Cosmology: Theory confronts observations”, held at IIT Kharagpur, January 1999. ## 1 Light neutrinos Neutrinos interact very feebly, only through weak interactions. Despite this fact, they were in thermal equilibrium in the very early universe because the density of particles was much higher at that time, and so the mean free path (or the average reaction time) was small. At that stage, their number density differed from that of the photons only because one is a fermion and the other is a boson, and the relation was $`n_\nu ={\displaystyle \frac{3}{4}}n_\gamma .`$ (1.1) This age of thermal equilibrium ended once the density of the particles became small enough in an expanding universe, and it happened when the temperature was about $`1`$ MeV. After that, the neutrinos followed only the overall expansion. The photon number density also fell by the cube of the scale factor, so the ratio remained the same. The only exception to this statement occurred when the temperature dropped somewhat below the electron mass, and $`e^+e^{}`$ annihilations produced new photons. This event increased the photon number density by a factor of $`11/4`$, so that in the present universe $`\left(n_\nu \right)_0={\displaystyle \frac{3}{4}}{\displaystyle \frac{4}{11}}\left(n_\gamma \right)_0110\mathrm{cm}^3.`$ (1.2) If the neutrinos are massive, the energy density is obtained by multiplying this by the mass. The total energy density of the universe is parametrized by $`10^4h^2\mathrm{\Omega }\mathrm{eV}/\mathrm{cm}^3`$, where $`h`$ is the Hubble parameter in units of $`100\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1`$. Thus, the fraction of energy density contributed by light neutrinos would be $`F_\nu ={\displaystyle \frac{m_\nu n_\nu }{10^4h^2\mathrm{\Omega }\mathrm{eV}/\mathrm{cm}^3}}=\left({\displaystyle \frac{m_\nu }{92\mathrm{eV}}}\right){\displaystyle \frac{1}{h^2\mathrm{\Omega }}}.`$ (1.3) Atmospheric neutrino data indicate neutrino masses of order $`10^1`$ eV. Solar neutrino data indicate even smaller masses. Although these do not apply for all neutrino species, it surely is suggestive of the fact that $`F_\nu 1`$. Before discarding light neutrinos as dark matter candidates on this ground, I would like to emphasize that there is a very strong assumption implicit in this entire argument, viz. that the neutrinos have no chemical potential so that their number is the same as that of the antineutrinos. If chemical potential is nonzero, the number density of neutrinos depend on it as well as the temperature, and one cannot even start with the simple relation of Eq. (1.1). If the neutrinos have chemical potential, they can contribute a larger fraction to the energy density of the universe. This issue has been analyzed recently . We show the results in Fig. 1. The horizontal axis corresponds to neutrino mass. The vertical axis is $`n_\nu n_{\overline{\nu }}`$ in units of the CMBR $`n_\gamma `$ in the present universe. The lines in this plot correspond to the values $`h^2\mathrm{\Omega }F_\nu =\frac{1}{8},\frac{1}{4}`$ and $`\frac{1}{2}`$ respectively, starting from the inner line. ## 2 Axions So far we talked about light neutrinos. They are known to exist. The other candidates to be discussed have not yet been detected in any experiment. Thus, I need some motivation for talking about them. For the case at hand, the motivation comes from strong interactions. These interactions are believed to be described by the gauge theory called QCD. In a gauge theory, one decides on some symmetry of the Lagrangian and writes down all terms that are cosistent with this symmetry. For strong interactions, this symmetry was found in the 1970s. If one cares to do only perturbative calculations with the Lagrangian obtained from this symmetry, there is no problem. If, however, one includes non-perturbative effects, one encounters one term which violates parity and time-reversal symmetries: $`_{\mathrm{QCD}}=_{\mathrm{pert}}+{\displaystyle \frac{\theta g^2}{32\pi ^2}}G_{\mu \nu }\stackrel{~}{G}_{\mu \nu },`$ (2.1) where $`G_{\mu \nu }`$ is the generalization of the electromagnetic field-strength tensor for the relevant symmetry. Here $`g`$ is the gauge coupling constant, and $`\theta `$ is an arbitrary parameter. The term involving this parameter is the generalization of a term $`\stackrel{}{E}\stackrel{}{B}`$, which violates parity and time-reversal symmetries. Among other things, this term will give rise to an electric dipole moment for the neutron. The experimental bounds are very strong on this, which restricts this parameter severely: $`\theta <10^{10}`$. Very small parameters always present certain conceptual difficulties in quantum field theories. If they are not zero, one has to justify them by some symmetry. Peccei and Quinn discovered such a symmetry by introducing a Higgs field $`\widehat{a}(x)`$ which has an interaction $`{\displaystyle \frac{g^2}{32\pi ^2}}{\displaystyle \frac{\widehat{a}(x)}{f_{\mathrm{PQ}}}}G_{\mu \nu }\stackrel{~}{G}_{\mu \nu },`$ (2.2) where $`f_{\mathrm{PQ}}`$ has the dimensions of mass. The field, of course, has other interactions, including a potential of its own, which is minimum at $`\widehat{a}=\theta f_{\mathrm{PQ}}`$. In general, now, we can write $`\widehat{a}(x)=\theta f_{\mathrm{PQ}}+a(x)`$. This field $`a(x)`$ vanishes at the minimum and is qualified to be a quantum field. It is called the axion field. Notice how it has solved our original problem. If we add the terms shown in Eqs. (2.1) and (2.2) and use the definition of the quantum field, we see that the parameter $`\theta `$ cancels out. Said another way, $`\theta `$ is forced to equal to zero by this procedure. We have an interaction of the axion obtained by replacing $`\widehat{a}`$ by $`a`$ in Eq. (2.2), but that does not violate parity and time reversal provided the axion is intrinsically negative under these symmetries. We now examine how the axion can be important as the dark matter. The equation of motion of the axion field, neglecting its interactions, is $`\ddot{a}+3H\dot{a}+m_a^2a=0,`$ (2.3) assuming the field to be spatially homogeneous. This is similar to the Klein-Gordon equation, but there are two differences. One of them is the second term on the left side, which shows the effects of the expansion of the universe. The other is hidden in the notation, and is the fact that the ‘mass’ $`m_a`$ appearing in the equation is not a constant. It arises due to co-operative effects, and depends on the temperature. At early times, when the temperature was very high, $`m_a`$ was zero. This is because all non-perturbative effects involve a factor $`\mathrm{exp}(1/g^2)`$, which is negligible since at high energies $`g`$ is very small for QCD. In this stage, the solution of Eq. (2.3) is given by $`a=\mathrm{constant}`$. As the universe cools down, $`g`$ increases, and becomes large enough to make the non-perturbative effects important when the temperature drops down sufficiently, say below a value $`T_{\mathrm{QCD}}`$. To find the solution after this time, let us first forget about $`H`$ and also the time-dependence of $`m_a`$. In that case, the solution would be $`a=A\mathrm{cos}m_at,`$ (2.4) where $`A`$ is a constant. If $`Hm_a`$ and also the time variation of $`m_a`$ is small, we can still try a solution like this, where now both $`A`$ and $`m_a`$ should be regarded as functions of time. Substituting the solution in Eq. (2.3), we now obtain $`{\displaystyle \frac{d}{dt}}(m_aA^2)=3Hm_aA^2,`$ (2.5) or $`m_aA^2R^3=\mathrm{constant},`$ (2.6) where $`R`$ is the scale factor of the universe, i.e., $`H=\dot{R}/R`$. The energy density in these oscillations, at the present time, would be given by $`\rho _0={\displaystyle \frac{1}{2}}\left(m_a^2A^2\right)_0.`$ (2.7) The axion mass in the present universe is determined by $`f_{\mathrm{PQ}}`$. It depends mildly on the axion models, and is given by $`(m_a)_0f_{\mathrm{PQ}}m_\pi f_\pi ,`$ (2.8) where the quantities on the right sides relate to properties of the pions, and are very well-known. However, we don’t know the magnitude of the oscillations in the present universe. Therefore, we use Eq. (2.6) to write $`\rho _0={\displaystyle \frac{1}{2}}\left(m_a\right)_0\left(m_aA^2\right)_T\times (R^3/R_0^3),`$ (2.9) where $`R`$ is the value of the scale factor at any arbitrary temperature $`T`$. The oscillations started, as we said, around $`TT_{\mathrm{QCD}}100`$ MeV. At that time, $`m_aHT^2/M_P`$, and $`Af_{\mathrm{PQ}}`$. Using these estimates and utilizing the fact that $`R/R_0=T_0/T`$, we obtain $`\rho _0{\displaystyle \frac{1}{2}}{\displaystyle \frac{m_\pi f_\pi }{f_{\mathrm{PQ}}}}\left({\displaystyle \frac{T_{\mathrm{QCD}}^2f_{\mathrm{PQ}}^2}{M_P}}\right)\left({\displaystyle \frac{T_0}{T_{\mathrm{QCD}}}}\right)^3,`$ (2.10) which gives $`\mathrm{\Omega }_a10^{12}{\displaystyle \frac{f_{\mathrm{PQ}}}{1\mathrm{GeV}}}.`$ (2.11) Thus, axions can be an important source of dark matter if $`f_{\mathrm{PQ}}`$ is close to $`10^{12}`$ GeV. Let us now see what is experimentally known about these axions. In Fig. 2, we show the present status . The horizontal axis is the axion mass. The vertical axis denotes its coupling with two photons, which allows the decay mode $`a\gamma \gamma `$. As I said earlier, there is slight model dependence in these quantities. The predictions of two compelling models are shown in the graph by the lines marked DFSZ and KSVZ. If the axion mass is in the eV range, the decay photons could be detected through telescopes, and the failure to do so has been indicated in the figure. However, if $`f_{\mathrm{PQ}}`$ indeed equals $`10^{12}`$ GeV and we take Eq. (2.8) as a strict equality, the axion mass comes out to be $`10^5`$ eV. So, if the axions has to contribute significantly to the energy density of the universe, we should find them around that value of mass. In this case, the two photon decay mode is so slow that the axion can be considered stable on cosmological time scales. However, because of the coupling, the axions may be resonantly converted into a microwave signal in a cavity where a strong magnetic field is present. These experiments are very promising. As we see, they are already probing very close to the interesting region. ## 3 WIMPs We now come to the WIMPs, which is an acronym for “Weakly Interaction Massive Particles.” There are a variety of ways in which such particles may be motivated. The most recent and most compelling of them is supersymmetry. Supersymmetric models have new particles — bosonic partners of known fermions and fermionic partners of known bosons. There are phenomenolgically strong reasons to suspect that these new particles, or “superpartners”, can be produced or annihilated only in pairs. If that is true, the lightest of these superpartners will be a stable particle. It will not decay spontaneously. These can be examples of WIMPs. There are other models also which predict WIMPs. The number density of these WIMPs in the early universe is governed by the Boltzmann equation. In the homogeneous and isotropic universe, this takes the form $`{\displaystyle \frac{dn}{dt}}=3Hn\sigma v(n^2n_{\mathrm{eq}}^2),`$ (3.1) where $`n_{\mathrm{eq}}`$ is the equilibrium number density for the temperature at any given time. The decrease of number density due to the expansion of the universe is represented in the first term in this equation, and the same due to pair annihilations goes in the second term. It is convenient to switch to dimensionless variables defined by $`y={\displaystyle \frac{n}{s}},x=m/T,`$ (3.2) where $`m`$ is the mass of the WIMP and $`s`$ is the entropy of the background photons at a temperature $`T`$. The evolution equation now becomes $`{\displaystyle \frac{dy}{dx}}={\displaystyle \frac{y_{\mathrm{eq}}}{x}}{\displaystyle \frac{\mathrm{\Gamma }_{\mathrm{ann}}}{H(x)}}\left[\left({\displaystyle \frac{y}{y_{\mathrm{eq}}}}\right)^21\right],`$ (3.3) where $`\mathrm{\Gamma }_{\mathrm{ann}}=n_{\mathrm{eq}}\sigma vm^3x^{3/2}e^x\sigma v`$, putting in the equilibrium number density for a non-relativistic particle. The important thing to notice here is that $`y`$ hardly changes if $`\mathrm{\Gamma }_{\mathrm{ann}}H(x)`$ at any given era characterized by a temperature determined by $`x`$. Using $`H(x)T^2/M_P=m^2/M_Px^2`$, we find $`{\displaystyle \frac{\mathrm{\Gamma }_{\mathrm{ann}}}{H(x)}}mM_Px^{1/2}e^x\sigma v.`$ (3.4) As $`x\mathrm{}`$, i.e., for large time, this ratio goes to zero. It means that after a while, the pair annihilation rate will be negligible, because the universe has meanwhile become large enough so that the particles do not find each other. The remnant density of the WIMPs can be obtained by numerical integration of Eq. (3.3), but a rough estimate can be obtained by the density at the era when $`\mathrm{\Gamma }_{\mathrm{ann}}=H(x)`$. The solution of $`x`$ Eq. (3.4) is now determined once the mass and the annihilation cross section are known. After this era, the number density just scales inversely as the volume of the universe. Let me now turn to the experimental searches . These are summarized it in Fig. 3. As I just said, the parameters which determine the density of WIMPs in the present universe are the mass and the annihilation cross section. The two axes in this plot are precisely these two parameters. For the cross section, only the cross section with the nucleons is probed in the experiments. The solid lines are exclusions plots from already published data, and the dotted lines are projections for future experiments. The scatter plots are expected values of the parameters for various choices of the basic parameters in the supersymmetric model. It does seem that the experiments in the near future will be able to decide whether WIMPs really dominate the energy density of the universe.
no-problem/9906/cond-mat9906301.html
ar5iv
text
# Spinodal decomposition of an ABv model alloy: Patterns at unstable surfaces ## I Introduction The phase separation of alloys and other unstable mixtures (binary fluids, glasses, polymer blends) has been extensively studied for decades . In a typical experiment, a stable mixture is quenched into an unstable state by a sudden change in an external control parameter, mostly temperature or pressure. Then, domains of the new equilibrium phases develop and coarsen, leading to a heterogeneous material. As the mechanical and transport properties of an alloy may depend considerably on its microstructure, the understanding of the dynamics of domain formation and growth is of great practical importance. In addition, phase separation is an interesting example of a process where spontaneous pattern formation occurs during the approach to equilibrium. In recent years, attention has been drawn to surface effects on spinodal decomposition, mainly due to interesting experimental results on polymer blends . Thin films of mixture are placed between glass plates, or between a substrate and vacuum, and quenched. If the interactions with the surface favor one of the components, this component rapidly segregates to the surface and triggers phase separation in the bulk: one observes surface-directed spinodal decomposition, during which concentration waves propagate from the surface into the sample. In some of the experiments, modulations parallel to the surface are observed, leading to a pattern along the surface with a distinct length scale . The surface effects are in competition with bulk spinodal decomposition, and only a thin layer is affected by the surface, whereas in the interior of the sample the usual bulk structures are found. Interesting questions are then what determines the type of structures found at the surface, and to predict typical length scales and growth rates of the patterns, as well as the thickness of the surface layer. The dynamics of spinodal decomposition near planar substrates has been investigated using continuum models , Monte Carlo simulations , and mean-field kinetic equations . On the other hand, very little is known about free surfaces, which can deform during the decomposition process . We have recently found that surface modes and spinodal waves at free surfaces can be observed in a simple lattice model , and the present paper is devoted to a detailed study of this phenomenon. The classical theory for the early stage of spinodal decomposition, based on out-of-equilibrium thermodynamics, was proposed by Cahn and Hilliard . The Cahn-Hilliard equation is obtained by postulating that the local interdiffusion currents are proportional to the gradients of the local chemical potential, and supposing that this chemical potential can be derived from a free energy functional of Ginzburg-Landau type. The proportionality constant is the atomic mobility, which is a phenomenological parameter in this theory. For a homogeneous initial state, the Cahn-Hilliard equation may be linearized around the average composition, and analyzed in terms of Fourier modes. Long wavelength perturbations grow exponentially with time, whereas short wavelength fluctuations are damped by the gradient terms, related to the surface tension. A typical length scale of the resulting domain pattern is then given by the wavelength of the fastest growing mode. It has been recognized that simplified lattice models are valuable tools to study the influence of microscopic dynamics on the process of phase separation in metallic alloys. They provide a convenient conceptual framework with a minimum number of parameters. Such lattice gas models assume the existence of a fixed crystal lattice, the sites of which can be occupied by the atoms of different species, or by vacancies. A configuration evolves by atomic jumps from site to site. Many studies have been carried out on the kinetic Ising model with Kawasaki spin exchange dynamics , which corresponds to the direct exchange of atoms. In most alloys, however, the dominant mechanism of diffusion is a vacancy mechanism , and several models of vacancy-mediated spinodal decomposition have been investigated . In our model, the rates for atomic jumps follow an Arrhenius law. A standard procedure to investigate the dynamics of such models are Monte Carlo simulations . The advantage of this approach is that the simulations constitute a genuine realization of the stochastic model, and all correlations in space and time are preserved. But for the same reason, an analytic understanding is difficult. Therefore, there have been numerous attempts to formulate microscopic kinetic equations which are analytically tractable. These equations are derived from the microscopic master equation by an approximation of mean-field type. While the linearized equation of Khatchaturyan is valid only near equilibrium, recently several authors have proposed fully nonlinear kinetic equations for Ising and lattice gas models . These equations can be cast in the form of generalized Cahn-Hilliard equations, where the atomic mobility depends explicitly on the details of the microscopic dynamics. Such equations have been applied successfully to the study of phase-separation dynamics in binary and ternary systems , and to dendritic growth . In particular, kinetic equations for models with vacancy dynamics have been proposed by Chen and Geng and by Puri and Sharma . We will follow the approach developed by one of the present authors , which allows to relate the microscopic equations in a particularly straightforward manner to the equations of out-of equilibrium thermodynamics. All the studies of vacancy-mediated phase separation we are aware of start from a homogeneous initial state. We study here the evolution of droplets of a dense mixture with few vacancies, immersed in a bath of vacancy-rich “vapor” phase. The interesting point in this situation is that we can have interfaces between an unstable mixture and a stable vapor which are “neutral”, that is without segregation of one of the components. Such interfaces are impossible in a binary model like the Ising model, where any stable phase below the critical temperature favors one of the components. In our model, the atoms situated near the surface of the droplets have a larger mobility than bulk atoms. Therefore, phase separation is fastest at the surface. We observe different structures at the surface, depending on the model parameters and the initial composition of the mixture. In the specially symmetric case of equal interaction energies and concentrations for both components, we obtain a regularly modulated surface mode which generates an ordered pattern in a surface layer. For asymmetrical interaction energies or compositions, surface-directed spinodal waves are observed. This difference can be explained by a competition between surface spinodal decomposition and surface segregation. We determine the characteristic length scales and growth rates of bulk and surface modes by a linear stability analysis of the mean-field kinetic equations. Our approach allows thus to relate morphological parameters to the interaction parameters of the microscopic model. The remainder of this article is organized as follows: in Sec. 2, we present the model and the derivation of the mean-field kinetic equations. Sec. 3 describes the simulations, which are analyzed in Sec. 4: a linear stability analysis is performed in the bulk and at the surface. Sec. 5 presents the discussion of the results. ## II Model and mean field kinetic equations We consider a simple cubic lattice in $`d`$ dimensions, with lattice constant $`a`$ and a total number of $`N`$ sites. The sites can be occupied by atoms of two species, $`A`$ and $`B`$, or by vacancies $`v`$. The occupation numbers at site $`i`$, $`n_i^\alpha `$, where $`\alpha `$ denotes $`A`$, $`B`$, or $`v`$, are equal to $`1`$ if site $`i`$ is occupied by species $`\alpha `$, and zero otherwise. Double occupancy is forbidden, which gives the constraint $`n_i^A+n_i^B+n_i^v=1i`$. Hence there are two independent variables per site; we choose in the following $`n_i^A`$ and $`n_i^B`$. We consider only interactions between nearest neighbor atoms: the energy of a configuration $`𝒞=\{n_i^A,n_i^B,i=1\mathrm{}N\}`$ is given by the Hamiltonian $$H(𝒞)=\underset{i,j}{}\left[ϵ_{AA}n_i^An_j^A+ϵ_{BB}n_i^Bn_j^Bϵ_{AB}(n_i^An_j^B+n_i^Bn_j^A)\right],$$ (1) where the sum is over all nearest neighbor pairs $`i,j`$, and $`ϵ_{\alpha \beta }`$ are the interaction energies between two atoms occupying two nearest neighbor sites. Vacancies do not interact with atoms or other vacancies. Unlike in a binary model, where the phase diagram is completely determined by the exchange energy $`ϵ_{AA}+ϵ_{BB}2ϵ_{AB}`$, here the interaction energies are independent. One of them sets the temperature scale, the other two are free parameters. It can be shown that Eq. (1) is equivalent to the Hamiltonian of a general ternary system (see appendix). The configuration evolves by jumps of atoms to a neighboring vacant site. This is a common diffusion mechanism in metals, and the associated energy barrier is usually much lower than for a direct interchange of atoms. We therefore will completely neglect the latter process. As appropriate for activated processes, the hopping rates are assumed to follow an Arrhenius law. The underlying physical picture is that the atoms are trapped in potential wells located around the sites of the lattice. Atoms spend most of their time near a lattice site, but from time to time they jump over the energy barrier between neighboring sites, the necessary energy being provided by the lattice phonons. We assume two contributions to the barriers: a constant activation energy $`U_\alpha `$ and the local binding energy, $`H/n_i^\alpha `$, which depends on the local configuration. The jump rate from site $`i`$ to $`j`$ is $$w_{ij}^\alpha =\nu _0^\alpha \mathrm{exp}\left[\frac{1}{kT}\left(U_\alpha +\frac{H}{n_i^\alpha }\right)\right].$$ (2) Here, $`T`$ is the temperature (constant throughout the system) and $`k`$ is Boltzmann’s constant. The prefactor $`\nu _0^\alpha `$ is related to a vibration frequency of the atom in the well (at least in the absolute Eyring regime). We can absorb the constant $`U_\alpha `$ in a prefactor which sets the time scale, $`w_0^\alpha =\nu _0^\alpha \mathrm{exp}(U_\alpha /kT)`$. In principle, $`w_0^A`$ and $`w_0^B`$ are different; however, simulation results indicate that qualitative results are unaffected by the ratio $`w_0^A/w_0^B`$ as long as it is not too far from unity . Therefore, we will take in the following $`w_0^A=w_0^B=w_0`$. The jump rates become $$w_{ij}^\alpha =w_0\mathrm{exp}\left[\frac{ϵ_{\alpha A}}{kT}\underset{a}{}n_{i+a}^A\frac{ϵ_{\alpha B}}{kT}\underset{a}{}n_{i+a}^B\right].$$ (3) Here and in the following, summation over $`a`$ means a sum over all nearest-neighbor sites. The jump rates in Eq. (3) define a stochastic process. To describe its time evolution, one may use the master equation for the probability distribution $`𝒫(𝒞,t)`$ of finding configuration $`𝒞`$ at time $`t`$: $$\frac{𝒫(𝒞,t)}{t}=\underset{𝒞^{}}{}\left(𝒫(𝒞^{},t)W(𝒞^{},𝒞)𝒫(𝒞,t)W(𝒞,𝒞^{})\right).$$ (4) The transition rates $`W(𝒞,𝒞^{})`$ from configuration $`𝒞`$ to $`𝒞^{}`$ are equal to $`w_{ij}^\alpha `$ given by Eq. (3) if the two configurations differ only by an exchange of a particle and a vacancy, and zero otherwise. Given a solution to the master equation, one can formally define a time-dependent average for any operator $`O(\{n_i^\alpha \})`$, function of the occupation numbers, by $$O(t)=\underset{𝒞}{}O(\{n_i^\alpha \})𝒫(𝒞,t).$$ (5) In particular, we define the time-dependent occupation probabilities of the sites: $$p_i^\alpha (t)=n_i^\alpha (t).$$ (6) These probabilities can also be interpreted as local concentrations. To obtain a kinetic equation for $`p_i^\alpha (t)`$, one differentiates both sides of the above equation with respect to time and uses the master equation. The result is a conservation law for the local occupation probability, $$\frac{p_i^\alpha }{t}=\underset{a}{}j_{ii+a}^\alpha .$$ (7) The currents $`j_{ij}^\alpha `$ through the link $`i,j`$ are given by $$j_{ij}^\alpha =n_i^\alpha (1n_j^An_j^B)w_{ij}^\alpha n_j^\alpha (1n_i^An_i^B)w_{ji}^\alpha .$$ (8) The prefactors of the jump rates assure that the start site is occupied by an $`\alpha `$-atom and the target site is empty. Up to now, the kinetic equation is equivalent to the complete master equation. To obtain a closed system of equations for the occupation probabilities $`p_i^\alpha `$, we make a mean-field approximation and replace the occupation numbers in the expressions for the currents by their averages $`p_i^\alpha `$. Clearly, this approximation is drastic: the resulting equations are a set of coupled deterministic differential equations. Fluctuations are suppressed, and the mean-field approximation does not take into account correlations. Nevertheless, as in the case of static mean-field approximations, one can use this method to get qualitative results on the dynamics of microscopic models. Similar approximations have been discussed by several authors . A possibility to improve systematically the simple mean-field equations is the use of the path probability method (PPM) devised by Kikuchi . The equations of the PPM, however, are considerably more complicated, and simulations on bulk spinodal decomposition have shown that the use of the PPM in the pair approximation leads to the same qualitative conclusions as the simple mean-field approximation . Therefore, we will study in the following only the mean-field case. We generalize to our ternary model the method developed for binary systems in Ref. . The equations for the currents can be written as a product of a prefactor $`S_{ij}`$, symmetric with respect to the interchange of $`i`$ and $`j`$, and the difference of two local terms $`C_j`$ and $`C_i`$, equivalent to chemical activities: $$j_{ij}^\alpha =S_{ij}^\alpha \left(C_j^\alpha C_i^\alpha \right),$$ (9) with $$S_{ij}^\alpha =w_0(1p_i^Ap_i^B)(1p_j^Ap_j^B)$$ (10) and $$C_i^\alpha =\frac{p_i^\alpha }{1p_i^Ap_i^B}\mathrm{exp}\left[\frac{ϵ_{\alpha A}}{kT}\underset{a}{}p_{i+a}^A\frac{ϵ_{\alpha B}}{kT}\underset{a}{}p_{i+a}^B\right].$$ (11) This factorization is not unique (see Ref. for more details). Our particular choice makes it straightforward to establish a connection to the phenomenological equations of out-of-equilibrium thermodynamics. We define a local chemical potential by $$\mu _i^\alpha =kT\mathrm{ln}C_i^\alpha .$$ (12) Then, the current becomes $$j_{ij}^\alpha =M_{ij}^\alpha \left(\mu _j^\alpha \mu _i^\alpha \right),$$ (13) where the mobility in the link $`ij`$ is given by $$M_{ij}^\alpha =S_{ij}^\alpha \frac{C_j^\alpha C_i^\alpha }{\mu _j^\alpha \mu _i^\alpha }.$$ (14) The explicit expression for the chemical potentials is $$\mu _i^\alpha =ϵ_{\alpha A}\mathrm{\Delta }_ap_i^Aϵ_{\alpha B}\mathrm{\Delta }_ap_i^Bzϵ_{\alpha A}p_i^Azϵ_{\alpha B}p_i^B+kT\mathrm{ln}\frac{p_i^\alpha }{1p_i^Ap_i^B},$$ (15) where $`\mathrm{\Delta }_ag_i=_a(g_{i+a}g_i)`$ for any site-dependent quantity $`g`$ is the discrete Laplacian, and $`z`$ is the coordination number of the lattice. This expression can also be obtained as the derivative of a free energy function $`F`$ with respect to the local occupation, $$\mu _i^\alpha =F/p_i^\alpha .$$ (16) The free energy function is a discrete analog of a Ginzburg-Landau functional: $$F=\underset{i}{}f(p_i^A,p_i^B,T)+\underset{a}{}\left(\frac{ϵ_{AA}}{4}(p_{i+a}^Ap_i^A)^2+\frac{ϵ_{AB}}{2}(p_{i+a}^Ap_i^A)(p_{i+a}^Bp_i^B)\frac{ϵ_{BB}}{4}(p_{i+a}^Bp_i^B)^2\right)$$ (17) with a local free energy density $`f(p^A,p^B,T)`$ $`=`$ $`{\displaystyle \frac{zϵ_{AA}}{2}}p_{}^{A}{}_{}{}^{2}zϵ_{AB}p^Ap^B{\displaystyle \frac{zϵ_{BB}}{2}}p_{}^{B}{}_{}{}^{2}`$ (19) $`+kT\left[p^A\mathrm{ln}p^A+p^B\mathrm{ln}p^B+(1p^Ap^B)\mathrm{ln}(1p^Ap^B)\right].`$ This free energy could have been obtained by a simple static mean field approximation of the Hamiltonian. Furthermore, for a closed system (no currents crossing the boundaries), $`F`$ can only decrease. This can be seen explicitly by taking its time derivative, using (13) and noticing that the mobility is always positive: $$\frac{dF}{dt}=\underset{i,\alpha }{}\frac{F}{p_i^\alpha }\frac{dp_i^\alpha }{dt}=\underset{i,j,\alpha }{}M_{ij}^\alpha \left(\mu _j^\alpha \mu _i^\alpha \right)^2.$$ (20) Therefore, the dynamics leads to a state which minimizes the static mean-field free energy. This final state may be the ground state (global minimum) or a metastable state (local minimum); we cannot describe nucleation events, unless we explicitly introduce fluctuations, for example by adding Langevin noise to the deterministic equations. Stated in terms of chemical potentials and mobilities, our kinetic equations have the form of generalized Cahn-Hilliard equations. In contrast to the phenomenological equations, the mobilities depend on the local configuration and are related to the details of the microscopic jump processes, albeit in an approximate manner. This allows in principle an application of these equations to situations far from equilibrium, and to situations where the mobility depends strongly of the local concentration. For the vacancy dynamics considered here, the concentrations of $`A`$ and $`B`$ are independent variables, but their evolution is coupled by the vacancy field. The mobilities are higher in regions with high vacancy concentration. To see this, consider the mobilities for a homogeneous system. For $`p_i^A\overline{p}^A`$ and $`p_i^B\overline{p}^B`$, Eq. (14) becomes $$M_{\mathrm{hom}}^\alpha (\overline{p}^A,\overline{p}^B)=w_0\overline{p}^\alpha (1\overline{p}^A\overline{p}^B)\mathrm{exp}\left[\frac{zϵ_{\alpha A}}{kT}\overline{p}^A\frac{zϵ_{\alpha B}}{kT}\overline{p}^B\right].$$ (21) This expression has a simple interpretation: the prefactor is the mean-field probability of finding an $`\alpha `$-atom and a vacancy on neighboring sites, and $`w_0`$ times the exponential term is the mean jump rate. It is worth noticing that in the mean-field approximation there are no “off-diagonal terms” in the mobility matrix: in general, one should obtain an expression for the currents of the form $`j^\alpha =M^{\alpha A}\mu ^AM^{\alpha B}\mu ^B`$. The reason for these terms missing is probably the suppression of all correlations in Eq. (8) by the mean-field approximation. In the present context, this deficiency seems to be of little importance. It should be noted that even if the off-diagonal mobilities are zero, the same is not true for the diffusion coefficients, because the chemical potentials involve the concentrations of both species. From the free energy density, we can obtain the phase diagram. We will consider only completely attractive models here, where no order-disorder transitions occur ($`ϵ_{AA}>0`$, $`ϵ_{BB}>0`$, $`ϵ_{AA}+ϵ_{BB}2ϵ_{AB}>0`$). The conditions for phase coexistence are that the chemical potentials and the grand potential, $`\mathrm{\Omega }=f\mu ^Ap^A\mu ^Bp^B`$ be equal in the two (or three) phases. This is equivalent to a “common tangent plane” construction: the free energy density, function of two variables, defines a surface in the space $`(p^A,p^B,f)`$. The (homogeneous) chemical potentials as functions of $`p^A`$ and $`p^B`$ define the orientation of planes tangent to this surface. The condition of equal grand potential implies that the points representing two (or more) phases in equilibrium must lie in the same plane. Thus we can obtain the phase diagram by constructing all the planes that are tangent to the free energy surface in at least two points. Various structures of phase diagrams can be obtained . For attractive interactions, quite generally the free energy surface has three minima for sufficiently low temperatures. Then, there exists exactly one plane which is tangent in three points: we have three-phase coexistence. Besides, there are families of double tangent planes which give the coexistence lines for two-phase coexistence. For $`ϵ_{AA}=ϵ_{BB}`$, our model is equivalent to the Blume-Emery-Griffiths model . We will focus in this paper on the specific example $`ϵ_{AB}=ϵ_{AA}/2`$. As we then have $`ϵ_{AA}=ϵ_{BB}=ϵ_{AA}+ϵ_{BB}2ϵ_{AB}`$, the phase diagram is completely symmetric with respect to the interchange of any two components and may be calculated using the analogy with the three state Potts model (see appendix). The phase diagram for $`kT/ϵ_{AA}=0.6`$ and $`z=4`$ (two dimensions) is shown in Fig. 1. There is a large region where an A-rich, a B-rich, and a vacancy-rich “vapor” phase coexist. In the regions of two-phase coexistence, the concentration of the third component is very low. From formula (21) we can immediately deduce that the mobility in the dilute “vapor”-phase, where $`p^A`$ and $`p^B`$ are small, is much larger than in the two dense phases. As we have $`ϵ_{AA}=ϵ_{BB}>ϵ_{AB}`$, the diffusion of the minority component is always faster than that of the majority component in the dense phases. ## III Simulations The part of the phase diagram which is most appropriate for the description of an alloy is the region of AB-coexistence with low vacancy concentration. All studies of lattice gas dynamics with vacancies we are aware of are limited to this area. We will investigate the behavior of finite “droplets” of such a material immersed in a stable “vapor”, a very natural situation which can arise for instance when droplets of a liquid mixture in coexistence with its vapor are rapidly quenched into an unstable state. Evidently, our model is not adapted to describe diffusion processes in a vapor, where diffusion does not take place via activated jumps to nearest neighbor sites. But the important feature of this “vapor” phase is that diffusion is much faster than inside the “solid”. Moreover, the atomic mobility decreases continuously across the vapor-mixture interface, and hence at the surface the diffusion is faster than inside the bulk. As we shall see, this induces fast surface modes. We integrated the mean-field kinetic equations by an explicit Euler scheme. The time step is limited by the numerical stability of the algorithm. For our inhomogeneous system, the “most dangerous” regions of the simulation domain are those where the diffusion is fastest, thus the vapor phase. For the diffusion equation with a diffusivity $`D_{\mathrm{vap}}`$, the maximum time step is $`D_{\mathrm{vap}}/2da^2`$ in $`d`$ dimensions. For the temperatures and vapor compositions we used, $`D_{\mathrm{vap}}/a^2w_0`$ is slightly less than unity. We integrated with a maximum time step of $`1/4w_0`$ in 2D and $`1/6w_0`$ in 3D without encountering numerical instabilities. When we started with step functions as initial conditions, the time step had to be chosen much smaller at the beginning and was then slowly increased to the maximum value. As initial state, we chose “droplets” (i. e. circular domains) or slabs of a mixture with few vacancies (typically some percents) and concentrations $`p_{\mathrm{sol}}^A`$ and $`p_{\mathrm{sol}}^B`$, immersed in a vapor of concentrations $`p_{\mathrm{vap}}^A`$ and $`p_{\mathrm{vap}}^B`$. To trigger the phase separation, we added small fluctuations to the initial state. To assure mass conservation, pairs of neighboring sites and a component (A or B) were randomly chosen, and the concentrations at the two sites were shifted by $`+A_0r`$ and $`A_0r`$, respectively, where $`r`$ is a random number uniformly distributed between $`1`$ and $`1`$. The noise amplitude $`A_0`$ ranged between $`10^5`$ and $`10^2`$. For 2D simulations, we used a lattice of size $`128\times 128`$ with periodic boundary conditions. All our simulations were carried out on workstations and took from 2 to 20 hours of CPU time. We also simulated two 3D-samples on a $`32\times 32\times 64`$ lattice; these took up to 100 hours CPU time. Let us first discuss critical quenches ($`p_{\mathrm{sol}}^A=p_{\mathrm{sol}}^B`$). Snapshot pictures from the time evolution of a droplet are shown in Fig. 2. Phase separation starts at the surface: a fairly regular modulation appears along the mixture-vapor interface. This surface mode triggers phase separation in adjacent regions and propagates into the interior of the sample with a constant velocity, leaving behind a checkerboard-like ordered structure. The domains often coalesce to form stripes. To see more in detail what happens at the interface, we plot in Fig. 3 several snapshots of the concentration profiles along a line which is normal to the surface at a randomly chosen point. The initial step profile quickly relaxes to a smooth shape and stays nearly stationary, until on a slower time scale B is enriched and A depleted; at other interface points, the opposite happens. An oscillatory concentration profile develops, with an amplitude which has its maximum in the interface and decays into the solid. We will show below that the envelope of this oscillation becomes a decaying exponential away from the surface. The fact that the perturbation decays exponentially with the distance from the surface, but grows exponentially with time, explains the constant propagation velocity of the decomposition front. Note that the concentrations in the vapor vary only very slightly: the vapor is a stable phase, and hence perturbations decay. Because of the fast diffusion in the vapor, the surroundings of the droplet act as a particle reservoir. As the surface mode develops, the mixture-vapor interface deforms, and fingers of vapor start to grow into the interior of the droplet. This is the result of a Mullins-Sekerka instability with respect to the vacancies. To clarify this point, we show in Fig. 4 snapshots of the vacancy concentration during the decomposition process. At the beginning, the vacancy concentration stays constant and equal to the initial value. Once the decomposition process reaches its nonlinear stage, however, vacancies are expelled from the domains of the new equilibrium phases. These excess vacancies have to diffuse to the surface of the droplet. A finger of vapor protruding into the mixture enhances the concentration gradients around its tip, and hence grows faster than a flat portion of the surface. In addition, the diffusion is faster in the initial mixture than in the phase-separated domains. The growth of the fingers stops when a layer of decomposed material has formed around their entire contour. The diffusion then takes place mainly along the domain boundaries, where the vacancies are enriched. The fingers are smoothed out by the subsequent coarsening process. The propagation of the surface mode stops when the bulk modes enter the nonlinear regime. The structures in the interior of the sample are the usual bicontinuous patterns of bulk spinodal decomposition at equal volume fractions. Both surface and bulk structures coarsen by the evaporation-condensation mechanism. This process is faster at the surface because of the rapid diffusion through the vapor. At the exterior surface of the droplet, there exist trijunction points where the three phases are in contact. The angles between the interfaces at these points are fixed by the local equilibrium between the three surface tensions of Av, Bv, and AB-interfaces. For our symmetric choice of interaction energies, all angles are $`120^{}`$ in local equilibrium. The initial values for the concentrations in our example represent a special choice: the chemical potentials of the two species and the grand potential have the same value in the mixture and in the vapor. There exists exactly one set of concentrations satisfying these conditions at a given temperature. In this special case, there is no net mass flux between vapor and mixture, and the interface is at rest. This choice was mainly made to simplify the stability calculations to be presented below. The existence of a surface mode, however, is not limited to this special case. For other initial conditions (but still $`p_{\mathrm{sol}}^A=p_{\mathrm{sol}}^B`$ and $`p_{\mathrm{vap}}^A=p_{\mathrm{vap}}^B`$), a surface mode develops while the mixture-vapor interface slowly moves. To obtain more quantitative information about the phase separation process, we repeated our simulation with the same parameters, but this time in a stripe geometry: a slab of mixture along the $`y`$-direction is immersed in the vapor; we thus have two straight interfaces normal to the $`x`$ axis. A convenient quantity for the analysis of phase separation processes is the dynamical structure factor. Of particular interest in our case is the difference between bulk and surface behavior. Therefore, we define a one-dimensional structure factor along the interface: $$S^\alpha (k_y,x,t)=\left|\frac{1}{L_y}\underset{\{j|x_j=x\}}{}p_j^\alpha e^{ik_yy_j}\right|^2,$$ (22) where $`x_j`$ and $`y_j`$ are the coordinates of lattice site $`j`$. The sum goes over a lattice plane (in 2D: a line) at a fixed $`x`$-coordinate, and $`k_y`$ is parallel to the surface. Figure 5 shows plots of this quantity for different times. We observe the characteristic Cahn-Hilliard behavior: in the beginning, linear superposition is valid, and the structure factor for each mode grows exponentially with a growth rate depending on $`k_y`$. In particular, all the structure factor curves intersect in one point, corresponding to the marginally stable mode. At later times, the nonlinear terms in the equations of motion couple the different modes, and the shape of the structure factor curve changes, as can be noted at the last time for the surface layer. The two sets of curves have very different amplitudes, and the maximum is located at $`ak_y0.3`$ at the surface, and at $`ak_y1`$ in the bulk. These differences between bulk and surface behavior will be addressed in Sec. 4. Very different structures appear when the concentration of the mixture is sufficiently off-critical. Fig. 6 shows the evolution of a slab of an AB mixture with a concentration ratio 60:40. The initial concentrations were again chosen to give equal chemical potentials and grand potential in the two bulk phases; note that now $`\mu ^A\mu ^B`$. The minority component rapidly segregates at the surface, triggering a “spinodal wave” normal to the surface. A surface mode with modulations along the interface is still present and leads to a destabilization of the first layer of the minority component, which disintegrates into regularly spaced droplets. These droplets coarsen rapidly. This time, there is no Mullins-Sekerka instability with respect to the vacancies, because the formation of the first decomposed layer rapidly blocks the exchange of vacancies between the interior of the sample and the vapor. Inside the sample, the surface-induced wave travels until the bulk modes reach their nonlinear regime. There is a competition of droplet and stripe patterns during the subsequent coarsening process. The droplets in the interior of the sample coarsen more rapidly than the stripes at the surface, not surprisingly as the driving force for the evaporation-condensation mechanism of coarsening is the curvature of the domain walls. Ultimately, the stripes start to break up and are “infected” by the droplet pattern. A simulation for a “droplet” geometry is shown in Fig. 7. The obviously symmetric configuration of the outer domains of B-rich phase is due to the anisotropy of the surface tensions introduced by the lattice . It is interesting to note that this effect is not immediately visible in the critical quenches. The symmetric configuration, however, is only transient: on the last snapshot picture, the smallest of the four outer B domains is about to evaporate. Note also the symmetry in the first and second ring of inner droplets; this pattern is later destroyed by the coarsening of the bulk structures. Figure 8 shows the evolution of the interface profiles. In contrast to the critical quench, there is no smooth, nearly stationary state at intermediate times. The profile immediately starts to show the onset of oscillations. Also, the evolution is faster: at $`t=12000w_0^1`$, when the separation of A and B is still tiny in Fig. 3, it is already well pronounced in the off-critical case. The difference between the two evolutions can be qualitatively understood by a look on the free energy surface, plotted in Fig. 9. In this figure, the points a, b, and c denote the initial compositions of the critical and off-critical mixtures and the vapor, respectively. In the symmetric case, we can connect the points a and b along a symmetry axis of the free energy density, with zero slope along the A-B-direction. This means that an interface which is situated completely on this line is stationary with respect to a separation of A and B. For the case of an off-critical concentration, we cannot find any such line going from b to c, and an unstable stationary interface does not exist. At the surface of the mixture, chemical potential gradients will always lead to a segregation of one of the components to the surface. Which component is attracted to the surface depends on the choice of concentrations in the vapor phase. Let us mention that this argument is not entirely complete, because it considers only the free energy density, whereas the complete free energy also contains the discrete gradient terms. In our simulations, however, we never observed any unstable stationary interface configuration for off-critical compositions (or asymmetric interaction energies). The transition from the checkerboard structures to the stripes is gradual. For slightly asymmetric concentrations, the segregation to the surface is slow, and the surface mode has enough time to grow. We observed checkerboard structures up to a concentration ratio of approximately 54:46. For this composition, checkerboard structures and stripes appear simultaneously on different portions of the surface. Similar findings are also valid when we vary the interaction parameters in our model: for $`ϵ_{BB}ϵ_{AA}`$, we have always observed surface-directed spinodal waves; however, a surface mode should appear in this case if for some asymmetric compositions the surface segregation becomes slow. These findings are consistent with calculations for mixtures near flat substrates using continuous equations of Cahn-Hilliard type: spinodal waves occur when one component is attracted to the substrate , whereas surface modes have been found in the case of a substrate which prefers neither of the components of the mixture . Our simple model shows that this general behavior is also valid for free surfaces. In 3D samples we also find fast surface modes, but because the vapor-mixture interface is now two-dimensional, the patterns occurring at the surface for critical quenches are those of 2D bulk spinodal decomposition (Fig. 10). A bicontinuous pattern forms at the surface and propagates into the sample, replicating itself in an oscillatory manner, that is we always find an oscillating concentration profile when we look normal to the surface. In the off-critical case, a spinodal wave occurs as in two dimensions (Fig. 11). In all the simulations presented so far, the domains of the A- and B-rich phases stick together, because the surface tension of an Av- or Bv-interface is more than the half of the one of an AB-interface. This changes if we lower the interaction energy $`ϵ_{AB}`$: for $`ϵ_{AB}=0`$, the droplet “explodes”: thin layers of vapor penetrate into the interior of the droplet along the forming AB-interfaces, and the domains of A and B are slowly drifting apart (Fig. 12). This reminds of the decomposition of a binary mixture in the presence of a surfactant. The presence of sharp corners and facets in the domain shapes of the last snapshot indicates that the surface tension anisotropy is quite large. ## IV Linear stability analysis ### A Bulk We will now study more in detail the checkerboard structures. To this end, we must calculate the growth rates of bulk and surface modes. We start with the bulk modes. A homogeneous system of overall composition $`\overline{p}^A`$ and $`\overline{p}^B`$ is perturbed by small fluctuations of the occupation probabilities: $$p_i^\alpha (t)=\overline{p}^\alpha +\delta _i^\alpha (t)(\alpha =A,B),$$ (23) with $`\delta _i^\alpha 1`$. We linearize the chemical potentials around the average concentrations. Introducing a vector notation with respect to the two species of particles, we obtain starting from Eqs. (15) $$\left(\begin{array}{c}\mu _i^A\\ \mu _i^B\end{array}\right)=\left(\begin{array}{c}\overline{\mu }^A\\ \overline{\mu }^B\end{array}\right)𝐄\left(\begin{array}{c}\mathrm{\Delta }_a\delta _i^A\\ \mathrm{\Delta }_a\delta _i^B\end{array}\right)+𝐒\left(\begin{array}{c}\delta _i^A\\ \delta _i^B\end{array}\right).$$ (24) Here, $`\overline{\mu }^\alpha `$ are the unperturbed values of the chemical potentials, the matrix $`𝐄`$ of the interaction energies is $$𝐄=\left(\begin{array}{cc}ϵ_{AA}& ϵ_{AB}\\ ϵ_{AB}& ϵ_{BB}\end{array}\right),$$ (25) and $`S`$ is the matrix of the second derivatives of the free energy density, taken at $`\overline{p}^A`$ and $`\overline{p}^B`$: $$𝐒=\left(\begin{array}{cc}S_{AA}& S_{AB}\\ S_{AB}& S_{BB}\end{array}\right)\mathrm{with}S_{\alpha \beta }=\frac{^2f}{p^\alpha p^\beta }|_{\overline{p}^A,\overline{p}^B}.$$ (26) The variations in the chemical potentials create currents. To obtain these currents to order one in $`\delta _i^\alpha `$, we may use the unperturbed values of the mobilities, $$M_{ij}^\alpha =M_{\mathrm{hom}}^\alpha (\overline{p}^A,\overline{p}^B)=\overline{M}^\alpha .$$ (27) The equations of motion become: $$\frac{d}{dt}\left(\begin{array}{c}\delta _i^A\\ \delta _i^B\end{array}\right)=\left(\begin{array}{c}\overline{M}^A\mathrm{\Delta }_a\mu _i^A\\ \overline{M}^B\mathrm{\Delta }_a\mu _i^B\end{array}\right).$$ (28) As a homogeneous system is translation invariant with respect to the lattice vectors, solutions of the linearized equations are of the form: $$\left(\begin{array}{c}\delta _j^A\\ \delta _j^B\end{array}\right)=\left(\begin{array}{c}\delta ^A\\ \delta ^B\end{array}\right)\mathrm{exp}\left(i\stackrel{}{k}\stackrel{}{x}_j+\omega (\stackrel{}{k})t\right),$$ (29) where $`\stackrel{}{x}_j`$ is the position vector of site $`j`$ in real space, and $`\stackrel{}{k}=(k_x,k_y)`$ is the wave vector of the perturbation. The growth rate $`\omega (\stackrel{}{k})`$ and the coefficients $`\delta ^\alpha `$ have to be determined by solving the eigenvalue problem $$\omega (\stackrel{}{k})\left(\begin{array}{c}\delta ^A\\ \delta ^B\end{array}\right)=A_\stackrel{}{k}\left(\begin{array}{cc}\overline{M}^A& 0\\ 0& \overline{M}^B\end{array}\right)\left(𝐒A_\stackrel{}{k}𝐄\right)\left(\begin{array}{c}\delta ^A\\ \delta ^B\end{array}\right).$$ (30) Here, the terms $$A_\stackrel{}{k}=4\mathrm{sin}^2(k_xa/2)4\mathrm{sin}^2(k_ya/2)$$ (31) arise from the discrete Laplacians. Eq. (30) is quadratic in $`\omega `$ and thus gives a stability spectrum with two branches. Each of these branches can be stable ($`\omega `$ is negative for all wave vectors) or unstable. In the latter case, positive growth rates occur for small values of $`|\stackrel{}{k}|`$. The number of unstable branches is equal to the number of negative eigenvalues of the matrix $`𝐒`$. This can be easily seen by taking the limit $`|\stackrel{}{k}|0`$ in Eq. (30). The terms proportional to $`A_\stackrel{}{k}^2`$ can be neglected, and we have $$\omega (\stackrel{}{k})\left(\begin{array}{c}\delta ^A\\ \delta ^B\end{array}\right)=|\stackrel{}{k}|^2𝐒\left(\begin{array}{c}\delta ^A\\ \delta ^B\end{array}\right).$$ (32) The matrix S is related to the curvature of the free energy surface. If both eigenvalues are positive, the surface is locally convex, and the homogeneous state is stable. For two eigenvalues of different sign, the surface has locally the structure of a saddle point, and we have partial instability: only fluctuations in the concave direction in concentration space are amplified. Finally, for two negative eigenvalues, the free energy surface is concave and all perturbations grow. The frontiers between these regions of different stability behavior are given by the spinodal surfaces, defined by the condition $$det𝐒=0.$$ (33) This generalizes the concept of a spinodal curve to our three-component system. For the special case of symmetric interaction energies and equal average concentration, Eq. (30) can be considerably simplified because then we have $`\overline{M}^A=\overline{M}^B=\overline{M}`$. In addition, the matrix $`𝐋`$ defined by $$𝐋=A_\stackrel{}{k}(𝐒A_\stackrel{}{k}𝐄)$$ (34) is symmetric. We obtain immediately $$\omega _\pm (\stackrel{}{k})=\overline{M}(L_{AA}\pm L_{AB}),$$ (35) and the associated eigenvectors are $$\left(\begin{array}{c}\delta ^A\\ \delta ^B\end{array}\right)=\left(\begin{array}{c}1\\ \pm 1\end{array}\right).$$ (36) This last result shows that of the two branches of the dispersion relation, the first describes the separation into a dense AB mixture and a dilute vapor, whereas the second gives a separation between A and B, leaving the local vacancy concentration unchanged. Fig. 13 shows a cut through the spinodal surfaces along the axis $`\overline{p}^A=\overline{p}^B`$. We also indicate the regions of different stability behaviors, and show the typical shapes of $`\omega _+(\stackrel{}{k})`$ and $`\omega _{}(\stackrel{}{k})`$ in these regions. The interior of our samples is in the region where only the mode separating A and B is unstable: the local vacancy concentration stays unchanged in the linear stage. Other modes of decomposition in a homogeneous ternary system have been studied by Chen . The anisotropy due to the lattice structure enters in the above formulae by the factor $`A_\stackrel{}{k}`$. For the long wavelength perturbations considered here ($`a|\stackrel{}{k}|<1`$), the relative variations of $`A_\stackrel{}{k}`$ with orientation are of the order of a percent. We will therefore neglect this dependence and use a wave vector along one of the lattice directions for comparisons to numerical results. For the example of the simulation shown in Fig. 2, i.e. with $`kT/ϵ_{AA}=0.5`$ and $`\overline{p}^A=\overline{p}^B=0.46464`$, we found a maximum bulk growth rate of $`\omega _b=1.06\times 10^4w_0`$ at a wave number $`k_b=1.003/a`$. ### B Surface We must now analyze the stability spectrum at the surface. This is more difficult than in the bulk because the initial state is now heterogeneous. To simplify the problem, we will treat the case of a flat interface which is normal to one of the lattice directions, say $`x`$. Then, in the initial state all lattice sites in a layer at a given $`x`$ coordinate have the same concentrations. In what follows, we will replace the site indices “$`i`$” used so far by a pair of indices “$`n,j`$”, where $`n`$ numbers the layer, and $`j`$ numbers the sites in the $`y`$ direction. We use the same values of the concentrations in the bulk phases as for our simulations. As the chemical potentials are equal in the two phases, the initial interface state can be obtained by fixing the chemical potentials to their appropriate values and numerically solving the one-dimensional version of the finite difference equations Eq. (15) for the concentrations in the $`n`$th layer $`\overline{p}_n^\alpha `$. Whereas the translation invariance is broken along the $`x`$-axis, it is preserved along $`y`$, and hence we can still use a Fourier representation. We write the perturbed state as $$p_{n,j}^\alpha (t)=\overline{p}_n^\alpha +\delta _{n,j}^\alpha (t),$$ (37) $$\delta _{n,j}^\alpha (t)=\delta _n^\alpha \mathrm{exp}\left(ik_yja+\omega (k_y)t\right),$$ (38) with two unknown coefficients $`\delta _n^A`$ and $`\delta _n^B`$ per layer. We start with the linearization of the chemical potentials. The discrete Laplacians appearing in Eq. (15) give contributions of the form $`\mathrm{\Delta }_a\delta _{n,j}^\alpha `$ $`=`$ $`\delta _{n+1,j}^\alpha +\delta _{n1,j}^\alpha +\delta _{n,j+1}^\alpha +\delta _{n,j1}^\alpha 4\delta _{n,j}^\alpha `$ (39) $`=`$ $`\left[\delta _{n1}^\alpha +\delta _{n+1}^\alpha +\left(A_\stackrel{}{k}2\right)\delta _n^\alpha \right]\mathrm{exp}\left(ik_yja+\omega (k_y)t\right).`$ (40) Denoting by $`\overline{\mu }^\alpha `$ the (constant) values of the unperturbed chemical potential, and by $`𝐒(n)`$ the matrix of the second derivatives, taken at $`\overline{p}_n^A`$ and $`\overline{p}_n^B`$, we find $$\left(\begin{array}{c}\mu _{n,j}^A\overline{\mu }^A\\ \mu _{n,j}^B\overline{\mu }^B\end{array}\right)=\left(𝐒(n)(A_\stackrel{}{k}2)𝐄\right)\left(\begin{array}{c}\delta _{n,j}^A\\ \delta _{n,j}^B\end{array}\right)𝐄\left(\begin{array}{c}\delta _{n1,j}^A\\ \delta _{n1,j}^B\end{array}\right)𝐄\left(\begin{array}{c}\delta _{n+1,j}^A\\ \delta _{n1,j}^B\end{array}\right).$$ (41) The next step is the linearization of the equations of motion. As in the homogeneous case, the mobilities can be taken in the unperturbed initial state. But because the concentrations vary through the interface, we cannot use the expression for the homogeneous system. The appropriate expressions can be obtained from Eq. (14) by taking the limit $`\mu _j\mu _i`$ with $`p_jp_i`$: $$\overline{M}_{mn}^\alpha =w_0\frac{(1\overline{p}_m^A\overline{p}_m^B)(1\overline{p}_n^A\overline{p}_n^B)}{kT}\mathrm{exp}\left(\frac{\overline{\mu }^\alpha }{kT}\right)$$ (42) (the two indices of $`\overline{M}_{nm}`$ are two layer indices). The linearized equation of motion becomes $$\frac{d}{dt}\delta _{n,j}^\alpha =\overline{M}_{nn}^\alpha \left(\mu _{n,j+1}^\alpha 2\mu _{n,j}^\alpha +\mu _{n,j1}^\alpha \right)+\overline{M}_{nn+1}^\alpha \left(\mu _{n+1,j}^\alpha \mu _{n,j}^\alpha \right)+\overline{M}_{nn1}^\alpha \left(\mu _{n1,j}^\alpha \mu _{n,j}^\alpha \right).$$ (43) Inserting Eqs. (41) and (38), one obtains an eigenvalue problem for $`\omega `$, the eigenvectors given by the set of unknown coefficients $`\delta _n^\alpha `$. This means that the matrix to diagonalize is of size $`2L\times 2L`$, where $`L`$ is the number of layers in the $`x`$ direction, and that the corresponding dispersion relation has $`2L`$ branches. This is a standard linear algebra problem and can be solved numerically. In the initial stage of the development, the surface mode is localized in a small number of layers around the surface. Hence we can simplify the problem by considering a small “solution region” around the interface. We assume perturbations outside a narrow region of $`N`$ layers centered around the interface to be zero. We diagonalized the resulting $`2N\times 2N`$ matrix numerically and obtained the maximum growth rate, $`\omega _s`$, and the corresponding wave vector $`k_s`$. As shown in Fig. 14, these quantities become independent of $`N`$ if a sufficient number of layers is included . To compare the results of this stability analysis to the simulations, we extracted the stability spectrum of the surface from the structure factor data of Fig. 5. For an exponentially growing Fourier mode, we have $`S(t)\mathrm{exp}(2\omega t)`$, and the growth rate can be obtained as $`\omega =\mathrm{ln}(S(t_2)/S(t_1))/2(t_2t_1)`$. Fig. 15 shows that the agreement with the theoretical prediction is excellent, except for very small wave numbers. This discrepancy can be explained by the fact that for large wavelengths, more layers have to be included, as the diffusion field in the vapor will be appreciably modified up to a distance from the interface which is comparable to the wavelength. In conclusion, the maximum growth rate and the corresponding wave vector can be accurately predicted by a calculation involving only a small number of layers. For the example of Fig. 2, we find a maximum growth rate of $`\omega _s=4.47\times 10^4w_0`$ at a wave number $`k_s=0.349/a`$. This surface mode grows more than four times faster than the fastest bulk mode, and at a wavelength about three times larger than the wavelength of the typical bulk pattern. We verified indeed in our simulations that the “incubation time”, i. e. the time after which the growing perturbations become visible in a given visualization, was about four times larger in the bulk than at the surface. Having verified that our method gives accurate results, we can apply it to study the behavior of bulk and surface modes for different temperatures. The results are given in Table I. All calculation were performed with $`N=16`$, except for the highest temperature where $`N=32`$ was used. The bulk growth rate has a maximum around $`kT/ϵ_{AA}=0.6`$ and decreases as the temperature is lowered, because the activated dynamics lead to small values of the mobility for low temperatures. On the other hand, for the highest temperature the system is near a spinodal surface, and the driving force for phase separation is small, leading also to a lower growth rate. The surface growth rate follows the same trends. At the highest temperature considered, $`\omega _s`$ is lower than $`\omega _b`$, which means that a proper surface mode does not exist any more. The ratio $`\omega _s/\omega _b`$ increases monotonically as the temperature is lowered. It should be noted that the mean-field approximation becomes increasingly inaccurate for low temperatures. Also, the interfaces become very sharp, which leads to strong lattice effects. It is also straightforward to repeat these calculations for three dimensions. In fact, the phase diagram stays the same if we scale the temperature by a factor of $`1.5`$. On the other hand, the interface shapes and the growth rates are different, because the balance between the discrete Laplacians and the local terms is altered in Eqs. (15) for the chemical potentials. In the stability calculations, the magnitude of the stability matrix $`𝐒`$ is modified whereas the terms arising from the Laplacians stay the same. The final results for the growth rates and wave vectors, shown in Table II, are similar to the two-dimensional case. ### C Propagation of the surface patterns The surface mode propagates into the interior of the sample, enforcing its characteristic wavelength in the direction parallel to the surface. This behavior can be understood by considering the solution of the eigenvalue equation away from the interface. Even if the surface modes are localized at the vapor-mixture interface, they are eigenmodes of the whole system and grow everywhere with the same growth rate $`\omega _s`$. But away from the interface, the initial state becomes homogeneous, and we can use Eq. (35), derived for the bulk, if we allow for a decay of the amplitude in the direction normal to the interface by introducing a complex wave number $`k_x`$. As the growth rate and the $`y`$-component of the wave vector are already known from the analysis of the surface instability, $`k_x`$ is the only remaining unknown. The relevant dispersion relation is $`\omega _{}(\stackrel{}{k})`$, and we obtain the equation $$A_\stackrel{}{k}^2\left(ϵ_{AA}ϵ_{AB}\right)A_\stackrel{}{k}\left(S_{AA}S_{AB}\right)+\frac{\omega _s}{\overline{M}}=0.$$ (44) This equation can either be solved numerically using the exact expression for $`A_\stackrel{}{k}`$ or analytically with the approximation $`A_\stackrel{}{k}k_x^2k_y^2`$. The latter method leads to a biquadratic equation for $`k_x`$. In both cases, there are four solutions of the form $$k_x=\pm k^{}\pm ik^{\prime \prime }$$ (45) with $`k^{}`$ and $`k^{\prime \prime }`$ real and positive. The two solutions with negative imaginary part diverge in the bulk and have to be discarded. The other two solutions lead to modes proportional to $`\mathrm{exp}(\pm ik^{}x+\omega _stk^{\prime \prime }x)`$, that is modes with an oscillatory concentration profile and an envelope which decays exponentially with the distance from the surface, but grows exponentially in time. One can define a propagation velocity of the decomposition front by the phase velocity of the envelope, $$v=\frac{\omega _s}{k^{\prime \prime }}.$$ (46) We compared the values for the wavelength normal to the surface, $`\lambda _x=2\pi /k^{}`$, and the propagation velocity obtained by this method to our simulations and found good agreement . The competition between surface and bulk modes leads to the interesting consequence that the thickness of the layer near the surface where the surface-directed patterns prevail depends on the strength of the initial fluctuations. This can be seen as follows. For an exponentially growing mode with initial amplitude $`\delta _0`$, the time to reach a threshold amplitude $`\delta _{nl}`$ where nonlinear couplings between different modes become important is $$\tau \frac{1}{\omega }\mathrm{ln}\frac{\delta _{nl}}{\delta _0}.$$ (47) Now, a surface mode “propagates” into the sample approximately between the time $`\tau _s`$ when it is well developed at the surface, and the time $`\tau _b`$ when the bulk modes reach the nonlinear stage. The distance $`d`$ the front propagates is therefore given by $$dv(\tau _b\tau _s)=\frac{1}{k^{\prime \prime }}\left(\frac{\omega _s}{\omega _b}1\right)\mathrm{ln}\frac{\delta _{nl}}{\delta _0}.$$ (48) The thickness of the surface structures depends logarithmically on the strength of the initial noise. This reasoning applies both to the symmetrical checkerboard structures and the surface-directed spinodal waves and was confirmed by our simulations. ## V Conclusions We have developed mean field kinetic equations to describe the dynamics of a lattice gas model of a binary alloy with vacancies (ABv model) in which diffusion takes place by the vacancy mechanism only. Despite the simplicity of the model, we observe a rich variety of phase separation patterns at free surfaces between an unstable mixture and a stable vapor. The most spectacular effect is a fast surface mode. It creates ordered patterns at the surface with a length scale which is clearly distinct from the characteristic scale of the bulk patterns. This mode appears in a small range of parameters around a symmetric point where the mixture-vapor interface is “neutral”, that is none of the components of the mixture segregates to the surface. On the other hand, if such segregation occurs and is rapid enough, the surface mode is suppressed, and instead surface-directed spinodal waves are observed which create a striped pattern along the surface. Both patterns, once they have formed at the surface, propagate into the sample over a distance which is related to the difference in surface and bulk growth rates and the strength of the initial fluctuations. Our approach starts from a minimal model with very few assumptions. Therefore, it cannot be used to model a particular experimental situation. But it allows to identify some basic ingredients necessary for the formation of such surface structures in a fairly well-defined setting, and we can draw some conclusions which should be generally valid. The existence of the fast surface mode is related to the fact that in our model the surface atoms are more mobile than in the bulk, an assumption which seems reasonable for many interfaces between a dense and a dilute phase. The characteristic growth rates of the surface modes can be calculated by a linear stability analysis starting from the initial interface profile. We have shown that accurate results can be obtained by solving the resulting eigenvalue problem in a small domain around the interface. The characteristics of the propagation of the decomposition front can be obtained by connecting the solution of the surface problem to the bulk solution. This method could be applied as well to continuum models of Ginzburg-Landau type by using an appropriate discretization. Which type of surface structure occurs is related to the time scales for segregation and surface phase separation. If the segregation is slow, surface spinodal decomposition is dominant, otherwise spinodal waves are observed. As in a realistic system, the interactions between different species are always different, spinodal waves are the more generic pattern. The surface mode should show up, however, in a small range of initial compositions if the interactions are not too different. More work is needed to clarify this point. A very interesting point is that the thickness of the layer of surface patterns depends on the initial fluctuation strength. This is not the case in bulk spinodal decomposition, where a rescaling of the initial fluctuations amounts simply to a shift in time. The arguments leading to Eq. (48) are fairly general and should apply to all systems where a fast surface mode is in competition with a slow bulk mode. In rapid quench experiments, the initial fluctuation spectrum is mainly determined by the temperature before the quench. Therefore, the thickness of the surface layer may depend on the initial temperature. In view of the logarithmic dependence of $`d`$ on the noise strength, this effect might be difficult to observe; however, if the initial state of the system is close to a critical point, the fluctuation amplitude is a rapidly varying function of temperature. Surface effects in spinodal decomposition have been studied recently in polymers . Evidently, our equations cannot be applied to polymers. But from our results it is simple to construct a Ginzburg-Landau theory in which surface modes occur: it is sufficient to introduce a mobility function which explicitly depends on the density and which enhances diffusion on the surface. It would be interesting to compare our findings to Monte Carlo simulations which naturally contain the fluctuations. It has been shown, however, that short-range lattice gas models do not exhibit linearly superposed, exponentially growing modes in the early stages of the phase separation. This can be attributed to the fact that the initial noise is too strong for a linearization of the equations of motion to be valid . On the other hand, for models with longer range interactions the Cahn-Hilliard behavior is restored , and we would expect surface modes in such models even with stochastic dynamics. In summary, our approach allows to explore the rich dynamics of the ABv model and to relate the structures that form spontaneously during phase separation to the parameters of the microscopic model. Here, we have only explored a small part of the possible behaviors in this model, because we have limited ourselves to attractive interactions. There are other interesting questions which could be addressed in the framework of this model and using mean field kinetic equations, for example the influence of the vacancy distribution on the coarsening behavior, or the interplay between phase separation, short range ordering and vacancy distribution. ###### Acknowledgements. We have benefited from valuable discussions with W. Dieterich, H.-P. Fischer, and P. Maass. We would like to thank Jean-François Colonna for his help with the generation of the 3D pictures. One of us (M.P.) was supported by a grant from the Ministère de l’Enseignement Supérieur et de la Recherche (MESR). Laboratoire de Physique de la Matière Condensée is Unité de Recherche Associée (URA) 1254 to CNRS. ## Phase diagram Let us first show that the Hamiltonian of the ABv-model can be derived from a general ternary Hamiltonian. We assume that the sites of the lattice can now be occupied by three different sorts of atoms, $`A`$, $`B`$, or $`C`$. In terms of the occupation numbers $`n_i^A`$, $`n_i^B`$, and $`n_i^C`$, the energy of a configuration is given by $$H=\underset{\alpha ,\beta }{}\underset{i,j}{}ϵ_{\alpha \beta }^{}n_i^\alpha n_j^\beta \underset{\alpha }{}\underset{i=1}{\overset{N}{}}\mu _{}^{}{}_{0}{}^{\alpha }n_i^\alpha .$$ (49) The interaction energies in the ternary system are primed to distinguish them from the $`ϵ`$’s in Eq. (1). Here, the $`\mu _{}^{}{}_{0}{}^{\alpha }`$ are external chemical potentials, equivalent in the spin language to “generalized magnetic fields” acting only on one possible spin state. Using the constraint of single occupancy, $`n_i^A+n_i^B+n_i^C=1i`$, we can eliminate the occupation numbers of one species, say $`C`$, and we obtain $$H=\underset{i,j}{}\left[ϵ_{AA}n_i^An_j^A+ϵ_{BB}n_i^Bn_j^B+ϵ_{AB}(n_i^An_j^B+n_i^Bn_j^A)\right]\underset{i=1}{\overset{N}{}}\left(\mu _0^An_i^A+\mu _0^Bn_i^B\right),$$ (50) where the effective interaction energies and chemical potentials appearing in this last equation are $`ϵ_{AA}`$ $`=`$ $`ϵ_{AA}^{}+ϵ_{CC}^{}2ϵ_{AC}^{},`$ (51) $`ϵ_{BB}`$ $`=`$ $`ϵ_{BB}^{}+ϵ_{CC}^{}2ϵ_{BC}^{},`$ (52) $`ϵ_{AB}`$ $`=`$ $`ϵ_{AB}^{}+ϵ_{CC}^{}ϵ_{AC}^{}ϵ_{BC}^{},`$ (53) $`\mu _0^A`$ $`=`$ $`\mu _{}^{}{}_{0}{}^{A}\mu _{}^{}{}_{0}{}^{C}zϵ_{CC}^{}+zϵ_{AC}^{}\mathrm{and}`$ (54) $`\mu _0^B`$ $`=`$ $`\mu _{}^{}{}_{0}{}^{B}\mu _{}^{}{}_{0}{}^{C}zϵ_{CC}^{}+zϵ_{BC}^{}.`$ (55) When the total number of particles of each species is conserved, the chemical potential terms in Eq. (50) are constants. Then, we are back to the Hamiltonian Eq. (1). Let us now consider the three state Potts model, with ternary interaction energies $`ϵ_{AA}^{}=ϵ_{BB}^{}=ϵ_{CC}^{}=1`$ and $`ϵ_{\alpha \beta }^{}=0`$ for $`\alpha \beta `$. This gives the effective interactions $`ϵ_{AA}=ϵ_{BB}=2,ϵ_{AB}=1`$; by changing the temperature scale, we obtain the values $`ϵ_{AA}=ϵ_{BB}=1,ϵ_{AB}=0.5`$ used in the present paper. Furthermore, from the symmetry of the Potts model it is obvious that three phase coexistence is possible below the critical temperature at zero magnetic fields. This is equivalent, in the ABv model, to $$\mu ^A=zϵ_{AA}p^Azϵ_{AB}p^B+kT\frac{p^A}{1p^Ap^B}=zϵ_{AA}/2$$ (56) and an equivalent equation for $`\mu ^B`$. These equations have to be solved numerically to obtain the equilibrium concentrations. The result for the A-rich phase is shown in Fig. 16. The coexistence of A-rich, B-rich, and “vapor”-phases terminates at a quadruple point: a fourth minimum in the free energy surface, located at the symmetric point $`p^A=p^B=p^v=1/3`$, develops, and at $`kT=1/(2\mathrm{ln}2)0.721`$ the four minima are exactly at the same level. Above this temperature, there is a narrow temperature range above which the four minima persist, but now the fourth minimum is lowest, and we have three different regions of three-phase coexistence (not shown in the figure). Above $`kT=0.75`$, only the “symmetric” minimum remains.
no-problem/9906/astro-ph9906020.html
ar5iv
text
# Where is SGR1806-20? ## 1 Introduction The four known soft gamma repeaters (SGRs) are neutron stars in or near radio or optical supernova remnants. SGR1806-20 was discovered in 1986 (Laros et al. 1986) and underwent a period of intense activity in 1987 (Laros et al. 1987, Kouveliotou et al. 1987) which led to its localization to an $``$ 400 arcmin<sup>2</sup> error ellipse (Atteia et al. 1987). Based on this position, Kulkarni & Frail (1993) suggested that the SGR was associated with the Galactic radio supernova remnant (SNR) G10.0-0.3. This was confirmed when the ASCA spacecraft observed and imaged the source in outburst, leading to a $``$ 1’ radius error circle (Murakami et al. 1994). ROSAT observations of the quiescent X-ray source associated with SGR1806-20 confirmed the ASCA data (Cooke 1993; Cooke et al. 1993). It is believed that the SGRs are ’magnetars’, i.e. single neutron stars in which the magnetic field energy dominates all other sources of energy, including rotation (Duncan & Thompson 1992). In the case of SGR1806-20, evidence for this model comes from observations of the period and period derivative of the quiescent soft X-ray emission (Kouveliotou et al. 1998). Studies of the radio nebula show evidence for changes in its morphology on $``$ year timescales, and suggest that the neutron star may be located at the non-thermal core of the radio emission (Frail et al. 1997). The position of the core also coincides with that of an unusual star, identified as a luminous blue variable (LBV) by van Kerkwijk et al. (1995). This appears to be the only case so far of an SGR with an optical stellar counterpart, and the connection between this object and the SGR has been unclear up to now. SGR1806-20 has remained active over the past several years, and many bursts have been detected by the Interplanetary Network (IPN), consisting primarily in this case of BATSE, Ulysses , and KONUS-WIND. However, only eight events have been intense enough to trigger both Ulysses and a near-Earth spacecraft, resulting in high time resolution data (the other bursts were recorded with lower time resolution by one or more instruments). It is these triggered events which lead to the most precise determination of the source position by triangulation. ## 2 Observations Details of the eight triggered bursts are given in Table 1. In each case, triangulation using Ulysses and either BATSE or KONUS results in a single annulus of width $``$23 - 28” which defines the possible arrival direction for the burst. Two such annuli define an error box, if the angular separation between their centers is sufficient to prevent the annuli from intersecting at grazing incidence. Over the $``$2 yr period analyzed here, the Ulysses -Earth vector moved sufficiently to define a non-degenerate error box. With three or more annuli, the problem of defining the source location becomes overdetermined, and we can use a statistical method to derive the most probable source location. This consists of defining a chisquare which is a function of an assumed source position in right ascension and declination, and of the parameters describing the eight annuli. Let $`\alpha ,\delta `$ be the right ascension and declination of the assumed source position, and let $`\alpha _i,\delta _i,\theta _i`$ be the right ascension, declination, and radius of the ith annulus. Then the angular distance d<sub>i</sub> between the two is given by $$d_i=\theta _i\mathrm{cos}^1(\mathrm{sin}(\delta )\mathrm{sin}(\delta _i)+\mathrm{cos}(\delta )\mathrm{cos}(\delta _i)\mathrm{cos}(\alpha \alpha _i))$$ (1) . If the 1 $`\sigma `$ uncertainty in the annulus width is $`\sigma _i`$, then $$\chi ^2=\underset{i}{}\frac{d_i^2}{\sigma _i^2}.$$ (2) The assumed source position is varied to obtain a minimum chisquare; 1, 2, and 3 $`\sigma `$ equivalent confidence contours in $`\alpha `$ and $`\delta `$ are found by increasing $`\chi _{min}^2`$ by 2.3, 6.2, and 11.8. We have tested this method on six IPN annuli for SGR1900+14 (Table 2), whose precise (sub-arcsecond) location is known from VLA observations of a particle outburst (Frail, Kulkarni, and Bloom 1999) following the giant flare of 1998 August 27 (Hurley et al. 1999a). The result is shown in figure 1. The 3$`\sigma `$ error ellipse has an area of $``$600 arcsec<sup>2</sup>, and the best fitting position for the SGR, at $`\alpha (2000)=19^\mathrm{h}07^\mathrm{m}14.3^\mathrm{s},\delta (2000)=9^\mathrm{o}19\mathrm{}19\mathrm{}`$, has a $`\chi ^2`$ of 1.05 for 4 degrees of freedom (six annuli, minus the two fitting parameters $`\alpha ,\delta `$). It lies $`0.6\mathrm{}`$ from the VLA position. The results of applying the method to SGR1806-20 are shown in figure 2. The best fit position is at $`\alpha (2000)=18^\mathrm{h}08^\mathrm{m}39.4^\mathrm{s},\delta (2000)=20^\mathrm{o}24\mathrm{}38.6\mathrm{}`$, with a $`\chi ^2`$ of 3.35 for 6 degrees of freedom (8 annuli minus two fitting parameters). It lies $``$15 ″from the center of the non-thermal core, and well outside it. The 3$`\sigma `$ error ellipse has an area of $``$ 230 arcsec<sup>2</sup>, making it the smallest burst error box determined to date (the 324 arcsec<sup>2</sup> error box of the 1979 March 5 burst was, until now, “the most precisely determined gamma-ray source error box in existence” - Cline et al. 1982). The position of the non-thermal core has a total $`\chi ^2`$ of 101. ## 3 Accuracy of the Method Since each individual annulus gives, in effect, an underdetermined source position, it is possible in principle that unknown systematic errors might affect the location accuracy. For example, timing errors of 96 to 206 ms in the Ulysses data could shift the positions of the annuli by different amounts and make them all consistent with that of the non-thermal core. Apart from the unlikely combination of errors which this would require (i.e., each annulus would have to be subject to a different error in such a way as to make the erroneous best fit position have an acceptable $`\chi ^2`$), there are several independent confirmations of the accuracy of the triangulation method. The first is the excellent agreement between the VLA and triangulated positions of SGR1900+14. The second is the agreement between IPN positions and the positions of gamma-ray bursts with optical counterparts (e.g. Hurley et al. 1997). The third and most stringent, however, is the confirmation of the Ulysses spacecraft timing and ephemeris by end-to-end timing tests, in which commands are sent to the GRB experiment at precisely known times, and the times of their execution onboard the spacecraft are recorded and compared with the expected times. Because of command buffering on the spacecraft, there are random delays in the execution of these commands, and the timing is verified to different accuracies during different tests. However, the tests before, during, and after the eight bursts in Table 1 took place on 1996 October 1, 1997 February 19, 1997 August 25, 1998 February 18, 1998 August 21, and 1999 March 7, and indicated that the timing errors at those times could not exceed 19, 21, 39, 29, 112, and 1 ms respectively. For comparison, the 3 $`\sigma `$ uncertainties in these triangulations have been taken to be 125 ms. This includes both the statistical errors, and a conservative estimate of unknown timing and spacecraft ephemeris errors. The low $`\chi ^2`$ values for the two SGR positions are probably due in part to this estimate. Thus the most likely explanation of our results is indeed that SGR1806-20 is not in the non-thermal core of G10.0-0.3, as has been assumed up to now. In this respect, SGR1806-20 resembles SGR1627-41, which also displays a significant displacement from the core of its radio SNR (Hurley et al. 1999b). ## 4 Discussion The association of van Kerkwijk et al.’s (1995) possible LBV with the SNR is compelling; they estimate that there are only several hundred stars this luminous in the galaxy, and this one lies within 1 $`\mathrm{}`$ of the radio peak. Indeed, in assuming a distance to the object of 6 kpc, they may have underestimated the star’s luminosity; a better distance estimate is now 14.5 kpc (Corbel et al. 1997), giving a luminosity of $`6\times 10^6\mathrm{L}_{}`$. The fact that this object has not yet been observed to vary is not an argument against the LBV identification: Humphreys & Davidson (1994) note that LBV’s do not always appear blue or variable. They are simply very luminous, unstable hot supergiants which undergo irregular eruptions. In a giant eruption, they may radiate as much luminous energy as a supernova, and eject a solar mass of material. But this does not explain the SGR bursts, the changing radio morphology, or the displacement between the radio core and the source of the bursts. We propose that the LBV drives the morphological changes. LBV’s are characterized by sporadic mass loss rates of up to $`10^4\mathrm{M}_{}/\mathrm{y}`$ (Humphreys & Davidson 1994) and more. Moreover, these flows may be bipolar or jetlike, as in the case of $`\eta `$ Car or P Cygni (Meaburn, Lopez & O’Connor 1999). Van Kerkwijk et al’s (1995) measurements of the possible LBV in G10.0-0.3 indicate an outflow velocity of 500 km/s. Coupled with a mass loss rate of $`10^4\mathrm{M}_{}/\mathrm{y}`$, this gives a total wind energy of $`2.5\times 10^{44}\mathrm{erg}/\mathrm{y}`$, or a factor of $``$30 greater than the rate of energy deposition into the radio nebula by the neutron star in the model of Frail et al (1997). Thus the LBV is easily capable of supplying the energy to explain the changing radio morphology. In the case of $`\eta `$ Car, the LBV not only changes the morphology of its radio nebula dramatically, but it also powers the (apparently non-thermal) radio nebula (Duncan et al. 1995; Duncan, White, & Lim 1997). We believe that the magnetar model is the best current explanation for the bursts. It is possible that the SGR is not associated with the radio nebula, and that we are simply observing a chance alignment of the two. But if the two are indeed associated, the SGR and the LBV may once have formed a binary system, which became unbound following the supernova explosion. In the magnetar model, the neutron star may be born with a kick velocity $`>`$1000 km/s (Duncan & Thomson 1992). If we assume that the distance to SGR1806-20 is 14.5 kpc (Corbel et al. 1997), that its age is 10,000 y, and that the neutron star originated at the position of the LBV/non-thermal core, its approximate transverse velocity is a rather modest 100 km/s. (This estimate is subject to large uncertainties due to the unknown age of the SNR; also, the actual space velocity could be much larger). This certainly does not strain the magnetar model, but it does raise another interesting question. Why did the SGR progenitor form a neutron star rather than a black hole, given that it must have been very massive to end its life earlier than the LBV? In any case, SGR1806 now appears to be similar to the other SGRs in that there is no associated radio emission at its position, except for the brief radio flare from SGR1900+14 (Frail et al. 1999). KH is grateful to JPL for Ulysses support under Contract 958056, and to NASA for Compton Gamma-Ray Observatory support under grant NAG 5-3811. We thank the referee, M. van Kerkwijk, for helpful comments, and P. Li for his analysis of the ROSAT data.
no-problem/9906/nucl-th9906058.html
ar5iv
text
# THE SHORT-RANGE BARYON-BARYON INTERACTION IN A CHIRAL CONSTITUENT QUARK MODEL ## I Introduction An intricate question of the intermediate energy nuclear physics is about the origin of the short-range repulsion in the nucleon-nucleon ($`NN`$) system or, more generally, in baryon-baryon systems. By now it is clear that the mechanism describing the $`NN`$ interaction should be related with the QCD dynamics responsible for the low-energy properties of the nucleon. It is also now evident that the most important QCD phenomenon in this case is the spontaneous breaking of chiral symmetry, which implies that at momenta below the chiral symmetry breaking scale the relevant quasiparticle degrees of freedom are constituent quarks and pseudoscalar mesons, which are Goldstone bosons of the broken chiral symmetry. Assuming that in this regime the dominant interaction between confined constituent quarks is due to Goldstone boson exchange (GBE) one can understand the structure of the whole low-lying baryon spectrum . A GBE interaction between constituent quarks is a natural interpretation of the t-channel iterations of point-like gluonic interactions between quarks which are responsible for dynamical breaking of chiral symmetry in QCD vacuum . In a previous work it was shown that the short-range part of the flavor-dependent spin-spin force between constituent quarks, which is due to GBE, and which is reinforced by the short-range part of vector-meson exchange (correlated two-GBE ), induces a strong short-range repulsion in the $`NN`$ system. The same interaction also implies a strong short-range repulsion in the 6Q system with “$`H`$-particle” quantum numbers , suggesting that a very existence of a deeply bound $`H`$-particle is impossible within the given picture. In the present work we extend our analysis to other flavor octet-octet B=2 systems and show that the short-range repulsion persists as a general case. In section II we revisit the $`NN`$ interaction and discuss in which respects the predictions of depend on the radial form of the short-range flavor-spin interaction. Section III is devoted to the classification of 6Q states relevant for the hyperon-nucleon ($`YN`$) and hyperon-hyperon ($`YY`$) systems and to a qualitative estimate of the short-range repulsion appearing in these systems. ## II The $`NN`$ interaction at short range - revisited The results of Ref. are based on the flavor-spin hyperfine interaction between two constituent quarks $`i`$ and $`j`$ which has a short-range part of the form $$\lambda _i^F\lambda _j^F\stackrel{}{\sigma }_i\stackrel{}{\sigma }_j,$$ (1) where $`\lambda ^F`$ are the flavor Gell-Mann matrices and an implicit summation over $`F`$ =1,2,…8 is understood. The operator (1) represents the short-range part of the Goldstone boson exchange interaction . Two correlated Goldstone bosons (vector meson exchange) enhance the effect of the short-range part of the one-boson exchange interaction, as shown in , so it can be incorporated in (1) as well. In Ref. it was shown that such a flavor-spin interaction leads to a short-range repulsion in the $`NN`$ interaction when the latter is treated as a 6Q system. This repulsion results from the fact that the energy of the most favourable compact 6Q configuration, $`s^4p^2[42]_O[51]_{FS}`$, is much above the two-nucleon threshold. Here and below $`[f]_O,[f]_{FS}`$, and $`[f]_F`$ denote Young diagrams specifying the permutational orbital (O), $`SU(6)_{FS}`$ (FS), and $`SU(3)_F`$ (F) symmetries and it is always assumed that the center-of-mass motion is removed from the shell model configurations. In the nonrelativistic Hamiltonian with the parametrization of the hyperfine interaction of Ref. has been used. That parametrization, being successful for baryon spectra, may lead, however, to some undesirable effects in the $`NN`$ system because it contains the shift parameter $`r_0=0.43`$ fm of the short-range hyperfine interaction from the origin: $$4\pi \delta (\stackrel{}{r}_{ij})\frac{4}{\sqrt{\pi }}\alpha ^3\mathrm{exp}(\alpha ^2(rr_0)^2).$$ (2) This shift enhances the quark-quark matrix elements with $`1p`$ relative motion and affects the coupling between the symmetry states chosen for the the diagonalization of the Hamiltonian. Here we try to find out to which extent the results of Ref. are modified when one takes $`r_0`$ = 0. Actually the baryon spectra can be described as well without such a shift with a properly chosen parametrization . To estimate the strength of the short-range $`NN`$ interaction we use the adiabatic (Born-Oppenheimer) approximation $$V_{NN}(R)=H_RH_{\mathrm{}},$$ (3) where $`R`$ is a generator coordinate, defined as the distance between two harmonic oscillator wells, each associated asymptotically to a $`3Q`$ cluster. In Eq. (3) $`H_R`$ is the lowest eigenvalue resulting from the diagonalization of the Hamiltonian at fixed $`R`$. Presently we are interested in the effective potential at $`R`$ = 0 only. In this case in the S-wave relative motion only the $`s^6[6]_O`$ and $`s^4p^2[42]_O`$ shell-model configurations are allowed . So to estimate the strength of the repulsion we diagonalize the Hamiltonian in the basis of the following four most important configurations : $$\begin{array}{ccc}|1>& =& |s^6[6]_O[33]_{FS}>\\ |2>& =& |s^4p^2[42]_O[33]_{FS}>\\ |3>& =& |s^4p^2[42]_O[51]_{FS}>\\ |4>& =& |s^4p^2[42]_O[411]_{FS}>.\end{array}$$ (4) The Hamiltonian and the confinement potential are taken from the ref. . Here we modify the radial dependence (see eqs. (17)-(18) of ref. ) of the meson-exchange potential as follows: (i) we drop the long-range Yukawa part $`\mu _\gamma ^2\mathrm{exp}(\mu _\gamma r_{ij})/r_{ij}`$ of the potential, as it contributes very little at short-range in the 6Q system; (ii) in the short-range part of the hyperfine interaction we put $`r_0`$ = 0; (iii) we keep the ratio of the singlet ($`g_0`$) to octet ($`g_8`$) coupling constants as in Refs. but readjust the absolute value of each coupling constant in order to reproduce the $`\mathrm{\Delta }N`$ mass splitting keeping in mind the modifications (i) and (ii). The root-mean-square matter radius $`\beta `$ of the $`s^3`$ nucleon and $`\mathrm{\Delta }`$, which coincides with the harmonic oscillator parameter in the 6Q basis, is obtained from the nucleon stability condition $$\frac{}{\beta }N|H|N=0.$$ (5) This procedure gives $`g_8^2/4\pi =2.11`$, which should be regarded as an effective coupling constant simulating the combined effect of both the pseudoscalar- and the vector-meson-like exchange short-range interaction (1) with a nonrelativistic $`s^3`$ ansatz for the baryon. The minimum of $`m_N`$ = $`N|H|N`$ = 1.324 GeV is achieved at $`\beta `$=0.373 fm. The absolute value of the “nucleon mass” $`m_N`$ is unimportant in the present context as it identically cancels out in (3). The calculated $`m_N`$ can also be shifted to a physical value by simply adding a constant contribution to the effective confining interaction. The results of the diagonalization for the $`{}_{}{}^{3}S_{1}^{}`$ and $`{}_{}{}^{1}S_{0}^{}`$ NN partial waves are shown in Tables I and II, which should be compared with Tables II and III of Ref. or more precisely to Tables V and VI of Ref. because the latter tables show the diagonalization results of a 4 $`\times `$ 4 matrix, as here. In Ref. the 5th basis vector used in has also been removed because it has no correspondence in the molecular orbital basis employed there. We see that with the present parametrization of the short-range QQ interaction the $`s^4p^2[51]_{FS}`$ configuration is again the lowest one with an energy of roughly 1 GeV below the energy of the orbitally unexcited configuration $`s^6[33]_{FS}`$. In this sense the conclusion of Ref. is reconfirmed. The lowest eigenvalue is about 1.4 GeV above the $`2m_N`$ threshold, which shows that the strong short-range repulsion in the $`NN`$ system persists with this new parametrization. However, the numerical values of the off-diagonal matrix elements are now different compared to those obtained with the parametrization of Refs. or . As a consequence the amplitude of the configuration $`s^4p^2[51]_{FS}`$ is somewhat smaller and one finds more mixing among the configurations $`|1|4`$ in the lowest state eigenvector, in contrast to Refs. or . But the configuration $`s^4p^2[51]_{FS}`$ remains dominant in the lowest state and it can induce additional effective repulsion as discussed in (see also ). While the mixing does not affect the conclusion about the repulsive core, it is important for the behaviour of the 6Q wave function at short range. When one projects this wave function on the $`NN`$ channel according to the procedure described in the node, predicted there (see Fig. 1), nearly disappears. Now the behaviour of the projection is similar to the $`NN`$ wave function obtained with usual repulsive core potentials. This behaviour comes from the destructive interference of the excited $`s^4p^2`$ and nonexcited $`s^6`$ configurations. Such a destructive interference has been observed earlier in other microscopical models and can be considered as a substantiation of the repulsive core in $`NN`$ potentials. However, contrary to any simple $`NN`$ potential model, the 6Q wave function is much richer and contains not only the $`NN`$ component, but also a variety of other components such as $`NN^{},N^{}N^{},\mathrm{\Delta }\mathrm{\Delta },\mathrm{\Delta }\mathrm{\Delta }^{},\mathrm{}`$ . Next we adress the question to which extent the height of the repulsion in the $`NN`$ system is sensitive to the contribution of the flavor-singlet $`\eta ^{}`$-like exchange. The $`\eta ^{}`$-like exchange tends to decrease the $`\mathrm{\Delta }N`$ splitting, while the $`\pi `$-like exchange works just in opposite direction. It means that in reproducing the $`\mathrm{\Delta }N`$ splitting the octet coupling constant will become smaller once the $`\eta ^{}`$-exchange interaction is dropped. We take the extreme limit $`g_0^2/4\pi =0`$ and repeat the steps (i)-(iii) from above. One obtains $`\beta `$ = 0.522 fm, $`g_8^2/4\pi =1.29`$ and $`m_N=1.4657`$ GeV. The results for the $`NN`$ system are given in Tables III and IV. Comparing them with those of Tables I and II we conclude that the height of the repulsive core is essentially smaller than before and the structure of the ground state eigenvector is changed. In the present case the configurations $`s^6[33]_{FS}`$ and $`s^4p^2[51]_{FS}`$ become approximately degenerate and still about 900 MeV above the $`2m_N`$ threshold. The lowest eigenvalue is about 600 MeV above the threshold, showing that the strength of the repulsion is reduced compared to the previous case, but still important. The calculation of the $`NN`$ interaction at short range within the adiabatic approximation above should be taken with some caution, however. In fact it represents only the diagonal kernel of a dynamical treatment such as the resonating group method (RGM). So an ultimate conclusion could only be drawn from the behavoiur of the $`NN`$ phase shifts calculated beyond the adiabatic approximation. Such phase shifts, obtained within an extended resonating group method, do indicate the presence of a very strong repulsion even in the case without any $`\eta ^{}`$-exchange interaction . ## III Short-range 6Q configurations in hyperon-nucleon and hyperon-hyperon systems In this section we discuss the issue whether or not the short-range repulsion in the $`NN`$ system, implied by the flavor-spin hyperfine interaction (1), persists in other baryon-baryon (flavor octet-octet) systems. We first construct the lowest possible symmetry states $`|[f]_O[f]_F[f]_{FS}>`$ compatible with the asymptotic two-baryon channels. For simplicity we consider the $`SU(3)_F`$ limit (see Eq. (6) below). Assuming that the orbital wave function of any of the octet baryons is described by the $`[3]_O`$ permutational symmetry, only two states $`[6]_O`$ and $`[42]_O`$ are allowed in the S-wave relative motion of a two-baryon system . Applying the inner product rules of the symmetric group both for $`[f]_O\times [f]_C`$, where $`[f]_C=[222]_C`$ is fixed by the color-singlet nature of the two-baryon system, and $`[f]_F\times [f]_S`$, where $`[f]_S`$ is fixed by the total spin $`S`$ of the 6Q system, one arrives at the symmetry states listed in Tables V and VI. We recall that for a given $`[f]_S`$ of $`SU(2)`$ where $`f_1`$ and $`f_2`$ represent the number of boxes in the first and second rows of the Young diagram $`[f]_S`$ respectively, with $`f_1+f_2=6`$ in the present case, the spin is given by $`S=1/2(f_1f_2)`$. When one considers a schematic model , where the interaction Hamiltonian is approximated as $$H_\chi =C_\chi \underset{i<j}{}\lambda _i^F\lambda _j^F\stackrel{}{\sigma }_i\stackrel{}{\sigma }_j,$$ (6) and the constant $`C_\chi `$ = 29.3 MeV is extracted from the phenomenological $`\mathrm{\Delta }N`$ splitting, then the expectation value of the interaction (6) for all symmetry states in Tables V and VI can be evaluated through the Casimir operator eigenvalues, see Appendix A of ref. . The results are given in Tables V and VI in units of $`C_\chi `$. One can immediately conclude from these Tables that far the lowest state is $`|s^4p^2[42]_O[321]_F[51]_{FS}>`$. This state is not allowed in the $`NN`$ system. Here it appears due to the presence of three distinct flavors $`u`$, $`d`$ and $`s`$. Its contribution is lower than that of the vector $`|3>`$ of the list (4) by about 10 $`C_\chi `$ units (recall that in the $`NN`$ case the flavor symmetry is $`[42]_F`$ for $`S=0`$ S-wave and $`[33]_F`$ for $`S=1`$ S-wave). By itself the $`|s^4p^2[42]_O[321]_F[51]_{FS}>`$ state guarantees that a strong effective repulsion will persist in the $`YN`$ and $`YY`$ systems, again related to a specific symmetry structure of the type $`s^4p^2[51]_{FS}`$. The ground state configuration $`s^6[6]_O`$ becomes even more “forbidden” by dynamics than in the $`NN`$ system (i.e. the weight of $`s^6[6]_O`$ should be expected to become smaller). This effect, however, cannot be obtained within the simple approximation considered below where only the expectation value of the $`|s^4p^2[42]_O[321]_F[51]_{FS}>`$ state is calculated. To have a rough qualitative idea about the strength of the interaction at short range in $`YY`$ and $`YN`$ systems we thus calculate the diagonal matrix element $$<s^4p^2[42]_O[321]_F[51]_{FS}|H_0+H_{conf}+H_\chi |s^4p^2[42]_O[321]_F[51]_{FS}>$$ (7) and compare it with the two-baryon threshold. With the coupling constant $`g_8^2/4\pi =2.11`$, and the ratio $`g_0^2/g_8^2`$, fixed in the previous section, we now use a harmonic oscillator parameter $`\beta =0.403`$ fm, which provides an equilibrium value for $`\mathrm{\Lambda }`$, as it follows from Table VII. In all cases we describe the kinetic energy of a $`6Q`$ system in a simple way $$<s^4p^2|H_0|s^4p^2>=\frac{19}{4}\mathrm{}\omega ,$$ (8) $$\mathrm{}\omega =\frac{\mathrm{}^2}{m_{ave}}\beta ^2.$$ (9) where $`m_{ave}`$ is an average quark mass defined for each system and the center-of-mass motion is removed. For example the $`\mathrm{\Lambda }\mathrm{\Lambda }`$ system has an average mass $`m_{ave}=(4m+2m_s)/6`$, where $`m`$ = 0.340 GeV and $`m_s`$ = 0.440 GeV. All the contributions from $`H_\chi `$ and $`H_{conf}`$ are calculated with the help of the fractional parentage technique, similar to and described in detail in Ref. . The $`SU(3)`$ Clebsch-Gordan coefficients for the flavor part of the wave function are taken from Ref. . By this technique one can reduce the six-quark matrix elements to linear combinations of two-quark matrix elements which allow immediate integration in the spin-flavor space by use of Eq. (3.3) of Ref. . In particular, for the confinement part $`H_{conf}`$ the orbital matrix elements can be easily calculated analytically. This gives $$H_{conf}=\frac{71}{6}\sqrt{\frac{2}{\pi }}C\beta .$$ (10) where $`C`$ is the string tension taken from and $`\beta `$ has been specified above. In an analogue way the confinement energy of a ground state baryon is $$H_{conf}=6\sqrt{\frac{2}{\pi }}C\beta .$$ (11) Thus the difference between the $`6Q`$ and two times the $`3Q`$ confinement energy is $`\sqrt{\frac{2}{\pi }}\frac{C\beta }{6}`$. This gives about $`5`$ MeV, which proves that the confinement contribution nearly cancels out in the baryon-baryon potential in the present approximation. Results for $`H_0`$, $`H_\chi `$ and $`H`$ are exhibited in Tables VIII and IX for $`S`$ = 0 and $`S`$ = 1 repectively. We define the separation energy (last column) of either table as the difference between $`H`$ and the lowest threshold two-baryon mass, calculated with the same hamiltonian, associated to a given $`YI`$. The lowest thresholds are $`\mathrm{\Lambda }\mathrm{\Lambda }`$ , $`\mathrm{\Sigma }\mathrm{\Sigma }`$ , $`N\mathrm{\Lambda }`$ and $`\mathrm{\Lambda }\mathrm{\Xi }`$ for $`YI`$ = 00, 01, 1 1/2 and -1 1/2 respectively. The separation energy is roughly interpreted as the value taken by the baryon-baryon interaction potential at zero separation distance in the Born-Oppenheimer approximation. The last column of Tables VIII and IX indicates repulsion of the order of 1 GeV in all cases. When one uses a Hamiltonian without $`\eta ^{}`$-exchange, like at the end of the above section, then the repulsion is weakened by roughly 200 - 300 MeV. ## IV Conclusions This paper closes a series of papers devoted to a qualitative study of the short-range baryon-baryon interactions within a chiral constituent quark model relying on meson-exchange dynamics. Our results indicate that the short-range flavor-spin interaction (1) between constituent quarks implies a strong short-range repulsion in NN and other flavor octet-octet systems. While we observe a strong repulsive core we cannot insist on numerical values for this repulsion because of the use of simplified approximations. The next stage of the study should invoke more detailed and refined dynamical approximations as well as the incorporation of multiple correlated GBE interactions (scalar and vector meson exchanges) in order to provide a realistic description of the middle- and long-range physics in baryon-baryon systems with all the necessary components, like tensor and spin-orbit forces. ## V Acknowledgement L.Ya.G. is indebted to the nuclear theory groups of KEK-Tanashi and Tokyo Institute of Technology for a warm hospitality. His work is supported by a foreign guestprofessorship program of the Ministry of Education, Science, Sports and Culture of Japan.
no-problem/9906/hep-th9906175.html
ar5iv
text
# Light Meson Mass Spectrum from Gauge Fields on Supergravity Backgrounds ## Abstract We propose that the spectrum of light mesons (the $`\pi `$ and $`\rho `$, together with their radial excitations) can be calculated in the limit of vanishing light quark masses by studying gauge theory (open string theory) on suitable higher dimensional background geometries. Using the metric proposed by Witten for glueball calculations as a paradigmatic example, we find a spectrum which is in startlingly good agreement with the masses tabulated by the Particle Data Group . These calculations have only one free parameter, corresponding to the overall mass scale. We make predictions for the next several particles in the spectrum. String theory arose as an attempt to understand the strong interactions. With the advent of gauge theories in the early 1970’s, string theory metamorphosed into a “theory of everything”, and, until recently, has had little new to say about strong interaction phenomenology. The last two years, however, have seen dramatic progress towards understanding the relationship between string theory and gauge theories. Maldacena’s conjecture of the relationship between $`𝒩=4`$ supersymmetric gauge theories in four dimensions and string theory (supergravity) on $`AdS^5\times S^5`$ has proven very fruitful . Witten’s conjectured extension to non-supersymmetric theories has provided the basis for calculations of glueball mass spectra which have been directly compared to the results of lattice calculations with good numerical agreement . Unfortunately, the calculations presented to date correspond to theories which do not contain states with light quarks, and so are not directly relevant to the real world. Indeed, the question as to how to create such models starting with string theory or M-theory is still unclear. There is, however, a qualitative line of reasoning to pursue. QCD flux tubes are presumably the starting point for a stringy picture of the strong interactions. The extra dimensions involved in the Maldacena/Witten conjectures plausibly correspond in some fashion to internal degrees of freedom (such as thickness) of the flux tube. It is thus reasonable that masses of glueballs, which are closed flux tubes in a QCD picture, can be calculated in terms of closed (super)string modes on non-trivial higher dimensional backgrounds, such as those proposed by Witten and others. Mesons, on the other hand, should be associated with the modes of open strings. This correspondance was made in the early days of string theory, and is also suggested by a constituent quark picture in which the quark and antiquark of a meson are connected by a QCD flux tube. Since the ends of open strings move at the speed of light, this picture suggests that we will be modeling states with massless quarks; i.e. chiral symmetry is spontaneously, but not explicitly, broken. If the dynamics of these meson flux tubes are locally the same as those for glueballs, then meson dynamics in the real world should correspond to open string dynamics on the same spaces considered in the glueball calculations. The purpose of this paper is to explore the implications of this suggestion. We do this by studying the mass spectrum associated with the lowest mode of the open (super)string: a massless gauge field $`A_\mu `$. Our calculations closely follow the calculations of the glueball mass, so our description will be brief. We consider the action $$I=\frac{1}{4}d^{10}x\sqrt{g}F_{\mu \nu }F^{\mu \nu }$$ (1) in Witten’s black hole metric $$\frac{ds^2}{l_s^2g_5^2N/4\pi }=\frac{dr^2}{r(1\frac{1}{r^6})}+r^3(1\frac{1}{r^6})d\tau ^2+r^3\underset{i=1}{\overset{4}{}}dx_i^2+rd\mathrm{\Omega }_4^2.$$ (2) Though the interacting theory will be expressed in terms of a non-abelian field strength, we only need the quadratic part of the Lagrangian to calculate masses and so we take $`F_{\mu \nu }=_\mu A_\nu _\nu A_\mu `$. We similarly suppress the isospin indices of the $`\pi `$ and $`\rho `$ fields, and the corresponding indices on the gauge field $`A_\mu `$. We choose a gauge in which $`A_r=0`$, and make an ansatz for the modes of interest: $`A_\tau `$ $`=`$ $`H(r)e^{ikx},`$ (3) $`A_i`$ $`=`$ $`ϵ_i(k)G(r)e^{ikx},`$ (4) where $`ϵ_i`$ defines the polarization of the vector mesons. The equations of motion then imply that $`ϵk=0`$ and that $`(r^61)_r[r^7_rH(r)]k^2r^9H(r)=0,`$ (5) $`_r[(r^7r)_rG(r)]k^2r^3G(r)=0.`$ (6) The physical $`\pi `$ and $`\rho `$, together with their radial excitations, thus correspond respectively to solutions of the equations (5) and (6) which are regular at $`r=1`$ (the horizon of the black hole), and which are normalizable, i.e. fall off sufficiently rapidly as $`r\mathrm{}`$. The masses are given by the corresponding eigenvalues, $`m_i^2=k_i^2`$. An interesting feature of eqn. (5), in contrast to eqn. (6) and the corresponding glueball equations, is that it has an acceptable solution for $`k^2=0`$ given by $`H(r)=A/r^6`$. There is thus a massless pion, as expected from the above general considerations. Other particle masses are found by numerical techniques. We have calculated the series solution about $`r=1`$ (a regular singular point of our equations), and used this to determine the functions and their derivatives at $`r=1.01`$. These values are then used as input to numerically integrate the equations. Since normalizable solutions vanish for large $`r`$, while non-normalizable ones do not, an eigenvalue is signaled by a change in sign of the solution at large $`r`$ as $`k^2`$ is changed. Successive eigenvalues should correspond to successive excitations of the $`\pi `$ (resp. $`\rho `$). We have tested this technique by calculating the glueball spectrum, and reproduce the precise results of to the accuracy quoted here. The overall mass scale is not determined by our equations. We fix this by comparison with the data, minimizing $`_i((m_i^{obs})^2\lambda (m_i^{calc})^2)^2`$ with respect to $`\lambda `$. The result, $`\lambda =.043`$, is not very different if, for example, one minimizes the differences between masses rather than squared-masses. In Table I we tabulate our results together with the observed masses of the corresponding particles. The resulting agreement is surprisingly good. The worst agreement is for the $`\pi (135)`$ and $`\rho (770)`$, and is presumably a consequence of the fact that there is no explicit chiral symmetry breaking in this model. The general pattern is consistent with quark-model expectations: the mass of the $`\rho `$ will be high if the mass of the $`\pi `$ is low, and these effects will be less important at higher masses. (This, incidentally, is why normalizing to the mass of the $`\rho `$ is probably not the best way of comparing the calculations to the data.) It is worth while making a few comments about the states that we have included. The eigenvalues of the equations (5) and (6) correspond, respectively, to masses of successive radial excitations of the $`\pi `$ and $`\rho `$. The $`\rho (1700)`$, which has the same overall quantum numbers as the $`\rho `$, is a D-wave state , and thus an orbital excitation of the $`\rho `$ which should not be reproduced by the present calculation. The $`\rho (2150)`$, which we have included, is presented in the full Meson Particle Listings but is omitted in the Data Summary Table of the Review of Particle Physics. Since we did not include this mode in determining the scale, it is in some sense a prediction of the present calculations. There are no other states in the Meson Particle Listings with quantum numbers such that they should be reproduced by this calculation. We do, however, predict additional radial excitations of the $`\pi `$ at 2.32 GeV and 2.88 GeV and of the $`\rho `$ at 2.64 GeV. Though it may be difficult to isolate such states, it would be extremely interesting to confirm (or deny) their existence. An infinite number of additional states are also predicted, but all have masses larger than 3 GeV. We believe that these results strongly suggest that present string theory technology is directly relevant to strong interaction phenomenology. A number of directions for further work immediately suggest themselves. First, we have not discussed the Kaluza-Klein modes associated with the $`S^4`$ directions. In the glueball case, modifications of the metric (2) corresponding to rotating D-branes decouple the unwanted Kaluza-Klein modes without significantly changing the masses of the physical glueballs . We expect similar results here, but this must be checked. Second, the fact that the pion is massless in this model is extremely interesting. In addition to studying the implications of this, one should also study the impact of adding an explicit mass term to the action (1). Generalizations of this may permit the extension of this model to $`K`$ and $`K^{}`$ mesons, and perhaps others. Beyond this, it may be possible to study interactions in this model. Since baryons can already be understood as topological excitations of pions in Skyrme models , perhaps a similar understanding is possible here. Finally, other modes of the open string should be studied; they should correspond to other mesons, including orbital excitations of the $`\pi `$ and $`\rho `$. We hope to report on some of these issues in the near future. We would like to thank Mark Trodden for helpful conversations, and Dan Akerib, Zlatko Dimcovic, Glenn Starkman and Tanmay Vachaspati for readings of the manuscript. We would also like to thank Juan Maldacena, Jeffrey Harvey and Antal Jevicki for useful background discussions. This work was supported in part by the National Science Foundation through grant PHY-9940415.
no-problem/9906/hep-th9906208.html
ar5iv
text
# THE DECONFINING PHASE TRANSITION AS AN AHARONOV-BOHM EFFECT 11footnote 1To appear in the proceedings of the workshop “Understanding Deconfinement in QCD”, Trento, Italy, March, 1999. ## 1 Introduction The first strong hints of the deconfined quarks at high temperature appeared more than ten years ago , the numerical confirmation followed soon. Subsequently a large number of details has been clarified but the driving force of the deconfinement transition, the confinement-deconfinement mechanism, remained elusive. A subjective and a rather sketchy list of remarks is presented here to indicate a unique and challenging aspect of this phase transition. Only one detail will be discussed in a slightly more detailed manner, the formal similarity between the density matrix of the Aharonov-Bohm (A-B) system and of QCD. One can distinguish two confinement mechanisms , a hard and a soft one. The hard mechanism is responsible for the linear potential between static test quarks in the absence of dynamical quarks. The soft, low energy mechanism is which screens a test quark and was thought by V. Gribov to be similar to the supercritical vacuum of QED. More precisely the infrared instability of the perturbative QCD, the source of the hard confinement mechanism, leads to strong gluon interactions at large distances. Sufficiently far from a test quark the coupling constants reaches a large enough value to ignite the spontaneous creation of the quark-anti quark pairs which in turn shield the test quark charge. Most of the remarks mentioned here refers to the hard confinement mechanism which is more elementary and should be clarified before embarking the study of the soft mechanism of full QCD. ## 2 Unusual or unique features 1. Different degrees of freedom: We find different degrees of freedom at the two sides of the phase transition. This happens in a number of other phase transitions, the Mott or the localisation-delocalisation transitions may serve as examples. Observe that the elementary degrees of freedom are recovered among the highly excited states in these cases. This does not happen in the hadronic phase. The relevance of this obvious remark becomes clear by considering the thermal average of an observable $`A`$, $`A=Z^1_nn|A|ne^{E_n/T}`$. In order to reproduce the thermal averages we use the color singlet asymptotic states $`|n`$ of the hamiltonian of the strong interactions. How can we recover the contributions of an isolated, deconfined quark to the given observable? The only way out of this problem is the modification of the Hilbert space at the phase transition. 2. Weakly or strongly coupled phase? The asymptotically free running coupling constants becomes small around the typical energy scale $`p=T`$ at high energies, $`T>\mathrm{\Lambda }_{QCD}T_c`$. Does that mean that the deconfined phase is weakly coupled at high enough temperature? The answer is known to be negative since long time . The small parameter of the perturbation expansion at high temperature stems from a non-perturbative quantity, the magnetic screening mass. This can be understood by recalling that the thermal bath breaks the Lorentz invariance. Though the typical energy scale is pushed up at high temperature $`ET`$, the (off-shell) momentum scale in the loop integrals is effected differently by the temperature and the infrared stabilization of the long wavelength modes remains a difficult question. This is because the partition function of the high temperature $`3+1`$ dimensional QCD can be approximated by a $`3`$ dimensional (classical) Yang-Mills-Higgs system and the infrared sensitivity of the partition function increases by lowering the dimension. Thus the fate of the perturbation expansion which is based on massless gluons depends on the screening mechanism. The usual strategy of dealing with the IR divergences, the separation of the scales $`T`$, $`gT`$ and $`g^2T`$, can not solve this problem because $`g`$ does not reach small enough values, $`g(m_{Planck})1/2`$. 3. Order parameter: The order parameter related to the hard confinement mechanism is the trace of the heavy quark propagator continued over complex time, $$\omega (\stackrel{}{x},t)=0|\psi _\alpha (\stackrel{}{x},t+\frac{i}{T})\overline{\psi }_\alpha (\stackrel{}{x},t)|0.$$ (1) In the high temperature phase where the time extent of the Euclidean space-time is shorter than the correlation length, $`1/T<\xi \mathrm{\Lambda }_{QCD}^1`$, the gluon field variables are correlated along the world line of the heavy quark and the order parameter develops a non-vanishing expectation value. A distinguishing feature of the deconfining transition is that its order parameter is not a canonical variable. It controls the symmetry with respect the global center <sup>2</sup><sup>2</sup>2The center $`C(G)`$ of the group $`G`$ is a subgroup of $`G`$. It consists of the elements which commute with $`G`$, $`[C,G]=0`$, e.g. $`C(SU(N))=Z_N`$. gauge transformations performed at the initial or the final state of a transition amplitude. It is important to keep in mind that the center of the global gauge transformations is the fundamental group of the gluonic configuration space , $`Z_3=\pi _1(SU(3)/Z_3)`$ <sup>3</sup><sup>3</sup>3Consider the gauge transformation $`\stackrel{}{A}(\stackrel{}{x})g(\stackrel{}{x})(\stackrel{}{}+\stackrel{}{A}(\stackrel{}{x}))g^{}(\stackrel{}{x})`$ acting on the anti-hermitean gauge field in the temporal gauge. The global gauge transformations which commute with other gauge transformations leave $`\stackrel{}{A}(\stackrel{}{x})`$ invariant.. The only other known dynamical breakdown of the fundamental group symmetry is the liquid-droplet quantum phase transition. 4. Finite volume effects: The ratio of the gluonic partition functions with and without a static quark is given by the expectation value of the order parameter, $`e^{(F_qF)/T}=\omega `$. Since the spontaneous symmetry breaking does not occur in a finite system, $`\omega =0`$ and the static quarks always appear confined, $`F_q=\mathrm{}`$, in finite volume. Where does the singular free energy density, $`F_q/V=\mathrm{}`$, come from? This problem is solved by taking into account the destructive interference between the homotopy classes in the gluonic configuration space. 5. Symmetry breaking by the kinetic energy: The spontaneous symmetry breaking mechanism is operating at low energy where the order parameter is driven to a non-symmetrical value due to the degenerate minima of the potential energy. The kinetic energy might drive a spontaneous, or more precisely dynamical symmetry breaking at high energies. The dynamical breakdown of the center symmetry results from such a mechanism . This can be understood by inspecting a quantum top, the baby version of the $`SU(2)`$ Yang-Mills model. The configuration space which consists of the $`3\times 3`$ orthogonal matrices, $`\{R\}=SO(3)=SU(2)/Z_2`$, is doubly connected and the wave functions are single and double valued in the integer (gluons) and the half-integer (quarks) spin subspaces, respectively. Consider now the transition amplitude $`𝒜(R^{},R)=R^{}|e^{itH/\mathrm{}}|R`$ as the function of the final state $`R^{}`$. Since an orientation of the top is undistinguishable from its $`2\pi `$ rotated copy the integer spin amplitude is doubly degenerate on the covering space $`SU(2)`$, $`𝒜(r^{},r)=𝒜(r^{\prime \prime },r)`$ where the final points $`r^{},r^{\prime \prime }SU(2)`$ differ in a rotation by $`2\pi `$, $`r^{}=r^{\prime \prime }`$ (center symmetry). Suppose that $`r^{}`$ is closer to the initial point $`r`$ than $`r^{\prime \prime }`$. Then the kinetic energy tends to suppress the propagation to $`r^{\prime \prime }`$ if the time available for the propagation is short (high temperature or energy). The result for an infinite top whose coordinate $`r`$ influences infinitely many degrees of freedom (global gauge transformations) is that the propagation to $`r^{\prime \prime }`$ is totally suppressed (center symmetry breakdown). The confinement can be understood as the destructive interference in the quark propagator between the different homotopy classes. In fact, the center symmetry of the pure gluon system yields identical amplitudes in different homotopy classes. But a particle in the fundamental representation of the gauge group $`SU(N)`$ propagating along the system picks up the phases $`e^{2i\pi n/N}`$, $`n=1,\mathrm{},N`$ which add up to zero. The result is the absence of these particles in the final states. We find here another characteristic feature of the deconfinement transition: it corresponds to a transition amplitude rather than to the vacuum. This is the key to find a synthesis between the high and the low energy scattering experiments, described in terms of the partons and the hadronic bound states, respectively. In other words, as the time of a collision process is shortened the transition matrix elements go over the “deconfined”, center symmetry broken phase and the elementary constituents (partons) appear. 6. Permanent confinement of triality : (i) The deconfining phase transition consists of the dynamical breakdown of the Gauss’ law and the modification of the Hilbert space for gluons, $$H=\{\begin{array}{cc}H_0\hfill & T<T_c,\hfill \\ H_0H_1H_1\hfill & T>T_c\text{,}\hfill \end{array}$$ (2) where the subscript stands for the triality, the center charge <sup>4</sup><sup>4</sup>4The wave functional $`\mathrm{\Psi }[\stackrel{}{A}(\stackrel{}{x})]H_{\mathrm{}}`$ changes by the phase factor $`e^{2i\pi \mathrm{}n/3}`$ when the global center gauge transformation $`e^{2i\pi n/3}`$ is performed on $`\stackrel{}{A}(\stackrel{}{x})`$. The multi-valued nature of the wave functional is to keep track of the global center gauge transformations, the elements of the fundamental group of the gluonic configuration space which are represented in a trivial manner on the gluon field.. Such a description of the phase transition is the resolution of the puzzle mentioned in point 1. (ii) The triality is permanently confined at any temperature. The deconfined quark seen in the numerical simulation is actually a composite particle containing a quark and its vacuum polarization cloud. The latter has a multi-valued wave functional in such a manner that the total (quark plus gluon) wave functional is single valued. The triality charge of the quark is screened by the unusual gluon state. (iii) The color-magnetic monopoles relate the rotations in the external and the color spaces. These monopoles acquire a half-integer spin in the gluonic states with multi-valued wave functional, a manner similar to the generation of the spin for skyrmions. The unusual gluonic screening cloud is the sum of states with odd and even number of monopoles. These components correspond to fermionic and bosonic exchange statistics. Thus the state of a deconfined quark is the sum of components with bosonic (odd number of monopoles) and fermionic (even number of monopoles) properties. The breakdown of the center symmetry leads to the mixing of the fermi and bose statistics for the deconfined quarks. 7. Triality-canonical ensemble: The transition between the canonical and the grand-canonical ensembles requires smooth enough dependence on the density. Due to the confinement mechanism the formal energy density diverges for non-integer baryon numbers, or non-vanishing triality charges (point 4.). It turns out that the triality-canonical ensemble predicts different center domain structure at the deconfining phase transition than the usual grand-canonical ensemble . This may happen because the center symmetry is broken spontaneously by the quark-anti quark see for $`T<T_c`$ and dynamically by the kinetic energy for $`T>T_c`$ in the canonical ensemble. This furthermore means that the formal center symmetry is preserved in the presence of dynamical quarks and the results mentioned in this talk remain valid in the triality-canonical ensemble with dynamical quarks. ## 3 Density matrix for the A-B system and for gluons A-B system: Consider a charged particle moving on the unit circle in periodic gauge where the wave function is periodic, $`\psi (\varphi +2\pi )=\psi (\varphi )`$. The hamiltonian is $`H=(i_\varphi \mathrm{\Theta }/2\pi )^2/2`$, where $`\mathrm{\Theta }=2\pi A_\varphi `$ stands for the magnetic flux of the circle. The eigenstates and the eigenvalues are $`\psi _n(\varphi )=e^{in\varphi }`$, and $`E_n=(n\mathrm{\Theta }/2\pi )^2/2`$, respectively. The density matrix is given by $`\rho (\alpha ,\beta )=Z^1_ne^{in(\alpha \beta )(n\mathrm{\Theta }/2\pi )^2/2T}`$, where $`Z`$ is the partition function, $`Z=_ne^{(n\mathrm{\Theta }/2\pi )^2/2T}`$. Notice that the probability density $`p(\varphi )=\rho (\varphi ,\varphi )`$ is real non-negative, as it should be. The periodicity of the wave functions gives $`\rho (\alpha ,\alpha +2\pi )=\rho (\alpha ,\alpha )`$. Let us go into an aperiodic gauge by performing the transformation $`\psi (\varphi )e^{i\varphi \mathrm{\Theta }/2\pi }\psi (\varphi )`$. The hamiltonian is simpler, $`H_\varphi ^2/2`$, but has the same spectrum as before because the wave functions are multi-valued, $`\psi (\varphi +2\pi )=e^{i\mathrm{\Theta }}\psi (\varphi )`$. In particular, the eigenvectors are $`\psi _n(\varphi )=e^{i\varphi (n\mathrm{\Theta }/2\pi )}`$. The density matrix transforms as $`\rho (\alpha ,\beta )e^{i(\alpha \beta )\mathrm{\Theta }/2\pi }\rho (\alpha ,\beta )`$, and becomes multi-valued, as well, $`\rho (\alpha ,\alpha +2\pi )=e^{i\mathrm{\Theta }}\rho (\alpha ,\alpha )`$ which makes the construction of the probability density non-trivial. In fact, the choice of different Riemann-sheets for the two coordinate variables yields complex probability and partition function. But notice that the complex factor is the same for each contribution, $$Z_{compl}=𝑑\varphi \rho (\varphi ,\varphi +2\pi )=e^{i\mathrm{\Theta }}𝑑\varphi \rho (\varphi ,\varphi ),$$ (3) and the imaginary part of the entropy is an overall constant which does not influence the thermalization and thermodynamics can be applied. QCD: A similar argument can easily be constructed for gluons yielding the following results: (i) The multi-valued nature of the gluonic wave functional of a deconfined quark is shown by the possible non-positive or complex expectation value of the order parameter $`\omega `$, the partition function of a quark. (ii) The density matrix for gluons and a deconfined quark is multi-valued as it happens for the A-B system in the aperiodic gauge. The change of the Riemann-sheet, $`\rho (\alpha ,\alpha )\rho (\alpha ,\alpha +2\pi )`$, corresponds to the center transformation. (iii) Thus the complex part of the free energy and the entropy of a deconfined quark is a simple kinematical constant which agrees for each contribution to the partition function and does not influence the thermalization and the applicability of the rules of thermodynamics. (iv) The complex part of the deconfined quark entropy may lead to observable effects in the triality-canonical ensemble which is more realistic than the grand-canonical one.
no-problem/9906/hep-th9906179.html
ar5iv
text
# NS Fivebranes in Type 0 String Theory Work supported by the European Commission TMR programme ERBFMRX-CT96-0045. ## 1 Introduction Type 0 string theories have recently attracted a lot of attention. They have been used to study the non-supersymmetric field theories living on their D-branes . Moreover, Bergman and Gaberdiel have conjectured that they are supersymmetry breaking orbifolds of M-theory, thus including them in the web of string dualities. For instance, certain circle compactifications of type 0 and type II are proposed to be T-dual. Other predictions of the conjectured duality are that type 0B string theory exhibits S-duality and that the type 0 theories contain non-perturbative fermionic states. In Ref. it has been studied how branes transform under type 0/type II T-duality. D-branes in type 0 theories have been extensively studied . Since their dynamics is governed by open strings ending on them, D-branes can be studied in perturbative string theory. This is not the case for solitonic (NS) fivebranes, which can be studied by examining the zero modes of the corresponding classical supergravity solutions . Each such zero mode corresponds to a massless field on the fivebrane worldvolume. In this paper we derive the massless field content of type 0 NS5-branes. Rather than doing the zero mode analysis directly for type 0A/B NS5-branes, we T-dualize them to Kaluza-Klein monopoles of type 0B/A, where the zero modes can be read off from the ten-dimensional bulk fields and the cohomology of the Euclidean Taub-NUT space. For both NS5-branes we find a non-chiral spectrum: an unrestricted antisymmetric tensor and six scalars in 0A and two gauge fields and four scalars in 0B. In type IIA the spectrum of the NS5 is chiral and anomalous. The gravitational anomaly on the NS5 must be cancelled by anomaly inflow from the bulk. The anomaly-cancelling bulk term can be shown to be a string one-loop effect . The fact that we find a non-anomalous spectrum on type 0A NS5-branes predicts that such an anomalous bulk term is not present in type 0A. Indeed, we will see that the type 0A GSO projection does not allow a CP-odd one-loop term. Having identified the rather unusual NS5-brane spectrum in type 0B, one naturally wonders what the NS5 transforms into under the conjectured type 0B S-duality. Following Ref. it seems natural to propose that the dual object is a bound state of an electric and a magnetic D5-brane. However, we will point out that this proposal gives rise to some puzzles. Finally, we propose an interpretation of the NS5-brane spectra in terms of “type 0 little strings”. This interpretation is suggested by the doubling of the gauge degrees of freedom on the NS-branes, which is reminiscent of (and, in fact, closely related to) the doubling of the Ramond-Ramond spectrum in going from type II to type 0. ## 2 Type 0 strings and D-branes In the Neveu-Schwarz-Ramond formulation, type II string theories are obtained by imposing independent GSO projections on the left- and right-moving parts. This amounts to keeping the following (left,right) sectors: $`\mathrm{IIB}:`$ $`(NS+,NS+),(R+,R+),(R+,NS+),(NS+,R+);`$ $`\mathrm{IIA}:`$ $`(NS+,NS+),(R+,R),(R+,NS+),(NS+,R),`$ (2.1) where for instance R$`\pm `$ is the Ramond sector projected with $`P_{\mathrm{GSO}}=(1\pm ()^F)/2`$, $`F`$ being the world-sheet fermion number. The type 0 string theories contain instead the following sectors: $`0\mathrm{B}:`$ $`(NS+,NS+),(NS,NS),(R+,R+),(R,R);`$ $`0\mathrm{A}:`$ $`(NS+,NS+),(NS,NS),(R+,R),(R,R+).`$ (2.2) These theories do not contain bulk spacetime fermions, which would have to come from “mixed” (R,NS) sectors. The inclusion of the NS-NS sector with odd fermion numbers means that the closed string tachyon is not projected out. The third difference with type II theories is that the R-R spectrum is doubled. As discussed in the introduction, the spectrum of type 0 theories contains two ($`p+1`$)-form R-R potentials for each even (0A) or odd (0B) $`p`$. We will denote these by $`C_{p+1}`$ and $`C_{p+1}^{}`$, where the unprimed potentials are the ones present in type II. For our purposes, more convenient combinations are $$(C_{p+1})_\pm =\frac{1}{\sqrt{2}}(C_{p+1}\pm C_{p+1}^{}).$$ (2.3) For $`p=3`$ these are the electric ($`+`$) and magnetic ($``$) potentials . We will adopt this terminology also for other values of $`p`$. There are four types of “elementary” D-branes for each $`p`$: an electric and a magnetic one (i.e. charged under $`(C_{p+1})_\pm `$), and the corresponding antibranes. In Ref. the interaction energy of two identical parallel ($`p+1`$)-branes was derived by computing the relevant cylinder diagram in the open string channel, analogously to the Polchinski computation in type II. Isolating, via modular transformation, the contributions due to the exchange of long-range fields in the closed string channel, it is found on the one hand that the tension of these branes is a factor $`\sqrt{2}`$ smaller than for type II branes. On the other hand, the R-R repulsive force between two like branes has the double strength of the graviton-dilaton attraction ; thus the type 0 branes couple to the corresponding R-R potentials $`(C_{p+1})_\pm `$ with the same charge as the branes in type II couple to the potential $`C_{p+1}`$. The open strings stretching between two like branes are bosons, just like the bulk fields of type 0. However, fermions appear from strings between an electric and a magnetic brane . ## 3 NS fivebrane spectra In type II string theories the massless spectrum on NS5-branes can be derived via an analysis of the zero modes of the classical supergravity NS5-brane solutions . We will find it convenient to do the analysis for the T-dual objects, since there the geometrical interpretation is manifest. Under T-duality the type IIA/B NS5-branes are mapped to type IIB/A Kaluza-Klein (KK) monopoles. A KK monopole is described by a supergravity solution with six worldvolume directions and a Euclidean Taub-NUT (ETN) metric in the transverse space. The massless fields on the KK monopole correspond to the zero modes of the bulk supergravity fields on ETN. The normalizable harmonic forms on ETN consist of one selfdual two-form. In the NS-NS sector there are three zero modes from broken translation invariance and one scalar from the NS two-form. These four scalars correspond to the translation zero modes on the corresponding NS5-branes. The R-R fields give additional bosonic zero modes: on the IIB KK monopole the R-R two-form gives rise to a scalar and the selfdual<sup>1</sup><sup>1</sup>1 By abuse of language we term the potential forms (anti-)selfdual, according to the (anti-)selfdual nature of their respective field strengths. R-R four-form potential leads to a selfdual two-form; the IIA KK monopole has a vector coming from the R-R three-form. In addition there are fermionic zero modes from broken supersymmetry: two fermions with opposite chirality for the IIA KK monopole and of the same chirality for IIB. All in all, this leads to an $`𝒩=(2,0)`$ tensor multiplet on the IIA NS5 and an $`𝒩=(1,1)`$ vector multiplet on the IIB NS5. Repeating this analysis for type 0 NS5-branes, it is clear that the fermionic zero modes disappear (there are no fermions in the bulk). Nothing changes as far as the NS-NS zero modes are concerned, but from the R-R fields we find extra zero modes from the doubled R-R spectrum in the bulk. On the 0B NS5 there is an extra vector (compared to the IIB NS5), whereas the 0A NS5 gets an extra scalar and an anti-selfdual tensor. In particular the spectrum is non-chiral on both branes. ## 4 Absence of anomalies At one string loop the type IIA tree level supergravity action is supplemented with the Wess-Zumino type term $`BX_8(R)`$ coupling the NS-NS two-form $`B`$ to four gravitons. $`X_8`$ is a quartic polynomial in the spacetime curvature. The original derivation of this one-loop term provided a non-trivial consistency check for six-dimensional heterotic-type II duality. Apart from that, this specific piece of the action is also responsible for gravitational anomaly cancellation on IIA NS fivebranes . Let us first highlight some aspects of the direct calculation in type II, which will enable us to draw conclusions on the analogous type 0 amplitude almost effortlessly. First, as the interaction contains the ten-dimensional $`ϵ`$ symbol, only odd spin structures on the torus can contribute, i.e. either the left-moving or the right-moving fermions but not both, have to be in the odd spin structure. Furthermore, the even spin structures are summed over to achieve modular invariance. From a different perspective this may also be seen as performing the GSO projection on the closed string states in the loop . In a Hamiltonian framework the torus vacuum diagram is written as a trace over the closed string Hilbert space, which decomposes in four different sectors as in Eq. (2.1). Chiral traces over these sectors can be expressed in terms of traces over fermion sectors with periodic (P) or antiperiodic (A) boundary conditions in the ($`\sigma `$,$`\tau `$) worldsheet directions: $`\mathrm{Tr}_{(NS,)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left((A,A)(A,P)\right);`$ $`\mathrm{Tr}_{(NS,+)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left((A,A)+(A,P)\right);`$ $`\mathrm{Tr}_{(R,)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left((P,A)(P,P)\right);`$ $`\mathrm{Tr}_{(R,+)}`$ $`=`$ $`{\displaystyle \frac{1}{2}}\left((P,A)+(P,P)\right).`$ Notice that the odd $`(P,P)`$ spin structure only occurs in the Ramond sector. From the GSO-projected spectrum in Eq. (2.1) only (R,NS) and (R,R) sectors contain (odd,even) pieces, yielding $$\frac{1}{4}((P,P),(A,A)+(A,P)+(P,A)).$$ (4.4) The right-moving combination is modular invariant, and in fact vanishes as a consequence of the “abstruse identity”. If one were to calculate the amplitude for one B-field and four gravitons, one would insert vertex operators in the appropriate pictures. Still, the right-moving even structures would be summed over as in Eq. (4.4). In contrast with the partition function however, in this case a non-vanishing result would be found so as not to contradict the alternative calculation in Ref. . It is also argued there that in type IIB the (odd,even) and (even,odd) contributions cancel out whereas they add in type IIA. So only in the latter is the probed interaction present. This difference between type IIA and IIB could have been inferred from their respective NS fivebrane worldvolume spectra, a chiral one in type IIA and a non-chiral one in type IIB. The chiral spectrum suffers from a gravitational anomaly in six dimensions, which is neatly cancelled by the standard anomaly inflow from the bulk through the above derived coupling . Let us repeat the argument for the one-loop correlator in type 0 string theories. Now it is only (R,R) sectors that contribute to the (odd,even) piece in the one-loop partition function, yielding $$\frac{1}{4}((P,P)(P,P),(P,A))$$ (4.5) both for 0A and 0B. As in type II one may now wish to insert vertex operators. However, it is obvious that the obligatory sum over spin structures as in Eq. (4.5) now yields a vanishing result for the (odd,even) part by itself. Analogously, the (even,odd) part will vanish by itself. Whence the absence of the anomaly-cancelling term in both type 0A and 0B. Reversing the anomaly inflow argument in type 0A the selfdual two-form must be supplemented with an anti-selfdual two-form, in order not to give rise to gravitational anomalies. This nicely agrees with the above derived unconstrained two-form on the NS fivebrane. From T-duality an additional vector (compared to the IIB NS5-brane) has to be expected on the type 0B NS5-brane. ## 5 S-duality ### Strings In type IIB string theory the fundamental string is mapped to the D-string under S-duality. The Green-Schwarz light-cone formulation of the fundamental type IIB string involves eight spacetime bosonic fields and sixteen spacetime fermionic ones. All fermions transform in the $`\mathrm{𝟖}_𝐒`$ representation of the transverse SO(8) rotation group. Upon quantization, their zero modes generate the 256-fold degenerate groundstate, transforming in the $`(\mathrm{𝟖}_𝐕+\mathrm{𝟖}_𝐂)(\mathrm{𝟖}_𝐕+\mathrm{𝟖}_𝐂)`$ of SO(8). The excitations of a D1-brane can be described in string perturbation theory by quantizing the open strings beginning and ending on it. Doing so, the NS sector gives eight massless bosons and the R sector sixteen massless fermions. Note that these massless excitations of both the fundamental string and the D-string can be interpreted as Goldstone modes for broken translation invariance and broken supersymmetry. It is believed that the IIB D-string at strong (fundamental) string coupling is the same object as a fundamental IIB string at weak coupling. The evidence for this conjecture is largely based on supersymmetry. For one thing, the ground state of the D-string is BPS, so that it can safely be followed to strong coupling and compared to the ground state of the weakly coupled fundamental string. Type 0B string theory is not supersymmetric, which complicates the analysis of its strong coupling limit enormously. Let us nevertheless try to proceed. The fundamental type 0B string can be obtained from the fundamental type IIB string by performing a $`(1)^{F_s}`$ orbifold, where $`F_s`$ is the spacetime fermion number. In the untwisted sector of this orbifold the massless states in the $`(\mathrm{𝟖}_𝐕\mathrm{𝟖}_𝐕)+(\mathrm{𝟖}_𝐂\mathrm{𝟖}_𝐂)`$ survive. The light modes in the twisted sector are a singlet tachyon and the extra massless R-R states in the $`\mathrm{𝟖}_𝐒\mathrm{𝟖}_𝐒`$. Bergman and Gaberdiel have conjectured that the fundamental type 0B string should be S-dual to a bound state an electric and a magnetic D1-brane. This proposal immediately raises two questions. First, when one quantizes the open strings on a coincident electric-magnetic D-string pair, one finds twice as many massless excitations as on a fundamental string: sixteen bosons and thirty-two fermions. This would lead to way too many states of the D-string pair compared to the fundamental string. How could the extra degrees of freedom disappear? Second, assuming that we have found a way to get rid of the superfluous degrees of freedom, how does the $`(1)^{F_s}`$ projection appear on the D-string pair? Concerning the first question, let us just note that this naive mismatch is no immediate contradiction: to compare the D-string pair to the weakly coupled fundamental string, one should take the former to strong (fundamental string) coupling. Along the way, the excitation spectrum may well change. As to the second question, the $`(1)^{F_s}`$ projection might be related to a gauging of a $`\text{ }\text{}_2`$ subgroup of the gauge symmetry on the D-string pair. However, it looks hard to make these ideas more precise. ### Fivebranes Type IIB NS5-branes are S-dual to D5-branes. These objects have the same light excitations (the reduction of an $`𝒩=1`$, D=10, U(1) vector multiplet to D=6). As in the case of IIB strings, part of the spectrum is protected by supersymmetry. Given the story for strings, it looks reasonable to propose that the type 0B NS5-brane is S-dual to an electric-magnetic D5-brane pair. This raises two puzzles. First, how do the spectra match? Is the doubled gauge spectrum of type 0B NS5-branes at weak coupling related to the two gauge fields on the D5-brane pair? This is not clear, since to compare the two objects one has to take one of them to strong coupling, where it is unclear what happens. For one thing, the doubling of the gauge field on the type 0B NS5 (compared to the IIB NS5) is related to the doubled R-R spectrum of the bulk of type 0A, as shown in Section 3. The latter doubling is not expected to persist at strong coupling . Are there fermionic zero modes on NS5-branes? We have not found any from a zero mode analysis of the bulk gravity theory, but this does not exclude that they could appear in a non-perturbative way. This would be analogous to the (not completely understood) case of the fundamental type 0 string, where it looks impossible to interpret the fermion zero modes as zero modes of bulk fields. Second, do type 0B NS5-branes carry anomalous gravitational couplings, which are known to be carried by type 0B D-branes ? Since these couplings are related to anomalies and thus to topology, we hope that their presence or absence could be invariant under continuously changing the coupling from zero to infinity. If this is the case, answering this question could provide a test of type 0B S-duality. The rest of this section will be devoted to analysing this question. We first remind the reader of the familiar story in type IIB. Consider, for instance, the intersection of two D5-branes on a string. The open strings from one brane to the other give rise to chiral fermions living on the intersection. These chiral fermions lead to gauge and gravitational anomalies, which are cancelled by anomaly inflow from the branes into the intersection . For this to happen, the D5-branes need, amongst others, a $$C_2p_1$$ (5.6) term in their Wess-Zumino action, where $`C_2`$ is the R-R two-form potential and $`p_1`$ the first Pontrjagin class of the tangent bundle of the D5-brane worldvolume. The presence of this term has been checked by an explicit string computation . Now, type IIB S-duality implies the presence of chiral fermions on the intersection of two NS5-branes on a string, and thus a $$Bp_1$$ (5.7) term on the worldvolumes of each NS5-brane ($`B`$ is the NS-NS two-form). Indeed, the following argument<sup>2</sup><sup>2</sup>2In fact, this is how the $`C_2p_1`$ coupling on D5-branes was first discovered. for the presence of this term was given in Ref. . Under T-duality in a transverse direction the NS5-brane is mapped to a Kaluza-Klein monopole in type IIA. This object is described by a Euclidean Taub-NUT metric in the transverse space. Since $`p_1(\mathrm{ETN})0`$ and the $`BX_8`$ term discussed in the previous section contains a piece proportional to $`B(p_1)^2`$, reducing this term gives rise to the required $`Bp_1`$ term on the IIA Kaluza-Klein monopole and thus on the IIB NS5-brane. What about type 0B? It has been argued that chiral fermions and thus anomalies are present on the intersection of electric and magnetic D5-branes. Hence, type 0 D5-branes have the $`p_1`$ couplings in their Wess-Zumino actions (this has also been checked via a string computation ). We expect that the number of chiral fermion zero modes on these intersections cannot change under a continuous increase of the string coupling. On the other hand, in this paper we have argued that there is no $`BX_8`$ term in type 0A, so we expect no $`Bp_1`$ coupling of type 0B NS5-branes, and thus no chiral fermions on intersections of NS5-branes. If this argument is correct, it could be a problem for type 0B S-duality. ## 6 A little string interpretation In type II string theories the fivebrane worldvolume theories have been conjectured to be properly described by “little strings”. These non-critical closed string theories have hitherto remained mysterious, although some qualitative features may be understood. We briefly recall some of these, focusing on the type IIB case. However, we prefer first to treat the more familiar example of a fundamental string, providing a guiding reference when dealing with the little strings. A fundamental string solution in supergravity breaks half of the 32 bulk supersymmetries. These broken symmetries give rise to fermion zero modes on the string, of which 8 are left-moving and 8 right-moving. They are identified with the Green-Schwarz space-time fermion fields on the worldvolume. Upon quantization of these fermions 128 bosons and 128 fermions are found, together building the on-shell $`𝒩`$ = 2 supergravity multiplet. This multiplet is also found from the RNS superstring, so the picture is consistent. In the latter formulation it is not hard to see that the NS-NS sector fields together with, say, the R-NS sector build an $`𝒩`$ = 1 supergravity multiplet. Likewise, the NS-R and R-R sector are combined into one multiplet of $`𝒩`$ = 1. Let us move on now to little strings. One way to derive properties of little strings in type IIB is to consider a gauge instanton on a D5 brane. Half of the bulk supersymmetries are unbroken by the D5-brane. Sixteen real supercharges are thus found, obeying a $`𝒩`$=(1,1) superalgebra. Of these, half are broken by the instanton. On the macroscopic string there are thus eight fermionic modes, four left-moving and four right-moving ones. Upon quantization of these zero modes one finds 8 bosonic and 8 fermionic states. They correspond to the on-shell degrees of freedom of one $`𝒩`$=(1,1) vectormultiplet.<sup>3</sup><sup>3</sup>3Since the little strings of type IIB are non-chiral, we will name them type iia little strings. Upon decomposition with respect to $`𝒩`$ = 1 six-dimensional supersymmetry one hypermultiplet and one vectormultiplet result. It is conceivable that a similar analysis in type IIA, if possible at all, would give a $`𝒩`$=(2,0) tensormultiplet, decomposing into one hypermultiplet and one tensormultiplet of $`𝒩`$=1 supersymmetry. We remark that the tensormultiplet has a bosonic content that consists of one scalar and one two-form potential with self-dual field strength. All in all, one may wish to draw the following analogies between little strings and type II strings. In both cases there is one universal sector, the hypermultiplet resp. the (NS-NS + R-NS) sector. The other sector depends on whether or not the supersymmetry in the “bulk” is chiral, where “bulk” means the fivebrane worldvolume for the little string and ten-dimensional spacetime for the type II string. For the non-chiral type IIA and little type iia the remaining bosonic fields are odd $`q`$-form potentials, coupling to even $`p`$ D$`p`$\- resp. d$`p`$-branes. As to chiral type IIB/iib the R-R sector has only even $`q`$-form potentials, of which one has a self-dual field strength. In these theories only odd $`p`$ D$`p`$ or d$`p`$-branes occur. So this analogy, however formal, is at least remarkable. As to the various fivebranes, the type IIA NS fivebrane theory is conjectured to be described by chiral iib little strings, whereas iia little strings would do the job for type IIB fivebranes. If one is willing to take the above for granted, one may then proceed and construct “little type 0a/b strings” by mimicking the procedure for the bulk strings: spacetime fermions are removed and there is a doubled R-R spectrum compared to the superstring. Concretely this would imply that the massless spectrum of little type 0a strings consists of the bosonic contents of one hypermultiplet and two vectormultiplets, in $`𝒩`$=1 language. This precisely matches with the spectrum on the type 0B NS fivebrane. Little type 0b strings would then generate the bosonic content of one hypermultiplet, one selfdual tensormultiplet and one anti-selfdual tensormultiplet. Also here this seems to be consistent with the spectrum of the type 0A NS fivebrane. ## 7 Conclusions From a zero mode analysis of type 0 Kaluza-Klein monopoles, we have derived the massless degrees of freedom of NS5-branes in type 0A and 0B. The massless spectrum is purely bosonic and non-chiral in both cases. The 0A NS5-brane contains an unrestricted tensor (rather than a selfdual one) whereas two gauge fields live on the 0B NS5-brane. Since the spectrum of a 0A NS5-brane is non-chiral, the anomaly-cancelling one loop bulk term present in type IIA should be absent in type 0A. This is indeed the case. We have pointed out some puzzles concerning type 0B S-duality. In particular, the question is raised whether the absence of a $`BX_8`$ term in the bulk of type 0A makes the duality of NS5-branes with D5-brane pairs impossible. It is tempting to interpret the spectrum we find in terms of “type 0 little strings”. Their relation to the usual type II “little strings” is very similar of the relation between the bulk type 0 and type II strings: fermions are projected out and the “Ramond-Ramond” spectrum is doubled. ###### Acknowledgments. We would like to thank M. Billó, M. Costa, M. Gaberdiel, R. Gopakumar, I. Klebanov, A. Sen, J. Troost and W. Troost for interesting discussions, comments and suggestions.
no-problem/9906/cond-mat9906297.html
ar5iv
text
# Thin superconducting disk with 𝐵-dependent 𝐽_𝑐 : Flux and current distributions ## I Introduction The critical state model (CSM) is widely accepted to be a powerful tool in the analysis of magnetic properties of type-II superconductors. For decades there have been numerous theoretical works devoted to CSM calculations in the parallel geometry, i. e., a long sample placed in parallel applied magnetic field, $`B_a`$. More recently, much attention has also been paid to the CSM analysis of thin samples in perpendicular magnetic fields. For this so-called perpendicular geometry explicit analytical expressions for flux and current distributions have been obtained for a long thin strip and thin circular disk assuming a constant critical current (the Bean model). From experiments, however, it is well known that the critical current density $`j_c`$ usually depends strongly on the local flux density $`B`$. This dependence often hinders a precise interpretation of various measured quantities such as magnetization, complex ac susceptibility, and surface impedance. It is therefore essential to extend the CSM analysis to account for a $`B`$-dependence of $`j_c`$. In the parallel geometry extensive work has already been carried out, and exact results for the flux density profiles and magnetization,, as well as ac losses have been obtained for different $`j_c(B)`$-dependences. In the perpendicular geometry the magnetic behavior is known to be qualitatively different. In particular, due to a strong demagnetization, the field tends to diverge at the sample edges, and the flux penetration depth and ac losses follow different power laws in $`B_a`$ for small $`B_a`$. Unfortunately, the theoretical treatment of the perpendicular geometry is very complicated, and we are not aware of any explicit expressions obtained for the CSM with a $`B`$-dependent $`j_c`$. However, it is possible to derive integral equations relating the flux and current distributions. Such equations have so far been obtained and solved numerically only for the case of a long thin strip. In this paper, we derive a CSM solution for a thin circular disk characterized by an arbitrary $`j_c(B)`$. The solution is presented as a set of integral equations which we solve numerically. In this way we obtain the field and current density distributions in various magnetized states. We present results for several commonly used functions $`j_c(B)`$. A special attention is paid to the low-field asymptotic behavior. The paper is organized as follows. In Sec. II the basic equations for the disk problem are derived. We consider here all states during a complete cycle of the applied field, including the virgin branch. Sec. III contains our numerical results for flux and current distributions as well as for the flux front position. A discussion of the results is presented. Finally, Sec. IV presents the conclusions. ## II Basic equations Consider a thin superconducting disk of radius $`R`$ and thickness $`d`$, where $`dR`$, see Fig. 1. We assume either that $`d\lambda `$, where $`\lambda `$ is the London penetration depth, or, if $`d<\lambda `$, that $`\lambda ^2/dR`$. In the latter case the quantity $`\lambda ^2/d`$ plays a role of two-dimensional penetration depth. We put the origin of the reference frame at the disk center and direct the $`z`$-axis perpendicularly to the disk plane. The external magnetic field $`𝐁_a`$ is applied along the $`z`$-axis, the $`z`$-component of the field in the plane $`z=0`$ being denoted as $`B`$. The current flows in the azimuthal direction, with a sheet current denoted as $`J(r)=_{d/2}^{d/2}j(r,z)𝑑z`$, where $`j`$ is the current density. To obtain expressions for the current and flux distribution we follow a procedure originally suggested in Ref. and then generalized in Ref. for the case of $`B`$-dependent $`J_c`$ in a thin strip. The procedure makes use of the Meissner state distributions for $`B`$ and $`J`$. In the Meissner state, where $`B=0`$ inside the disk, the field outside the disk is given by $$B_M(r,R)=B_a+\frac{2B_a}{\pi }\left[\frac{R}{\sqrt{r^2R^2}}\mathrm{arcsin}\left(\frac{R}{r}\right)\right],$$ (1) and the current is distributed according to $$J_M(r,R)=\frac{4B_a}{\pi \mu _0}\frac{r}{\sqrt{R^2r^2}},r<R.$$ (2) ### A Increasing field We begin with a situation where the external field $`B_a`$ is applied to a zero-field-cooled disk. The disk then consists of an inner flux-free region, $`ra`$ , and of an outer region, $`a<rR`$, penetrated by magnetic flux. According to the critical state model, the penetrated part will carry the critical sheet current $`J_c`$ corresponding to the local value of magnetic field, $$J(r)=J_c[B(r)],a<r<R$$ (3) Now following Refs. , we express the field and current as superpositions of the Meissner-state distributions, (1) and (2), i.e.: $`B(r)`$ $`=`$ $`{\displaystyle _a^{\mathrm{min}(r,R)}}𝑑r^{}B_M(r,r^{})G(r^{},B_a).`$ (4) $`J(r)`$ $`=`$ $`{\displaystyle _{\mathrm{max}(a,r)}^R}𝑑r^{}J_M(r,r^{})G(r^{},B_a),`$ (5) where $`G(r,B_a)`$ is a weight function. Since $`B(r)`$ and $`B_M(r,R)B_a`$ at $`r\mathrm{}`$ we have the normalization condition $$_a^R𝑑rG(r,B_a)=1.$$ (6) Substituting Eq. (3) and Eq. (2) into Eq. (5) yields an integral equation for the function $`G(r,B_a)`$ which can be inverted to obtain $$G(r,B_a)=\frac{B_c}{B_a}\frac{d}{dr}_r^R\frac{dr^{}}{\sqrt{r^2r^2}}\frac{J_c[B(r^{})]}{J_{c0}},$$ (7) where $$B_c=\mu _0J_{c0}/2,J_{c0}J_c(B=0).$$ (8) Note that due to the similar form of the function $`J_M(r,R)`$ in (2) and the Meissner-state current in the strip case our weight function $`G(r,B_a)`$ is also similar to that for a strip, see Ref. From Eqs. (5) and (7) it then follows that the current distribution in a disk is given by $$J(r)=\{\begin{array}{cc}\frac{2r}{\pi }_a^R𝑑r^{}\sqrt{\frac{a^2r^2}{r^2a^2}}\frac{J_c[B(r^{})]}{r^2r^2},\hfill & \hfill r<a\\ J_c[B(r)],\hfill & \hfill a<r<R\end{array}$$ (9) This equation is supplemented by the Biot-Savart law, which for a disk reads, $$B(r)=B_a+\frac{\mu _0}{2\pi }_0^RF(r,r^{})J(r^{})𝑑r^{}.$$ (10) Here $`F(r,r^{})=K(k)/(r+r^{})E(k)/(rr^{})`$, where $`k(r,r^{})=2\sqrt{rr^{}}/(r+r^{})`$, while $`K`$ and $`E`$ are complete elliptic integrals defined as $`E(z)=_0^{\pi /2}\left[1z^2\mathrm{cos}^2(x)\right]^{1/2}𝑑x`$ and $`K(z)=_0^{\pi /2}\left[1z^2\mathrm{cos}^2(x)\right]^{1/2}𝑑x`$. The relation between the flux front location $`a`$ and applied field $`B_a`$ is obtained by substituting Eq. (7) into Eq. (6), giving $$B_a=B_c_a^R\frac{dr^{}}{\sqrt{r^2a^2}}\frac{J_c[B(r^{})]}{J_{c0}}.$$ (11) For a given $`B_a`$ and for a specified $`J_c(B)`$ we need to solve the set of three coupled equations (9)-(11). In the case of $`B`$-independent $`J_c`$, the Eq. (11) acquires the simple form $$a/R=1/\mathrm{cosh}(B_a/B_c),$$ (12) and the Eqs. (9) and (10) lead to the Bean-model results derived in Refs. and . Note that the equations can be significantly simplified at large external field where $`a0`$ proportionally to $`\mathrm{exp}(B_a/B_c)`$. Then $`B(r)`$ is determined by the single equation $$B(r)=B_a\frac{\mu _0}{2\pi }_0^RF(r,r^{})J_c[B(r^{})]𝑑r^{},$$ (13) following from Eq. (10). ### B Subsequent field descent Consider now the behavior of the disk as $`B_a`$ is reduced after being first raised to some maximal value $`B_{am}`$. Let us denote the flux front position, the current density and the field distribution at the maximum field as $`a_m`$, $`J_m(r)`$ and $`B_m(r)`$, respectively. Obviously, $`J_m(r)`$, $`B_m(r)`$, and $`a_m`$ satisfy Eqs. (9)-(11). During the field descent from $`B_{am}`$ the flux density becomes reduced in the outer annular region $`a<r<R`$, see Fig. 2. The central part of the disk $`r<a`$ remains frozen in the state with $`B_a=B_{am}`$. Let us specify the field and current distributions in this remagnetized state as $$B(r)=B_m(r)+\stackrel{~}{B}(r),J(r)=J_m(r)+\stackrel{~}{J}(r),$$ (14) and derive the relation between $`\stackrel{~}{B}(r)`$ and $`\stackrel{~}{J}(r)`$. For that one can use a procedure similar to the one described in Sec. II A. The only difference is that in the region $`a<r<R`$ we now have to use $`J(r)=+J_c[B(r)]`$. In this way we obtain $$\stackrel{~}{J}(r)=\stackrel{~}{J}_c(r),a<r<R,$$ where we define $$\stackrel{~}{J}_c(r)=J_c[B_m(r)+\stackrel{~}{B}(r)]+J_c[B_m(r)].$$ (15) Note that the function $`\stackrel{~}{J}_c(r)`$ depends on the coordinate only through the field distributions $`B_m(r)`$ and $`\stackrel{~}{B}(r)`$. Instead of Eq. (9), the additional current satisfies $$\stackrel{~}{J}(r)=\{\begin{array}{cc}\frac{2r}{\pi }_a^R𝑑r^{}\sqrt{\frac{a^2r^2}{r^2a^2}}\frac{\stackrel{~}{J}_c(r^{})}{r^2r^2},\hfill & \hfill r<a\\ \stackrel{~}{J}_c(r),\hfill & \hfill a<r<R\end{array}$$ (16) with the complementary equation $$\stackrel{~}{B}(r)=B_aB_{am}+\frac{\mu _0}{2\pi }_0^RF(r,r^{})\stackrel{~}{J}(r^{})𝑑r^{}.$$ (17) Furthermore, similarly to Eq. (11), we have $$\frac{B_aB_{am}}{B_c}=_a^R\frac{dr}{\sqrt{r^2a^2}}\frac{\stackrel{~}{J}_c(r)}{J_{c0}},$$ (18) which completes the set of equations describing the remagnetized state. Again, for $`B`$-independent $`J_c`$ the equations reproduce the Bean-model results If the field is decreased below $`B_{am}`$ the memory of the state at $`B_a=B_{am}`$ is completely erased, and the solution becomes equivalent to the virgin penetration case. If the difference $`B_{am}B_a`$ is sufficiently large then $`a0`$ rapidly, and the critical state $`J(r)=J_c(r)`$ is established throughout the disk. In this case the field descent is described by Eq. (13) with the opposite sign in front of the integral. We emphasize that the expressions derived here (Sec.II A and B) are readily converted to the long thin strip case. This is due to the similarity of (2) and the expression for the Meissner-current in a strip, $$J(x)=\frac{2B_a}{\mu _0}\frac{x}{\sqrt{w^2x^2}},w<x<w,$$ (19) where $`x`$ is the coordinate across the strip. Thus, making in this paper the substitutions $$rx,Rw,F(r,r^{})\frac{2x^{}}{x^2x^2},B_c\frac{\mu _0J_{c0}}{\pi }$$ one immediately arrives at the set of equations valid for a thin long strip. In that case some of the integrals can be done analytically to yield the expressions obtained in Ref. . A difference between the derivation in Ref. and the present one is that for decreasing fields we calculate only the additional field $`\stackrel{~}{B}(r)`$ rather than the total field $`B(r)`$. This allows us to use only one weight function (7) to calculate the flux distributions both for increasing and decreasing fields. This simplifies the numerical calculations significantly. ### C Numerical procedure Given the $`J_c(B)`$-dependence, the magnetic behavior is found by solving the derived integral equations (9)-(11) numerically using the following iteration procedure. With $`B_a`$ increasing a flux front position $`a`$ is first specified, and an initial approximation for $`B(r)`$, e.g., the Bean model solution, is chosen. At each step the $`n`$th approximation, $`B^{(n)}(r)`$, is used to calculate $`J^{(n)}(r)`$ from Eq. (9) and $`B_a^{(n)}`$ from Eq. (11). They are then substituted into Eq. (10) yielding the next approximation, $`B^{(n+1)}(r)`$. The iterations are stopped when $`\left(R^1𝑑r\left[B^{(n+1)}(r)B^{(n)}(r)\right]^2\right)^{1/2}10^6B_c`$. With $`B_a`$ decreasing, the same procedure is used to find first $`J_m(r)`$, $`B_m(r)`$, $`B_{am}`$ for a given $`a_m`$. Then, Eqs. (16)-(18) are solved for a fixed $`a`$ yielding the functions $`\stackrel{~}{B}(r)`$ and $`\stackrel{~}{J}(r)`$ and also the applied field $`B_a`$. ## III Flux and current distribution ### A General features In the numerical calculations we used the following dependences $`J_c(B)`$, $`J_c`$ $`=`$ $`J_{c0}/(1+|B|/B_0)\text{(Kim model),}`$ (20) $`J_c`$ $`=`$ $`J_{c0}\mathrm{exp}(|B|/B_0)\text{(exponential model).}`$ (21) Shown in Fig. 3 (a,b) are the field and current distributions for increasing field with $`a/R=0.2`$ for the Kim model with different parameters $`J_{c0}`$ and $`B_0`$. They are chosen in a way to keep the position $`a`$ fixed for all the curves for a given value of $`B_a`$. This allows us to follow the variations in the profile shape as the $`B`$-dependence of $`J_c`$ changes. The chosen parameters $`J_{c0}`$ and $`B_0`$ correspond to the set of $`J_c(B)`$-curves shown in Fig. 3 (c). The Bean model results are also plotted in Fig. 3. Several major deviations between the Kim and the Bean model can be noticed. In the Kim model we see that (i) the current $`J(r)`$ is not uniform at $`a<r<R`$ – it is minimal at the disk edge where $`|B|`$ is maximal; (ii) the current has a cusp-like maximum at $`r=a`$ since the magnetic field vanishes at this point with infinite derivative; (iii) compared to the Bean model, the $`B(r)`$ profiles are steeper near the flux front, whereas the peaks at the edges are less sharp. Qualitatively similar results are obtained for the exponential model, see Fig. 4. Also here, by changing the model parameters one can produce a variety of flux and current profiles which are quite different from the Bean model predictions. When comparing the Kim and exponential model, however, it turns out to be very difficult to find clear distinctions. During field descent the $`B`$\- and $`J`$-profiles become more complicated. For brevity we show only profiles for the Kim model with $`B_0/B_c=3`$ and for the Bean model at different values of $`B_a`$, see Fig. 5. Again, the Kim model gives a non-uniform current density at $`r>a`$. Contrary to the increasing-field states, the current density can now either decrease, or increase towards the edge depending on $`B_a`$. Figure 6 shows profiles for fully-penetrated decreasing-field state. In the Bean model the current remains constant, while the profile of the flux distribution is fixed, although shifted according to the applied field. In contrast, the Kim-model profiles are strongly dependent on $`B_a`$. There is a peak in the current profile and an enhanced gradient of $`B(r)`$ near the point where $`B=0`$. ### B Flux penetration depth To analyze quantitatively the role of a $`B`$-dependent $`J_c`$ let us consider the position of the flux front during increasing field. A circle with radius $`a`$ then limits the Meissner region $`B=0`$, and is also the location of maximum gradient in $`B`$. These features can be measured directly in experiments on visualization of magnetic flux distribution, e. g., magneto-optical imaging. Let us first recall the CSM expression for the flux front location in a long circular cylinder in a parallel field , $$\frac{a}{R}=1\frac{1}{\mu _0R}_0^{B_a}\frac{dB}{j_c(B)}.$$ (22) At small applied fields, $`B_a`$, it can be expanded as $$\frac{a}{R}1\frac{B_a}{B_c}+\frac{\mu _0R}{2}j_c^{}(0)\left(\frac{B_a}{B_c}\right)^2,$$ (23) where for a cylinder $`B_c\mu _0j_{c0}R`$. Note that the $`B`$-dependence of $`j_c`$ enters the expansion first in the second-order term. Consequently, for a long cylinder the low-field behavior of $`a`$ is well described by the Bean model where the penetration depth increases linearly with the applied field. For a thin disk, the penetration of flux proceeds differently. In the Bean model the location of the flux front, Eq. (12), is for small $`B_a`$ given by $$a/R1(1/2)(B_a/B_c)^2.$$ (24) For an arbitrary $`B`$-dependence of $`J_c`$ the expression (11) relating $`a`$ and $`B_a`$ cannot easily be expanded in powers of the ratio $`B_a/B_c`$. The physical reason for this is the singular behavior of the magnetic field near the disk edge. There, the local field diverges at any finite $`B_a`$, and an expansion of $`J_c(B)`$ in powers of $`B`$ is not everywhere convergent. To clarify the behavior of the flux front we have therefore performed numerical calculations of the dependences $`a(B_a,B_0)`$. Shown in Fig. 7 are the results for the Kim (upper panel) and the exponential (lower panel) models. Note that the limit of large $`B_0`$ represents the Bean model. For small $`B_a`$ all the models seem to yield a parabolic relation between the penetration depth and the applied field. This is illustrated in more detail in Fig. 8, where all the graphs in the log-log plot have a slope of 2 in the low-field region. We therefore conjecture that any $`B`$-dependence of $`J_c`$ leads to the same quadratic law (24) as for the Bean model, although with different coefficients in front of $`(B_a/B_c)^2`$. The overall behavior of the penetration depth can be fitted well by the full form Eq. (12), provided one makes the substitution $`B_cB_c^{\text{eff}}`$, i.e., $$a/R=1/\mathrm{cosh}(B_a/B_c^{\text{eff}}).$$ (25) We find that the effective $`B_c`$ satisfies the relation $$\frac{B_c^{\text{eff}}}{B_c}=1\alpha \frac{B_c}{B_0},$$ (26) if the ratio $`B_0/B_c`$ is of the order 1, or larger. Here $`\alpha =0.42`$ for the exponential model, and $`\alpha =0.36`$ for the Kim model. The same relations (25) and (26) are found to hold true for a long thin strip with $`\alpha =0.60`$ and 0.51 for the exponential and the Kim model, respectively. We believe that for many purposes a Bean-model description with an effective critical current is appropriate for thin samples of any shape, both in applied field and under transport current. Indeed, strong demagnetization effects always lead to a divergence of magnetic field at the sample edge. This implies that in the sample there is always present a wide range of $`B`$ values up to infinity. As a result, the sample behavior is determined by the whole $`J_c(B)`$-dependence. In particular, the value $`J_c(0)`$ is not governing the magnetic behavior of thin samples, even when the applied field is very small. ## IV Conclusion A set of integral equations for the magnetic flux and current distributions in a thin disk placed in a perpendicular applied field is derived within the critical state model. The solution is valid for any field-dependent critical current, $`J_c(B)`$. By solving these equations numerically it is demonstrated that both the flux density and current profiles are sensitive to the $`J_c(B)`$-dependence. In particular, compared to the Bean model, the $`B(r)`$-profiles are steeper near the flux front, whereas the peaks at the edges are less sharp. Since the local magnetic field at the disk edge is divergent for any value of the applied field, $`B_a`$, a field dependence of $`J_c`$ affects the flux distribution even in the limit of low $`B_a`$. Our numerical calculations show that the flux penetration depth at small fields has the same quadratic dependence on $`B_a`$ as for the Bean model, however with different coefficient. The overall behavior of the flux penetration depth is well described by the Bean model expression with an effective value of the critical current. These results are believed to be qualitatively correct for thin superconductors of any shape. The behavior differs strongly from the case of a long cylinder in a parallel field, where the front position at low $`B_a`$ is not affected by the $`B`$-dependence of the critical current density. ###### Acknowledgements. The financial support from the Research Council of Norway is gratefully acknowledged.
no-problem/9906/cond-mat9906057.html
ar5iv
text
# Rate-and-State Theory of Plastic Deformation Near a Circular Hole ## I Introduction Since the work of Hart in the 1960’s, scientists have understood that a satisfactory theory of plastic deformation in solids must include dynamic variables that describe the internal states of materials. The deformation field itself cannot be sufficient. It cannot, in any natural way, describe the irreversible changes that lead to hysteretic stress-strain curves, or to the transition from nonlinear viscoelastic to viscoplastic behavior with increasing applied stress. Conventional theories of plasticity cope with these limitations by specifying phenomenological rules to suit various situations and histories of deformation. For example, strain hardening curves, viscoplastic laws, or the distinctions between loading and unloading behaviors are determined from experiment and used as needed in computations. In most treatments there is also a sharp distinction between time-independent and time-dependent behaviors, with little or no indication of how these properties may be related to one another. We believe that a deeper, more nearly fundamental level of phenomenology is required for modern applications, for example, for computing deformations near moving crack tips. Plasticity is an intrinsically time dependent phenomenon; time-independent descriptions should emerge as static limits of fully dynamic theories. In a recent paper , M. Falk and one of the present authors (JSL) proposed a theory of plastic deformation in amorphous solids in which they introduced an internal state variable to describe the orientations of what they called “shear-transformation zones.” (We refer to this as the “STZ” theory.) The resulting “rate-and-state” equations (a concept that is widely used in the seismological literature and in recent theories of friction ) successfully describe the full range of viscoelastic and viscoplastic phenomena, including hysteretic effects. The basic structure of the STZ theory appears to be broadly applicable. The new internal state variable might equally well describe, for example, anisotropy in the way dislocations pile up near defects in crystalline materials. Reference discusses plasticity only in spatially uniform situations. Our purpose here is to apply a simple version of the STZ theory to a spatially inhomogeneous situation and to make contact with conventional plasticity theory. We especially want to learn whether the conventional picture of a time-independent plastic zone appears in the static limit. Looking ahead to fully dynamic situations such as fracture, we also want to understand the dynamics of plastic flow in regions of concentrated stress. We shall show that the conventional time-independent concepts — the “yield-surface” hypotheses — do emerge from dynamic theories in an approximate way in many normal situations. As we shall argue, however, the rate-and-state theory is simpler, richer and more general than the conventional approaches. The problem of describing spatially inhomogeneous plastic deformation is best approached by looking at a simple example where questions of time dependence and compatibility are not obscured by mathematical details. A growing circular hole in a very large stressed plate satisfies the criterion of simplicity. It shares important features with the case of plastic deformation near a crack tip; the tractions are applied at a distance and the stresses are concentrated near the hole. Several researchers have addressed the problem of dynamic hole growth in the context of ductile fracture and spallation . Carroll and Holt and later Johnson included inertial effects but neglected rate dependent plasticity . Bodner and Partom used a version of a conventional rate dependent plasticity theory but did not study a stress controlled situation . The scheme of this paper is as follows. In Section II, we derive equations of motion for the radius of the circular hole and the surrounding stress field, assuming a general form for the constitutive relation that governs the rate of plastic deformation. Section III contains a brief summary of several results of conventional plasticity theory which will serve as points of comparison for the rate-and-state analysis. A simplified version of the STZ model is introduced in Section IV A where, for completeness, we outline some important properties of this model that were reported previously in and . We describe the dynamic behavior of the STZ model for the hole problem in Section IV B. The paper concludes with remarks about the implications of these results in Section V. ## II Dynamic Plasticity in a Circular Geometry Throughout this analysis, we consider only a two-dimensional solid in a state of plane strain, and assume that inertial effects are negligible. Suppose that a circular hole in this system has radius $`R(t)`$ at time $`t`$. Outward tractions at the distant edges of the plate cause the pressure $`p`$ far from the hole to be $`p\sigma _{\mathrm{}}`$. We introduce polar coordinates $`r`$ and $`\theta `$ that define an Eulerian reference frame, so that $`u(r,t)`$ is the radial displacement, measured from some initial reference state, of the material currently at position $`r`$. The function $`u(r,t)`$ is the only degree of freedom in the problem. The total rate-of-deformation tensor (including both elastic and plastic parts) is diagonal with components: $$𝒟_{rr}^{tot}=\frac{v}{r};𝒟_{\theta \theta }^{tot}=\frac{v}{r},$$ (1) where $`v`$ is the material velocity: $$v=\frac{u/t}{1u/r}.$$ (2) For small strains, the tensor $`𝒟^{tot}`$ is approximately equal to the total strain-rate tensor $`\dot{\epsilon }^{tot}`$. Eq.(1) implies that the components of $`𝒟^{tot}`$ satisfy the compatibility condition: $$\frac{}{r}\left(r𝒟_{\theta \theta }^{tot}\right)=𝒟_{rr}^{tot}.$$ (3) The stress tensor has the form: $$\sigma _{rr}=ps,\sigma _{\theta \theta }=p+s.$$ (4) Here, $`s=(\sigma _{\theta \theta }\sigma _{rr})/2`$ is the deviatoric stress. The elastic stress-strain relations are: $$2\mu \epsilon _{rr}^{el}=(12\nu )ps,2\mu \epsilon _{\theta \theta }^{el}=(12\nu )p+s,$$ (5) where $`\mu `$ is the shear modulus and $`\nu `$ is Poisson’s ratio. In the absence of inertial effects, balance of forces implies $$\frac{p}{r}=\frac{1}{r^2}\frac{}{r}(r^2s).$$ (6) The boundary condition at $`r=R`$ is $$\sigma _{rr}(R,t)=p(R,t)+s(R,t)=0.$$ (7) For simplicity, we neglect surface tension. Also at $`r=R`$, we have $$\frac{\dot{R}}{R}=𝒟_{\theta \theta }^{tot}[s(R),\mathrm{}].$$ (8) Here and elsewhere, dots above symbols denote derivatives with respect to time $`t`$. For incompressible plasticity, the constitutive relation has the form: $$𝒟_{\theta \theta }^{pl}=𝒟_{rr}^{pl}𝒟(s,\mathrm{}),$$ (9) where $`𝒟`$ is some function of $`s`$ and possibly other variables as indicated by the ellipsis. Combining (9) with the compatibility condition (3), the elasticity equations (5), and force balance (6), we find $$𝒟(s,\mathrm{})+\frac{(1\nu )}{\mu }\dot{s}=\frac{C(t)}{r^2},$$ (10) where $`C(t)`$ is an $`r`$-independent constant of integration. Further analysis using the boundary conditions (7) and (8) at $`r=R`$ yields $$C(t)=\left[1+\left(\frac{12\nu }{\mu }\right)s(R,t)\right]R\dot{R};$$ (11) and $$\left[1\frac{1}{\mu }s(R,t)\right]\frac{\dot{R}}{R}=2_R^{\mathrm{}}\frac{dr}{r}𝒟[s(r,t),\mathrm{}].$$ (12) Eqs. (10) through (12), supplemented by equations of motion for other arguments of $`𝒟(s,\mathrm{})`$, constitute a coupled set of first-order differential equations suitable for computing the time evolution of $`s(r,t)`$ and $`R(t)`$. If $`R(t)`$ is the only length scale in the problem, we can look for self-similar solutions in which $`\dot{R}/R=\omega =\mathrm{constant}`$ so that the hole radius is growing exponentially. All functions depend only on $`\xi =r/R(t)`$. Combining the expression for $`C(t)`$ in (11) with (10), and transforming to the scaling variable $`\xi `$, we find $$𝒟(\stackrel{~}{s},\mathrm{})\frac{(1\nu )}{\mu }\omega \xi \frac{d\stackrel{~}{s}}{d\xi }=\frac{\omega }{\xi ^2}\left[1+\left(\frac{12\nu }{\mu }\right)\stackrel{~}{s}(1)\right].$$ (13) Force balance (6) plus the boundary condition (7) imply $$\sigma _{\mathrm{}}=2_1^{\mathrm{}}\frac{d\xi }{\xi }\stackrel{~}{s}(\xi ).$$ (14) If these self-similar solutions exist for $`\omega >0`$, they describe unbounded plastic failure of the material. ## III Conventional Theories In a typical time-independent approach to this problem , one assumes that there exists a maximum value of $`s`$, say $`s_y`$, and that, in any sufficiently slow deformation, the material adjusts its state so that $`ss_y`$ everywhere. Technically, the condition $`s=s_y`$ is a special case of a Tresca yield surface in the space of stress components. For present purposes, we take this assumption to mean that the hole is surrounded by a plastic zone, $`R<r<R_1`$, within which the force-balance equation (6) remains valid but the condition $`s=s_y`$ replaces Hookean elasticity (5). Outside this zone, $`r>R_1`$, $`p=\sigma _{\mathrm{}}`$ and $`s=s_yR_1^2/r^2`$. Continuity of stress at $`R_1`$ means that, within the zone, $`p=\sigma _{\mathrm{}}2s_y\mathrm{ln}(r/R_1)`$. Then the boundary condition (7) at $`r=R`$ implies that $$\mathrm{ln}\frac{R_1}{R}=\frac{1}{2}\left(\frac{\sigma _{\mathrm{}}}{s_y}1\right).$$ (15) Thus, these assumptions predict that a stationary state with a non-vanishing plastic zone exists for $`\sigma _{\mathrm{}}>s_y`$. A calculation of the displacements similar to that described by Hill shows that the hole radius $`R`$ diverges as $`\sigma _{\mathrm{}}`$ approaches an upper threshold stress $`\sigma _{\mathrm{}}^{th}`$ which (for the case $`s_y/\mu 1`$) is given by: $$\frac{\sigma _{\mathrm{}}^{th}}{s_y}=1+\mathrm{ln}\left(\frac{\mu }{2s_y(1\nu )}\right).$$ (16) To see what happens above this stress, we must consider a time-dependent theory. The simplest conventional time-dependent hypothesis is the Bingham law which, for $`s0`$, we can write in the form: $$𝒟(s)=\{\begin{array}{cc}\alpha (ss_y)\hfill & \text{for }ss_y\hfill \\ 0\hfill & \text{otherwise,}\hfill \end{array}$$ (17) where $`\alpha `$ is a response coefficient. Essentially by definition, the combination of (17) with (10)–(12) describes an elastic perfectly-plastic material whose stationary states, for $`s_y<\sigma _{\mathrm{}}<\sigma _{\mathrm{}}^{th}`$, are the same as those described in the preceding paragraph. We have checked that these states are stable attractors by using (17) to integrate (10)–(12), using the initial condition $`s(r,0)=\sigma _{\mathrm{}}R^2(0)/r^2`$ (for $`r>R(0)`$). We found no surprises; a plastic zone consistent with (15) forms around the growing hole. We also can compute the self-similar solutions of (13) for the Bingham model. For ease of analysis, we write these for the case of incompressible elasticity, $`\nu =1/2`$. In the outer, elastic region, $`\xi >\xi _1=(\mu /s_y)^{1/2}`$: $$\stackrel{~}{s}(\xi )=s_y\left(\frac{\xi _1}{\xi }\right)^2;$$ (18) and, in the plastic region, $`ss_y`$, $`1<\xi <\xi _1`$: $$\stackrel{~}{s}(\xi )=s_y+\left(\frac{\mu \omega }{\mu \alpha +\omega }\right)\frac{1}{\xi ^2}\left[1\left(\frac{\xi }{\xi _1}\right)^{\beta +2}\right],$$ (19) where $`\beta (\omega )=2\mu \alpha /\omega `$. Note that the exponent $`\beta `$ becomes indefinitely large in the limit of small $`\omega `$; thus the second term in the square brackets in (19) produces a function $`\stackrel{~}{s}(\xi )`$ that is sharply bent but continuous at $`\xi =\xi _1`$. To complete the calculation, we use (19) to evaluate the right-hand side of (14). After rearranging and taking the limit of small $`\omega `$, we find: $$\omega \frac{\alpha \mu }{s_y}(\sigma _{\mathrm{}}\sigma _{\mathrm{}}^{th}),$$ (20) where $`\sigma _{\mathrm{}}^{th}`$ is the upper threshold defined in (16). Thus, as expected, the dynamic failure modes start where the time-independent theory breaks down. ## IV The STZ Model ### A Basic properties For present purposes, it will be sufficient to use the simplified version of the STZ model that we introduced in an earlier one-dimensional analysis . As in the Bingham case, the model is specified by the function $`𝒟(s,\mathrm{})`$ defined in (9). For two-dimensions, with circular symmetry, we write: $$𝒟(s,\mathrm{\Delta })=\frac{1}{\tau }(\lambda s\mathrm{\Delta }),$$ (21) and supplement this by an equation of motion for the state variable $`\mathrm{\Delta }`$: $$\dot{\mathrm{\Delta }}=𝒟(s,\mathrm{\Delta })(1\gamma s\mathrm{\Delta }).$$ (22) $`\mathrm{\Delta }(r,t)`$ is the single independent element of a diagonal, traceless tensor which describes the local anisotropy of the shear-transformation zones. We are omitting the other state variable in that describes the density of STZ’s on the assumption that this quantity quickly reaches its equilibrium value. More importantly, this simplified version of the STZ model omits the strongly $`s`$-dependent rate factor that governs memory effects. This version is qualitatively sensible if we load the system only once in one direction, but it does not behave properly if the loading is cycled in any way. The inverse stress $`\gamma `$ can be eliminated in favor of a group of parameters that plays the role of a dynamic yield stress, specifically, $`s_y=1/\sqrt{\lambda \gamma }`$. Note what is happening here for spatially uniform situations. For $`s<s_y`$, $`\mathrm{\Delta }(t)`$ has its stable fixed points on the line $`\mathrm{\Delta }=\lambda s`$, where $`𝒟\dot{\epsilon }_{\theta \theta }^{pl}=0`$. In this region, the material is nonlinearly viscoelastic. For $`ss_y`$ or, equivalently, $`\gamma 0`$, it obeys a conventional creep-compliance law : $$\epsilon ^{tot}(t)=(1+\lambda )s(t)\lambda _{\mathrm{}}^t𝑑t^{}\mathrm{exp}\left[\frac{1}{\tau }(tt^{})\right]\dot{s}(t^{}).$$ (23) We obtain a particularly important result by supposing that the system is initially in a state with $`\epsilon _{\theta \theta }^{pl}=\mathrm{\Delta }=0`$ and that a stress $`s<s_y`$ is applied suddenly at time $`t=0`$. A simple calculation then yields: $$\epsilon _{\mathrm{final}}^{pl}\epsilon _{\theta \theta }^{pl}(t\mathrm{})=\frac{\lambda s_y^2}{s}\mathrm{ln}\left(1\frac{s^2}{s_y^2}\right).$$ (24) $`\epsilon ^{pl}(t)`$ approaches $`\epsilon _{\mathrm{final}}^{pl}`$ exponentially in time with a relaxation time $`\tau _{\mathrm{relax}}`$ that diverges as $`ss_y`$: $$\tau _{\mathrm{relax}}=\frac{\tau }{1(s/s_y)^2}.$$ (25) Eq. (24) is a strain-hardening curve; that is, $`\epsilon _{\mathrm{final}}^{pl}`$ is the non-recoverable plastic strain produced after an infinitely long time by the deviatoric stress $`s`$. In the limit $`\lambda 0`$, with $`s_y`$ held constant, $`\epsilon _{\mathrm{final}}^{pl}`$ vanishes for $`s<s_y`$ but can have any positive value for $`ss_y`$. Thus the parameter $`\lambda `$ is a measure of the deviation from perfect plasticity. The diverging relaxation time near $`s=s_y`$, however, has no simple analog in conventional descriptions of strain hardening. This one-parameter fit (24) to the shape of the strain-hardening curve is much too simple even for the fully nonlinear STZ model, where the behaviors at small stresses and at stresses near $`s_y`$ are determined by different groups of parameters. Moreover, the small-$`s`$ behavior of this $`t\mathrm{}`$ curve is not what is measured experimentally. In real materials and in the full STZ theory, the plastic deformation rates at small stresses are too small to be observed, and the material behaves as if it were purely elastic. For an illustration of this behavior, see Fig. 5 in . For $`s>s_y`$, $`\mathrm{\Delta }`$ goes to $`1/\gamma s`$ and $$\dot{\epsilon }_{\theta \theta }^{pl}\frac{\lambda }{\tau s}\left(s^2s_y^2\right)\frac{2\lambda }{\tau }(ss_y).$$ (26) In dynamic situations at large stress, therefore, the STZ model looks like a Bingham plastic. In short, even this highly simplified version of the STZ model describes much of both static and dynamic plasticity. ### B Dynamically growing hole We return now to the circle problem. As a first investigation, we have numerically integrated (10)–(12), supplemented by (22), to find the time evolution of $`s(r,t)`$, $`\mathrm{\Delta }(r,t)`$, and $`R(t)`$. In Figs. 1 and 2, we show what happens if we suddenly apply the stress $`s(r,0)=\sigma _{\mathrm{}}R^2(0)/r^2`$. We set $`\lambda s_y=0.005`$, $`s_y=0.1\mu `$, and $`\sigma _{\mathrm{}}=0.2\mu `$ and assume a previously undeformed system, $`\mathrm{\Delta }(r,0)=0`$. For these values of the parameters, the hole grows for a while and then stops, and a plastic zone with $`ss_y`$, $`\mathrm{\Delta }\lambda s_y`$ forms around it. Apart from the fact that the outer boundary of the plastic zone is smooth rather than sharply defined, this static limit of the time-dependent deformation is qualitatively consistent with conventional, time-independent plasticity theory. In Fig. 3, we show an analogous set of curves for a substantially larger value of $`\lambda `$, specifically, $`\lambda s_y=0.5`$. According to (24), this system deviates appreciably from perfect plasticity. A plastic zone does form around the hole, but it has no sharp outer boundary. If we estimate the position of this boundary, say, by finding the point of inflection in the final curve $`s(r)`$, we find that the relation (15) is strongly violated. As in the Bingham model discussed in Section III, the STZ model predicts the existence of unbounded failure modes for sufficiently large $`\sigma _{\mathrm{}}`$. The stationary states of the kind shown in Figs. 1 and 3 cease to exist beyond $`\sigma _{\mathrm{}}`$. Indeed, as we show in Fig. 4) the radius of the hole diverges at $`\sigma _{\mathrm{}}^{th}`$. We find these dynamically growing states via the scaling analysis of Eqs. (13) and (14). It is useful to make the following changes of variables: $$\frac{R^2(t)}{r^2}=\frac{1}{\xi ^2}=\zeta ;s(r,t)=s_y\psi (\zeta );\mathrm{\Delta }(r,t)=\lambda s_y\phi (\zeta ).$$ (27) We find: $$\psi \phi +\frac{2(1\nu )\omega \tau }{\mu \lambda }\zeta \frac{d\psi }{d\zeta }=\frac{\omega \tau }{\lambda s_y}\zeta \left[1+(12\nu )\frac{s_y}{\mu }\psi (1)\right].$$ (28) The equation of motion for $`\mathrm{\Delta }`$, i.e. (22), becomes: $$2\omega \tau \zeta \frac{d\phi }{d\zeta }=(\psi \phi )(1\phi \psi ).$$ (29) Finally, (14) becomes: $$\sigma _{\mathrm{}}=s_y_0^1\frac{d\zeta }{\zeta }\psi (\zeta ).$$ (30) As a first step in interpreting these equations, we compute the threshold $`\sigma _{\mathrm{}}^{th}`$ by taking the limit $`\omega 0`$. To avoid unnecessary complication, we again set $`\nu =1/2`$. In this case, (13) and (29) reduce to $`\psi \phi `$ and, after a simple integration: $$\lambda s_y\mathrm{ln}\left(\frac{1+\psi }{1\psi }\right)+\frac{s_y}{\mu }\psi =\zeta .$$ (31) Note that, if $`s_y/\mu 1`$ (which is generally true for realistic situations), then $$\psi (\zeta )\mathrm{tanh}\left(\frac{\zeta }{2\lambda s_y}\right).$$ (32) This solution exhibits a plastic zone with a smooth elastic-plastic boundary only for $`\lambda s_y1`$, in which case there is a region between $`\zeta =2\lambda s_y`$ and $`\zeta =1`$ in which $`\psi 1`$. It is easy to compute $`\psi (\zeta )`$ without making the latter approximation and, via (30), to obtain the threshold stress $`\sigma _{\mathrm{}}^{th}`$. If $`(s_y/\mu )(1+2\lambda \mu )1`$, then $$\frac{\sigma _{\mathrm{}}^{th}}{s_y}1+\mathrm{ln}\left(\frac{\mu }{s_y}\right)\mathrm{ln}(1+2\lambda \mu ),$$ (33) which agrees with (16) when $`\lambda =0`$ and $`\nu =1/2`$, i.e. in the limit of perfect plasticity. Note that we have chosen the parameters in Fig. 1 to lie within this range. If, on the other hand, $`\lambda s_y1`$, then $`\sigma _{\mathrm{}}^{th}1/2\lambda `$. Here the threshold lies below $`s_y`$; the material is highly deformable and conventional plasticity theory has no range of validity. We illustrate this point in Fig. 5. Expanding the solutions of (28) and (29) to first order in $`\omega `$, we find the following behavior near threshold: For $`s_y/\mu 2\lambda s_y1`$, $$\omega \frac{2\lambda }{\tau }[1+2\lambda s_y\mathrm{ln}(\lambda s_y)](\sigma _{\mathrm{}}\sigma _{\mathrm{}}^{th});$$ (34) and, for $`\lambda s_y1`$, $$\omega \frac{\lambda }{\tau }(\sigma _{\mathrm{}}\sigma _{\mathrm{}}^{th}).$$ (35) In each of the last two equations, the quantity $`\sigma _{\mathrm{}}^{th}`$ has the value computed in the corresponding limit in the previous paragraph. We have studied the behavior of the similarity solutions (28)–(30) numerically. Just as in the Bingham plastic, the deviatoric stress at the surface of the hole in the STZ material grows with the hole expansion rate $`\omega `$ as shown in Fig. 6. While intuitively obvious, this result is relevant to understanding stress transmission to brittle crack tips. In Fig. 7 we show the $`\lambda `$-dependence of the stress for a slowly expanding hole. This solution is essentially the $`\omega 0`$ limit obtained in Eq. (31). In a “softer” material with larger $`\lambda `$, the plastic zone shrinks and disappears completely for a large enough $`\lambda .`$ This softening of the material for larger $`\lambda `$ leads to the decrease in the threshold stress $`\sigma _{\mathrm{}}^{th}`$ above which dynamic failure modes exist. ## V Discussion Our principal point is that the STZ model, a simple example of a rate-and-state theory, provides an extremely compact and physically motivated description of essentially all of plasticity theory, both static and time dependent. In just two constitutive relations, (21) and (22), containing just two dimensionless groups of parameters, $`s_y/\mu `$ and $`\lambda s_y`$, plus one time constant $`\tau `$, we capture linear viscoelasticity at small stress, strain hardening at larger stress, and a dynamic transition to viscoplasticity at a yield stress. All of these properties have been described previously in for homogeneous situations and in for an inhomogeneous one dimensional case. In this paper, our principal interest has been to make contact with conventional theories of plasticity by looking at deformation near a circular hole. As a rule, we recover conventional results when both $`s_y\mu `$ and $`\lambda s_y1`$. The first of these conditions, that the yield stress must be much less than the shear modulus, is generally true for ordinary materials. The second condition, according to (24), says that the plastic strain at which large-scale deformations start to occur must be much smaller than unity. This too is ordinarily satisfied by rigid solids. However, it remains to be seen whether this $`\lambda `$-dependent condition might be modified in more realistic parameterizations of rate-and-state theories. One place where the rate-and-state theory differs qualitatively from conventional plasticity is in the interpretation of the strain hardening curve (24). Here, this curve must be interpreted as the $`t\mathrm{}`$ limit of the dynamic inelastic response to an externally applied stress, not as an instantaneous response that is mathematically equivalent to nonlinear elasticity. Moreover, the parameter $`\lambda `$ that determines the deviation from perfect plasticity in (24) is the same parameter that appears in the viscoplastic law (26), and is also the same parameter that determines both the shape of the plastic zone and the threshold $`\sigma _{\mathrm{}}^{th}`$ at which static solutions give way to time-dependent, unbounded failure modes. Perhaps the single most important advantage of the STZ theory is that the material in the plastic zone is characterized not just by the stress and corresponding displacement fields but also by the state variable $`\mathrm{\Delta }`$. This variable tells us in a natural way that the material inside the zone will respond differently to subsequent changes in stress than will the undeformed material outside it. To take advantage of this feature, however, we must go beyond the truncated version of the STZ model that we have used here and include the strongly nonlinear, $`s`$-dependent rate factor derived in . Without doing any further calculations, we can see how the nonlinear rate factor produces memory effects. Suppose that we start in a stressed configuration such as the one shown in Figs. 1 and 2 in which a plastic zone has formed around the hole; and suppose that we then unload the system by quickly reducing the external stress to zero. At $`s=0`$, the rate factor becomes extremely small; in the absence of a stress that can induce transitions from one orientation to another, the zones remain as they were in the previously stressed state. That is, $`\mathrm{\Delta }(r)`$ remains unchanged on unloading. The system relaxes elastically, but the plastic degrees of freedom do not revert to their earlier values. Accordingly, there must be residual stresses in the region near the hole. The material will “remember” the magnitude and direction of its previous loading because it “knows” the function $`\mathrm{\Delta }(r)`$, which will determine via the nonlinear generalizations of the constitutive equations (21) and (22) what happens when new stresses are applied. The other feature that we have emphasized in this analysis is the dynamic failure that occurs at $`\sigma _{\mathrm{}}^{th}`$, a feature that also occurs in the conventional time-dependent theories. We believe that the dynamic growth modes may provide a clue for solving a long-standing puzzle in the theory of fracture, specifically, the question of how breaking stresses can be transmitted through plastic zones to crack tips. This puzzle pertains to quasi-static crack advance, which would seem to be impossible if, as in conventional plasticity theory, the stresses near a crack tip are constrained to be less than or equal to the yield stress. The new modes, especially those that occur when there is appreciable deviation from perfect plasticity, raise the possibility that, like the growing hole, crack growth could occur via plastic flow near the tip. We hope to explore this possibility in a subsequent report. ###### Acknowledgements. We thank Zhigang Suo for suggesting that we study the hole problem as a way to understand the issues raised in this paper. This research has been supported primarily by U.S. DOE Grants DE-FG03-84ER45108 and DE-FG03-99ER45762, and in part by the MRSEC Program of the NSF under award number DMR96-32716.
no-problem/9906/quant-ph9906093.html
ar5iv
text
# The influence of density of modes on dark lines in spontaneous emission ## I Introduction It is now well understood that spontaneous emission of a quantum system depends crucially on the nature of the reservoir with which the system interacts. Spontaneous emission can be modified in a “tailored” manner by changing the density of modes of the reservoir . An effective method to achieve this is to place atoms in waveguides , microcavities or photonic band gap materials , where the density of modes differs substantially from that of the free space vacuum. The latter has also attracted attention for its potential for modifying the absorption and dispersion properties of a system . In the above studies the typical scheme involves the interaction of a two-level atom with a reservoir with modified density of modes. Spontaneous emission of this two-level system differs considerably from the free space result and the usual Weisskopf-Wigner exponential decay is violated. In specific cases complete inhibition of spontaneous decay has even been predicted . In addition to studies of two-level atoms, there are also investigations of multi-level atoms, where the spontaneous emission from an atomic transition in the modified reservoir influences the spontaneous emission of another atomic transition that interacts with the normal free space vacuum . It is schemes such as these that concerns us in this article. Specifically, we study two distinct multi-level atomic schemes where one transition is considered to decay spontaneously in a modified reservoir and the other decays to a normal free space vacuum. The main interest here is the existence of ‘dark lines’ (complete quenching of spontaneous emission for specific vacuum modes) in the spontaneous emission spectrum of the free space transition due to the coupling of the other transition with the modified reservoir. We note that dark lines in spontaneous emission have been predicted in several laser driven schemes and have been experimentally observed . In this article we employ specific models for the modified reservoir density of modes, such as, for example, that which results from an isotropic photonic band gap model with (or without) defects. We show that dark lines appear in the spontaneous emission spectrum as a consequence of the structure of the modified reservoir. This article is organized as follows: in the next section we consider a three-level, $`\mathrm{\Lambda }`$-type atomic system, with one transition spontaneously decaying in the normal free space vacuum and the other transition decaying in a modified reservoir. We study the spontaneous emission spectrum of the free space transition for four different density of modes of the modified reservoir and show that dark lines can occur in the spectrum due to the structure of the modified reservoir. We note that John and Quang have studied the same atomic system as us, using specifically the appropriate density of modes obtained near the edge of an isotropic photonic band gap \[see Eq. (12)\]. However, in their study they focussed on the phenomenon of dynamical splitting in the spectrum and not on the existence dark lines, which is the main phenomenon discussed in our article. Zeros and splittings in spectra should not be confused, although they relate to each other in the limit of strong coupling. In section III we consider a laser-driven extension of the previous system and show that dark lines can occur in this system, too. In this case the dark lines originate from either laser-induced or modified vacuum-induced mechanisms. Finally, we summarize our findings in section IV. ## II First case: $`\mathrm{\Lambda }`$-type scheme We begin with the study of the $`\mathrm{\Lambda }`$-type scheme, shown in Fig. 1(a). This system is similar to that used by Lewenstein et al and by John and Quang . The atom is assumed to be initially in state $`|2`$. The transition $`|2|1`$ is taken to be near resonant with a modified reservoir (this will be later referred to as the non-Markovian reservoir), while the transition $`|2|0`$ is assumed to be occurring in free space (this will be later referred to as the Markovian reservoir). The spectrum of this latter transition is of central interest in this article. The Hamiltonian which describes the dynamics of this system, in the interaction picture and the rotating wave approximation (RWA), is given by (we use units such that $`\mathrm{}=1`$), $`H`$ $`=`$ $`{\displaystyle \underset{\lambda }{}}g_\lambda e^{i(\omega _\lambda \omega _{20})t}|20|a_\lambda +{\displaystyle \underset{\kappa }{}}g_\kappa e^{i(\omega _\kappa \omega _{21})t}|21|a_\kappa +\text{H.c.}.`$ (1) Here, $`g_\kappa `$ denotes the coupling of the atom with the modified vacuum modes $`(\kappa )`$ and $`g_\lambda `$ denotes the coupling of the atom with the free space vacuum modes $`(\lambda )`$. Both coupling strengths are taken to be real. The energy separations of the states are denoted by $`\omega _{ij}=\omega _i\omega _j`$ and $`\omega _\kappa `$ $`(\omega _\lambda )`$ is the energy of the $`\kappa `$ $`(\lambda )`$-th reservoir mode. The description of the system is given using a probability amplitude approach. We proceed by expanding the wave function of the system, at a specific time $`t`$, in terms of the ‘bare’ state vectors such that $`|\psi (t)=b_2(t)|2,\{0\}+{\displaystyle \underset{\lambda }{}}b_\lambda (t)|0,\{\lambda \}+{\displaystyle \underset{\kappa }{}}b_\kappa (t)|1,\{\kappa \}.`$ (2) Substituting Eqs. (1) and (2) into the time-dependent Schrödinger equation we obtain $`i\dot{b}_2(t)`$ $`=`$ $`{\displaystyle \underset{\lambda }{}}g_\lambda b_\lambda (t)e^{i(\omega _\lambda \omega _{20})t}+{\displaystyle \underset{\kappa }{}}g_\kappa b_\kappa (t)e^{i(\omega _\kappa \omega _{21})t},`$ (3) $`i\dot{b}_\lambda (t)`$ $`=`$ $`g_\lambda b_2(t)e^{i(\omega _\lambda \omega _{20})t},`$ (4) $`i\dot{b}_\kappa (t)`$ $`=`$ $`g_\kappa b_2(t)e^{i(\omega _\kappa \omega _{21})t}.`$ (5) We proceed by performing a formal time integration of Eqs. (4) and (5) and substitute the result into Eq. (3) to obtain the integro-differential equation $`\dot{b}_2(t)`$ $`=`$ $`{\displaystyle _0^t}𝑑t^{}b_2(t^{}){\displaystyle \underset{\lambda }{}}g_\lambda ^2e^{i(\omega _\lambda \omega _{20})(tt^{})}{\displaystyle _0^t}𝑑t^{}b_2(t^{}){\displaystyle \underset{\kappa }{}}g_\kappa ^2e^{i(\omega _\kappa \omega _{20})(tt^{})}.`$ (6) Because the reservoir with modes $`\lambda `$ is assumed to be Markovian, we can apply the usual Weisskopf-Wigner result and obtain $$\underset{\lambda }{}g_\lambda ^2e^{i(\omega _\lambda \omega _{20})(tt^{})}=\frac{\gamma }{2}\delta (tt^{}).$$ (7) Note that the principal value term associated with the Lamb shift which should accompany the decay rate has been omitted in Eq. (7). This does not affect our the results, as we can assume that the Lamb shift is incorporated into the definition of our state energies. For the second summation in Eq. (6), the one associated with the modified reservoir modes, the above result is not applicable as the density of modes of this reservoir is assumed to vary much quicker than that of free space. To tackle this problem, we define the following kernel $$K(tt^{})=\underset{\kappa }{}g_\kappa ^2e^{i(\omega _\kappa \omega _{21})(tt^{})}g^{3/2}𝑑\omega \rho (\omega )e^{i(\omega \omega _{21})(tt^{})}.$$ (8) which is calculated using the appropriate density of modes $`\rho (\omega )`$ of the modified reservoir. In Eq. (8), $`g`$ denotes the coupling constant of the atom to the non-Markovian reservoir. Using Eqs. (7) and (8) into Eq. (6) we obtain $`\dot{b}_2(t)`$ $`=`$ $`{\displaystyle \frac{\gamma }{2}}b_2(t){\displaystyle _0^t}𝑑t^{}b_2(t^{})K(tt^{}).`$ (9) The long time spontaneous emission spectrum in the Markovian reservoir is given by $`S(\delta _\lambda )|b_\lambda (t\mathrm{})|^2`$, with $`\delta _\lambda =\omega _\lambda \omega _{20}`$ . We calculate $`b_\lambda (t\mathrm{})`$ with the use of the Laplace transform of the equations of motion. Using Eq. (4) and the final value theorem we obtain the spontaneous emission spectrum as $`S(\delta _\lambda )\gamma |lim_{si\delta _\lambda }B_2(s)|^2`$, where $`B_2(s)`$ is the Laplace transform of the atomic amplitudes $`b_2(t)`$ and $`s`$ is the Laplace variable. This in turn, with the help of Eq. (9), reduces to $$S(\delta _\lambda )\frac{\gamma }{\left|i\delta _\lambda +\gamma /2+\stackrel{~}{K}(si\delta _\lambda )\right|^2},$$ (10) where $`\stackrel{~}{K}(s)`$ is the Laplace transform of $`K(t)`$, which yields $$\stackrel{~}{K}(s)=g^{3/2}𝑑\omega \frac{\rho (\omega )}{s+i(\omega \omega _{21})}.$$ (11) Therefore, in order to calculate the spontaneous emission spectrum in the Markovian reservoir we need to calculate $`\stackrel{~}{K}(s)`$. This will be done for different models of the density of modes $`\rho (\omega )`$ of the non-Markovian reservoir. We begin by considering the non-Markovian reservoir to be that obtained near the edge of an isotropic photonic band gap model . Then, $$\rho (\omega )=\frac{1}{\pi }\frac{1}{\sqrt{\omega \omega _g}}\mathrm{\Theta }(\omega \omega _g),$$ (12) where $`\omega _g`$ is the gap frequency and $`\mathrm{\Theta }`$ is the Heaviside step function. In this case, Eq. (11) leads to $$\stackrel{~}{K}(s)=\frac{g^{3/2}}{\sqrt{is\delta _g}},$$ (13) with $`\delta _g=\omega _g\omega _{21}`$. The spontaneous emission spectrum then reads $$S(\delta _\lambda )\gamma \left|\frac{\sqrt{\delta _\lambda \delta _g}}{(i\delta _\lambda +\gamma /2)\sqrt{\delta _\lambda \delta _g}+g^{3/2}}\right|^2.$$ (14) Obviously, the spectrum exhibits a zero (i.e. predicts the existence of a dark line), if $`\delta _\lambda =\delta _g`$. This is purely an effect of the above density of states, and the non-Markovian character of the reservoir. In the case of a Markovian reservoir the spontaneous emission spectrum would obtain the well-known Lorentzian profile and no dark line would appear in the spectrum. The behaviour of the spectrum is shown in Fig. 2 for different values of the detuning from the threshold. The spectrum has two well-separated peaks and the dark line appears at the predicted value. We note that a spectrum of this form has also been derived by John and Quang ; however, in their study they did not mention the existence of the dark line (i.e. the zero in the spectrum and its spectral position), which is of central importance in this article. The density of modes of Eq. (12) is a special case of a more general family of density of modes those given by $$\rho (\omega )=\frac{1}{\pi }\frac{\sqrt{\omega \omega _g}}{ϵ+\omega \omega _g}\mathrm{\Theta }(\omega \omega _g),$$ (15) where $`ϵ`$ is usually referred to as the smoothing parameter. Such a density of modes has been used in studies of atoms in waveguides microcavities and photonic band gap materials . Eq. (12) is recovered by taking the limit $`ϵ0`$ in Eq. (15). For the above density of modes, Eq. (15) we obtain $`\stackrel{~}{K}(s)`$ $`=`$ $`{\displaystyle \frac{g^{3/2}}{i\sqrt{ϵ}+\sqrt{is\delta _g}}},`$ (16) $`S(\delta _\lambda )`$ $``$ $`\gamma \left|{\displaystyle \frac{i\sqrt{ϵ}+\sqrt{\delta _\lambda \delta _g}}{(i\delta _\lambda +\gamma /2)(i\sqrt{ϵ}+\sqrt{\delta _\lambda \delta _g})+g^{3/2}}}\right|^2.`$ (17) So, in this “smoothed” case, no dark line appears in the spectrum. This is shown in Fig. 3, where the zero disappears from the spectrum. The only case for which spontaneous emission spectrum will give zero is that of $`ϵ=0`$ and $`\delta _\lambda =\delta _g`$ which, of course, reduces to the previous result of Eq. (14). The above two density of modes describe “pure” materials (i.e. materials with no defects). In reality, defects exist in waveguides, microcavities or photonic band gap materials which exhibit gaps in their density of modes. These defects lead to narrow-linewidth, well-localized modes in the gaps of the above structures. Their density of modes can be described by either a delta function or a narrow Lorentzian . We will now combine the density of modes given by Eq. (12) and that given by a defect structure. In the case that we choose a delta function density of modes for the defect structure we obtain $`\stackrel{~}{K}(s)`$ $`=`$ $`{\displaystyle \frac{g^{3/2}}{\sqrt{is\delta _g}}}+{\displaystyle \frac{g_1^2}{s+i\delta _c}},`$ (18) $`S(\delta _\lambda )`$ $``$ $`\gamma \left|{\displaystyle \frac{\sqrt{\delta _\lambda \delta _g}}{(i\delta _\lambda +\gamma /2)\sqrt{\delta _\lambda \delta _g}+g^{3/2}+ig_1^2\frac{\sqrt{\delta _\lambda \delta _g}}{\delta _\lambda \delta _c}}}\right|^2,`$ (19) where $`\delta _c=\omega _c\omega _{21}`$ is the defect mode-atom detuning (defect mode at frequency $`\omega _c`$) and $`g_1`$ is the defect mode-atom coupling constant. In this case the spectrum exhibits two dark lines one at $`\delta _\lambda =\delta _g`$ and another at $`\delta _\lambda =\delta _c`$. In Fig. 4 we present the spontaneous emission spectrum described by Eq. (19) for the same parameters that used in Fig. 2. The spectrum obtains now a very pronounced third peak and two dark lines at the predicted values. For a Lorentzian density of modes of the defect structure we get $`\stackrel{~}{K}(s)`$ $`=`$ $`{\displaystyle \frac{g^{3/2}}{\sqrt{is\delta _g}}}+{\displaystyle \frac{g_1^2}{s+i\delta _c+\gamma _c/2}},`$ (20) $`S(\delta _\lambda )`$ $``$ $`\gamma \left|{\displaystyle \frac{\sqrt{\delta _\lambda \delta _g}}{(i\delta _\lambda +\gamma /2)\sqrt{\delta _\lambda \delta _g}+g^{3/2}+g_1^2\frac{\sqrt{\delta _\lambda \delta _g}}{i(\delta _c\delta _\lambda )+\gamma _c/2}}}\right|^2,`$ (21) with $`\gamma _c`$ being the width of the Lorentzian describing the defect mode. As in the case of the density of modes given by Eq. (12), in this case too, there is a single dark line in the spectrum at $`\delta _\lambda =\delta _g`$. However, the behaviour of the spectrum is quite different in this case compared to that shown in Fig. 2, as can be seen in Fig. 5. ## III Second case: Laser-driven scheme In this section we turn to the study of the laser-driven system shown in Fig. 1(b). This system is composed of a ground state $`|3`$ which is coupled by a laser field to the excited state $`|2`$. State $`|2`$ can spontaneously couple to either state $`|0`$ via interaction with a Markovian reservoir, or to state $`|1`$ via interaction with a non-Markovian reservoir, as in the previous section. The spontaneous emission spectrum in the Markovian reservoir is our main concern in this section, too. The Hamiltonian of this system, written in the interaction picture and under the RWA reads $`H`$ $`=`$ $`\mathrm{\Omega }|32|e^{i\delta t}+{\displaystyle \underset{\lambda }{}}g_\lambda e^{i(\omega _\lambda \omega _{20})t}|20|a_\lambda +{\displaystyle \underset{\kappa }{}}g_\kappa e^{i(\omega _\kappa \omega _{21})t}|21|a_\kappa +\text{H.c.}.`$ (22) Here, $`\mathrm{\Omega }`$ is the Rabi frequency (assumed real) and $`\delta =\omega \omega _{23}`$ the laser detuning, with $`\omega `$ being the laser field angular frequency. For the description of this system too we use the probability amplitude approach and expand the wave function of the system as $`|\psi (t)=b_3(t)e^{i\delta t}|3,\{0\}+b_2(t)|2,\{0\}+{\displaystyle \underset{\lambda }{}}b_\lambda (t)|0,\{\lambda \}+{\displaystyle \underset{\kappa }{}}b_\kappa (t)|1,\{\kappa \}.`$ (23) We substitute Eqs. (22) and (23) into the time-dependent Schrödinger equation and apply the elimination of the vacuum modes outlined in the previous section to obtain $`i\dot{b}_3(t)`$ $`=`$ $`\delta b_3(t)+\mathrm{\Omega }b_2(t),`$ (24) $`i\dot{b}_2(t)`$ $`=`$ $`\mathrm{\Omega }b_3(t)i{\displaystyle \frac{\gamma }{2}}b_2(t)i{\displaystyle _0^t}𝑑t^{}b_2(t^{})K(tt^{}),`$ (25) $`i\dot{b}_\lambda (t)`$ $`=`$ $`g_\lambda b_2(t)e^{i(\omega _\lambda \omega _{20})t},`$ (26) $`i\dot{b}_\kappa (t)`$ $`=`$ $`g_\kappa b_2(t)e^{i(\omega _\kappa \omega _{21})t}.`$ (27) The long time spontaneous emission spectrum in the Markovian reservoir is given by $`S(\delta _\lambda )|b_\lambda (t\mathrm{})|^2`$, and is calculated in a closed form with the use of the Laplace transform and the final value theorem as $$S(\delta _\lambda )\gamma \left|\frac{(\delta _\lambda \delta )b_2(0)+\mathrm{\Omega }b_3(0)}{(\delta _\lambda \delta )[\delta _\lambda +i\gamma /2+i\stackrel{~}{K}(si\delta _\lambda )]\mathrm{\Omega }^2}\right|^2.$$ (28) The spectrum depends, in this case too, from the Laplace transform of the kernel, which in turn depends on the density of modes of the modified reservoir \[see Eq. (11)\]. We use the four models of the density of modes described in the previous section to calculate the above spectrum. In the case of an isotropic photonic band gap material, with the use of Eq. (13) we find, $$S(\delta _\lambda )\gamma \left|\frac{(\delta _\lambda \delta )b_2(0)+\mathrm{\Omega }b_3(0)}{(\delta _\lambda \delta )(\delta _\lambda +i\gamma /2)+ig^{3/2}\frac{\delta _\lambda \delta }{\sqrt{\delta _\lambda \delta _g}}\mathrm{\Omega }^2}\right|^2.$$ (29) Let us assume first that the atom starts from its ground state, i.e. $`|b_3(0)|^2=1`$, $`|b_2(0)|^2=0`$. Then, the spectrum obtains a dark line at $`\delta _\lambda =\delta _g`$, except in the case that $`\delta =\delta _g`$, so in the case that the laser detuning becomes equal to the detuning from the band edge no dark line occurs, if the system is initially in state $`|3`$. We note that in the case when the transition $`|2|1`$ occurs in a Markovian reservoir, then no dark line appears in the spectrum . If now the atom starts in a superposition of the ground $`(|3)`$ and excited states $`(|2)`$ with real expansion coefficients $`[b_i(0)]`$ then two dark lines can appear in the spectrum. The first dark line appears at $`\delta _\lambda =\delta \mathrm{\Omega }b_3(0)/b_2(0)`$ and is attributed to the laser-atom interaction . The second dark line occurs at $`\delta _\lambda =\delta _g`$ and is attributed to the interaction with the non-Markovian reservoir. In this case too, the second dark line disappears if $`\delta =\delta _g`$. The behaviour of the above spectrum for the case that the system is initially in state $`|2`$ is shown in Fig. 6. We have chosen $`\delta _g\delta `$ therefore, as predicted, two dark lines appears in the spectrum. We note the similarity of this spectrum and that of Fig. 4. This should be expected as the delta function density of modes (used in Fig. 4) represents a pure Jaynes-Cummings interaction . If the more general, smoothed density of states of Eq. (15) is used (with $`ϵ0`$), the spontaneous emission spectrum is given by $$S(\delta _\lambda )\gamma \left|\frac{(\delta _\lambda \delta )b_2(0)+\mathrm{\Omega }b_3(0)}{(\delta _\lambda \delta )(\delta _\lambda +i\gamma /2)+ig^{3/2}\frac{\delta _\lambda \delta }{i\sqrt{ϵ}+\sqrt{\delta _\lambda \delta _g}}\mathrm{\Omega }^2}\right|^2.$$ (30) Then, in the case that the atom starts in the ground state no dark line exists in the spectrum. However, if the atom starts in a superposition of the ground and excited states (with real expansion coefficients) the laser-induced dark line appears at the same frequency as before, i.e. at $`\delta _\lambda =\delta \mathrm{\Omega }b_3(0)/b_2(0)`$. This is verified in Fig. 7, where only a single dark line appears in the spectrum at $`\delta _\lambda =\delta `$, as $`b_3(0)=0`$. In the case of an isotropic photonic band gap with a defect with delta function density of modes then the spectrum reads $$S(\delta _\lambda )\gamma \left|\frac{(\delta _\lambda \delta )b_2(0)+\mathrm{\Omega }b_3(0)}{(\delta _\lambda \delta )(\delta _\lambda +i\gamma /2)+ig^{3/2}\frac{\delta _\lambda \delta }{\sqrt{\delta _\lambda \delta _g}}g_1^2\frac{\delta _\lambda \delta }{\delta _\lambda \delta _c}\mathrm{\Omega }^2}\right|^2.$$ (31) and one obtains, in general, two dark lines if the atom starts from the ground state, one at $`\delta _\lambda =\delta _g`$ and the other at $`\delta _\lambda =\delta _c`$. Any of these dark lines can disappear if either $`\delta =\delta _g`$, or $`\delta =\delta _c`$ so the spectrum exhibits only one dark line in this case. If now the atom starts in a superposition of the ground $`(|3)`$ and excited states $`(|2)`$ with real expansion coefficients then a third dark line can appear in the spectrum at $`\delta _\lambda =\delta \mathrm{\Omega }b_3(0)/b_2(0)`$. The spectrum for the case that the atom is initially in the excited state $`|2`$ (as in the previous cases), is displayed in Fig. 8. A very rich behaviour of the spectrum is find. The spectrum now has three dark lines at the predicted values and four different peaks can be seen. Finally, in the case of an isotropic photonic band gap with a defect with Lorentzian density of modes the spectrum gets the form, $$S(\delta _\lambda )\gamma \left|\frac{(\delta _\lambda \delta )b_2(0)+\mathrm{\Omega }b_3(0)}{(\delta _\lambda \delta )(\delta _\lambda +i\gamma /2)+ig^{3/2}\frac{\delta _\lambda \delta }{\sqrt{\delta _\lambda \delta _g}}+ig_1^2\frac{\delta _\lambda \delta }{i(\delta _c\delta _\lambda )+\gamma _c/2}\mathrm{\Omega }^2}\right|^2.$$ (32) The dark lines in the spectrum appear at the same frequencies as those of the simple isotropic photonic band gap model, discussed above. However, the shape of the spectrum in this case differs from that of Fig. 6 as it is shown in Fig. 9. ## IV Summary In this article we have investigated the spontaneous emission properties of two distinct atomic models with one transition coupling to a Markovian reservoir while another transition coupling to a non-Markovian reservoir. Of specific interest to us were the existence of dark lines in the Markovian spontaneous emission spectrum. We have shown that dark lines can occur if the non-Markovian reservoir is described by certain densities of modes. In the case of the laser-driven scheme of Fig. 1(b) laser-induced dark lines can co-exist with non-Markovian reservoir-induced dark lines. Overall, spontaneous emission in the Markovian transition can be efficiently controlled (and even suppressed) by appropriately engineering the density of modes of the non-Markovian reservoir. ## Acknowledgments E.P. would like to thank Niels J. Kylstra for useful discussions in the subject. The contribution of D.G.A. to this work was done in partial fulfilment of the MSc. requirements at the Department of Physics, University of Crete, Greece. This work was funded by the UK Engineering and Physical Sciences Research Council (EPSRC) and the European Commission Cavity QED TMR Network ERBFMRXCT96066. D.G.A. wishes to acknowledge the financial support of the Hellenic State Scholarship Foundation (SSF).
no-problem/9906/nucl-th9906065.html
ar5iv
text
# Particle Ratios at CERN, BNL and GSI : Unified description of Freeze-Out Parameters.11footnote 1Presented at Quark Matter 99, Torino, Italy ## 1 Particle Ratios : General Remarks Information about the chemical freeze-out parameters, the temperature, $`T_{ch}`$, and baryon chemical potential $`\mu _B^{ch}`$ can be obtained from ratios of integrated particle yields . This is because various effects like transverse flow or particle production from a superposition of fireballs cancel out in such ratios provided the freeze-out parameters are unique. We will discuss in succession particle ratios at GSI/SIS, BNL/AGS and SPS/CERN. We will then discuss common properties, in particular, the observation that all of them correspond to an average energy of 1 GeV per hadron independent of the beam energy . ## 2 Particle ratios at GSI/SIS A systematic study of the particle ratios measured at GSI/SIS energies has been done recently . In figure 1 we show the results obtained in for Ni-Ni at 1.0 AGeV. The particle ratios are consistent with the interpretation that the hadronic composition of the final state is fixed at a unique temperature and baryon chemical potential. Since the temperature is low, the number of particles created is small and it is therefore necessary to take into account the exact conservation of strangeness because strange particles are always produced in pairs and it is more difficult to create two particles in a small cold system than it is to create one particle. There is no necessity to introduce any other parameters and a good fit can be achieved with full chemical equilibrium including strange particles, i.e. with $`\gamma _s=1`$. In view of the fact that kaons are produced below threshold at SIS this is remarkable. ## 3 Particle ratios at BNL/AGS Particle ratios at BNL/AGS have been reanalyzed recently in references and . The results are consistent with those obtained previously, e.g. ref. obtains $`T_{ch}=118.4\pm 11.6`$ MeV and $`\mu _B^{ch}522`$ MeV. The ratios together with error bands are shown in figure 2 . This temperature is about double the one extracted from SIS results, the baryon chemical potential is clearly much smaller. Because of the higher temperature, it is no longer necessary to treat strangeness in a special way and corrections from the more exact canonical treatment are negligible. There is agreement that a good fit can be achieved with $`\gamma _s=1`$ and the data are again consistent with chemical equilibrium. ## 4 Particle ratios at CERN/SPS Several papers have appeared recently analyzing the particle ratios measured in Pb-Pb collisions at CERN . In reference it is found that a strangeness suppression factor $`\gamma _s=0.55`$ is needed to describe the data while in reference it is argued that a good description is possible using $`\gamma _s=1`$. Despite this serious difference both papers arrive at very similar results : $`T_{ch}=165`$ MeV in vs $`T_{ch}=170`$ MeV in , i.e. the difference is less than 3 percent. The disagreement over the value of $`\gamma _s`$ will be settled when more precise data become available, indications from this conference are that a value below 1 is necessary but a full analysis is still outstanding. The analysis of reference introduces new parameters and is not directly comparable to the one discussed here. ## 5 Particle Ratios : Common Properties at SPS, AGS and SIS The results from GSI, BNL and CERN show a striking systematic behavior (see figure 3) : the GSI/SIS results have the lowest temperature and the highest baryon chemical potential, as the beam energy is increased a clear shift towards higher temperatures and lower baryon chemical potentials occurs. The points have in common that the average energy per hadron is approximately 1 GeV . Chemical freeze-out is thus reached when the energy per particle drops below 1 GeV per hadron. When this value is reached inelastic collisions are no longer important and the abundances of the various hadronic species are fixed. The consequences of this are discussed in detail in reference . Recent results from E895 collaboration presented at this conference are compatible with the above result. ## 6 Conclusions Bringing together results obtained at very different beam energies shows that the hadronic abundances seen in the final state of relativistic heavy ion collisions are fixed once the average energy per hadron drops below 1 GeV. We expect the results from the SPS beams at 40 and 80 GeV to follow this observation. It will be interesting to see how the results from RHIC will relate to this observation. We thank P. Braun-Munzinger, B. Friman, W. Nörenberg, H. Oeschler, H. Satz and J. Stachel for fruitful discussions.
no-problem/9906/hep-ph9906297.html
ar5iv
text
# References BA-99-49 FERMILAB-Pub-99/159-T hep–ph/9906297 June 1999 Bimaximal Mixing in an $`SO(10)`$ Minimal Higgs Model Carl H. Albright<sup>1</sup> Department of Physics Northern Illinois University, DeKalb, IL 60115 and Fermi National Accelerator Laboratory P.O. Box 500, Batavia, IL 60510 S.M. Barr<sup>2</sup> Bartol Research Institute University of Delaware Newark, DE 19716 ## Abstract An $`SO(10)`$ SUSY GUT model was previously presented based on a minimal set of Higgs fields. The quark and lepton mass matrices derived fitted the data extremely well and led to large $`\nu _\mu \nu _\tau `$ mixing in agreement with the atmospheric neutrino data and to the small-angle MSW solution for the solar neutrinos. Here we show how a slight modification leading to a non-zero up quark mass can result in bimaximal mixing for the atmospheric and solar neutrinos. The “just-so” vacuum solution is slightly favored over the large-angle MSW solution on the basis of the hierarchy required for the right-handed Majorana matrix and the more nearly-maximal mixing angles obtained. PACS numbers: 12.15.Ff, 12.10.Dm, 12.60.Jv, 14.60.Pq Key words: bimaximal mixing, quark and lepton mass matrices <sup>1</sup>E-mail: albright@fnal.gov; <sup>2</sup>E-mail: smbarr@bartol.udel.edu Atmospheric neutrino data reveal a large and, in fact, nearly maximal mixing between $`\nu _\mu `$ and some other neutrino, which most plausibly is assumed to be $`\nu _\tau `$, though it may also be a sterile neutrino. The solar neutrino problem , on the other hand, can be solved either by a small mixing ($`\mathrm{sin}^22\theta _{e\mu }6\times 10^3`$) or by a nearly maximal mixing of $`\nu _e`$ with $`\nu _\mu `$. In the former case one has the small-angle MSW solution , while in the latter case one has either the large-angle MSW solution or the vacuum oscillation solution . Cases where both the atmospheric and solar neutrino anomalies are solved by nearly maximal mixing are often called “bimaximal.” There have been several interesting suggestions about how bimaximal mixing might arise theoretically . Here we suggest a scheme that has certain novel features. The basic idea, which at first sounds artificial but which actually can emerge quite simply and naturally as will be seen, is that the large mixing of $`\nu _\mu `$ and $`\nu _\tau `$ originates from transformation of the charged lepton mass matrix, whereas the large mixing of $`\nu _e`$ and $`\nu _\mu `$ originates in the neutrino mass matrix itself. (Of course, it only makes sense to draw this distinction if there is a preferred basis of families, which is here provided by some underlying theory of flavor.) The idea that the large $`\nu _\mu \nu _\tau `$ mixing arises from the mass matrix of the charged leptons was proposed in , where it emerged as part of a complete model for the quark and lepton masses and mixings. One of the virtues of this idea is that it allows a simple resolution of the supposed paradox that the mixing of the second and third families is small for the quarks, $`V_{cb}0.04`$, and large for the leptons, $`V_{\mu 3}0.7`$. The model developed in was only a model for the heaviest two families with the first family being approximated as massless. In this model was extended to the first family in a very economical way that gave several additional predictions, among which were that the $`\nu _e\nu _\mu `$ mixing angle is small, and in fact precisely in the presently allowed range for the small-angle MSW solution. This small $`\nu _e\nu _\mu `$ angle, like the large $`\nu _\mu \nu _\tau `$ angle, arose from diagonalization of the charged lepton mass matrix. In the present paper we show that a slight alteration of that model leads to an equally predictive scheme that has bimaximal mixing. All the predictions for quark and charged lepton masses, for the CKM parameters, and for the $`\nu _\mu \nu _\tau `$ mixing are left essentially unaffected; however, the $`\nu _e\nu _\mu `$ mixing moves from the small-angle MSW value to become maximal. We will first very briefly review the model of and , noting the features relevant to lepton mixing, and then proceed to show how bimaximal mixing can arise in it. The model is based on supersymmetric $`SO(10)`$ and leads to the following matrices: $`U^0`$ for up-type quarks, $`D^0`$ for down-type quarks, $`L^0`$ for the charged leptons, and $`N^0`$ for the Dirac neutrino masses, where the superscript 0 refers to the matrices at the unification scale. $$\begin{array}{cc}U^0=\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& ϵ/3\\ 0& ϵ/3& 1\end{array}\right)M_U,\hfill & D^0=\left(\begin{array}{ccc}0& \delta & \delta ^{}\\ \delta & 0& \sigma +ϵ/3\\ \delta ^{}& ϵ/3& 1\end{array}\right)M_D,\hfill \\ & \\ N^0=\left(\begin{array}{ccc}0& 0& 0\\ 0& 0& ϵ\\ 0& ϵ& 1\end{array}\right)M_U,\hfill & L^0=\left(\begin{array}{ccc}0& \delta & \delta ^{}\\ \delta & 0& ϵ\\ \delta ^{}& \sigma +ϵ& 1\end{array}\right)M_D.\hfill \end{array}$$ (1) These matrices arise from simple diagrams in the $`SO(10)`$ unified model, which involve only five effective Yukawa terms that have the forms $`(\mathrm{𝟏𝟔}_3\mathrm{𝟏𝟔}_3)\mathrm{𝟏𝟎}_H`$, $`(\mathrm{𝟏𝟔}_2\mathrm{𝟏𝟔}_3)\mathrm{𝟏𝟎}_H\mathrm{𝟒𝟓}_H`$, $`[\mathrm{𝟏𝟔}_2\mathrm{𝟏𝟔}_H]`$ $`[\mathrm{𝟏𝟔}_3\mathrm{𝟏𝟔}_H^{}]`$, $`[\mathrm{𝟏𝟔}_1\mathrm{𝟏𝟔}_2][\mathrm{𝟏𝟔}_H\mathrm{𝟏𝟔}_H^{}]`$, and $`[\mathrm{𝟏𝟔}_1\mathrm{𝟏𝟔}_3][\mathrm{𝟏𝟔}_H\mathrm{𝟏𝟔}_H^{}]`$. These lead, respectively, to the entries in Eq. (1) that are denoted $`1`$, $`ϵ`$, $`\sigma `$, $`\delta `$, and $`\delta ^{}`$. The numerical subscripts are family indices, while the subscript H denotes a Higgs multiplet. The notation $`[\mathbf{16\hspace{0.33em}16}]`$ implies the spinors are contracted into an $`SO(10)`$ vector. The VEV, $`\mathrm{𝟏𝟔}_HM_G`$, lies in the $`SU(5)`$ singlet direction and helps break $`SO(10)`$ down to the Standard Model, while $`\mathrm{𝟏𝟔}_H^{}M_W`$ lies in the weak doublet direction and helps break the electroweak interactions. The expectation value of the adjoint ($`\mathrm{𝟒𝟓}_H`$) is proportional to the $`SO(10)`$ generator $`BL`$, as required for the Dimopoulos-Wilczek mechanism to provide the doublet-triplet splitting . The foregoing information is sufficient to derive the matrices in Eq. (1). The family hierarchy results from the smallness of the parameters $`ϵ0.14,\delta 0.008`$, and $`\left|\delta ^{}\right|0.008`$. The parameter $`\sigma 1.8`$ is not small, however, and is the key to understanding many of the qualitative and quantitative features of the quark and lepton spectrum, including the fact that the second and third families of leptons have a large mixing while that of the quarks is small. The point is that $`SU(5)`$ relates left-handed (right-handed) down quarks to right-handed (left-handed) charged leptons, and consequently relates the $`ij`$ element of $`D^0`$ to the $`ji`$ element of $`L^0`$. That is why in Eq. (1) the large parameter $`\sigma `$ appears in $`L_{32}^0`$ where it leads to large $`\mu ^{}\tau ^{}`$ mixing and hence large $`\mathrm{sin}^22\theta _{\mu \tau }`$, whereas it does not appear in $`D_{32}^0`$ where it would give large $`V_{cb}`$, but rather appears in $`D_{23}^0`$ where it affects only the unobservable mixing of right-handed quarks. The great difference in magnitude between $`V_{cb}`$ and $`\mathrm{sin}^22\theta _{\mu \tau }`$ is thus a consequence of $`D^0`$ and $`L^0`$ being highly asymmetric and the peculiarities of $`SU(5)`$ invariance. As shown in and , the matrices in Eq. (1) give a remarkably good fit to all the known quark and lepton masses and mixings, in particular, when the small-angle MSW solution is relevant for the solar neutrino problem. We now review the lepton results and explore the possibility of large-angle $`\nu _e\nu _\mu `$ mixing in this $`SO(10)`$ unified model. The lepton mixing matrix is given by $$V_{lepton}=U_L^{}U_\nu ,$$ (2) where $`U_L`$ is the unitary transformation of the left-handed charged leptons required to diagonalize $`L^0`$, and $`U_\nu `$ is the complex orthogonal transformation of the left-handed neutrinos required to diagonalize the light-neutrino mass matrix, $`M_\nu =N^TM_R^1N`$. In $`SO(10)`$ the Dirac neutrino mass matrix $`N^0`$ and up quark mass matrix $`U^0`$ are related, and as given in Eq. (1) have vanishing first rows and first columns. This is to be regarded as only an approximation to the real world, but as we shall see later, it is a very good approximation. With this particular texture for $`N^0`$, no matter what form $`M_R`$ assumes, the light-neutrino mass matrix $`M_\nu `$ will also have vanishing first row and column, and we can write $$U_\nu =\left(\begin{array}{ccc}1& 0& 0\\ 0& c& s\\ 0& s& c\end{array}\right),$$ (3) where the parameters $`s`$ and $`c`$ are complex, in general, with $`c^2+s^2=1`$. From the form of $`N^0`$ it is easy to see that, formally speaking, $`\left|s\right|=O(ϵ)`$. If the unknown matrix $`M_R`$ is parametrized by $`(M_R^1)_{ij}=a_{ij}\mathrm{\Lambda }_R^1`$, then, in units of $`(M_U^2/\mathrm{\Lambda }_R)`$, one has $`(M_\nu )_{1j}=(M_\nu )_{j1}=0`$, $`(M_\nu )_{22}=ϵ^2a_{33}`$, $`(M_\nu )_{23}=(M_\nu )_{32}=ϵa_{33}ϵ^2a_{23}`$, and $`(M_\nu )_{33}=a_{33}2ϵa_{23}+ϵ^2a_{22}`$. This gives $$\mathrm{tan}2\theta _{23}^\nu 2sc/(c^2s^2)=2ϵ\left(\frac{a_{33}ϵa_{23}}{a_{33}2ϵa_{23}+ϵ^2a_{22}ϵ^2a_{33}}\right),$$ (4) and thus $`\theta _{23}^\nu ϵ`$, unless the parameters $`a_{ij}`$ are fine tuned to have a special relationship to each other. We see then, that the contributions to the leptonic mixings coming from $`U_\nu `$ are either zero or small. The leptonic mixings thus arise from $`U_L`$. This is good news, since $`L^0`$ in Eq. (1) is known in this model. With the values of the parameters $`ϵ`$, $`\sigma `$, $`\delta `$, and $`\delta ^{}`$ given earlier as determined by fitting known quantities, one finds that $$U_L^{}=\left(\begin{array}{ccc}c_{12}& s_{12}c_{23}& s_{12}s_{23}\\ s_{12}& c_{12}c_{23}& c_{12}s_{23}\\ 0& s_{23}& c_{23}\end{array}\right),$$ (5) where $$\begin{array}{c}\mathrm{tan}\theta _{23}^L=s_{23}/c_{23}\sigma +ϵ1.9,\hfill \\ \\ \mathrm{sin}\theta _{12}^L=s_{12}\delta \sqrt{\sigma ^2+1}/ϵ\sigma 0.07.\hfill \end{array}$$ (6) Note that the mixing $`\theta _{23}^L`$ is large and the mixing $`\theta _{12}^L`$ is small. As can be seen from Eq. (1), the smallness of the angle $`\theta _{12}^L`$ is related to the smallness of the corresponding angle for the quarks. The largeness of $`\theta _{23}^L`$, on the other hand, is due to the large “lopsided” entry $`\sigma `$. There is no similar entry for the mixing of the first family. One could imagine introducing one, but one would find that doing so would make it hard to fit many known quantities such as $`V_{us}`$, $`V_{ub}`$, $`m_e/m_\mu `$, and $`m_d/m_s`$. It would seem, then, that this model must give large $`\nu _\mu \nu _\tau `$ mixing and small $`\nu _e\nu _\mu `$ mixing. However, as we shall now show, the same model actually can give bimaximal mixing. We will suppose now that the up quark matrix element $`U_{11}^0`$ is not exactly zero, but is given by $`\eta U_{33}^0`$. If $`m_u(1\mathrm{GeV})4\mathrm{MeV}`$, then $`\eta 6\times 10^6`$, which is a thousand times smaller than the smallest of the other model parameters, $`\delta `$ and $`\delta ^{}`$. It is in this sense that the vanishing of the first row and column of $`U^0`$ is an excellent approximation. In $`SO(10)`$ the simplest possibility is that $`N_{11}^0`$ is also given by $`\eta M_U`$. Thus, we assume for $`N^0`$ the form $$N^0=\left(\begin{array}{ccc}\eta & 0& 0\\ 0& 0& ϵ\\ 0& ϵ& 1\end{array}\right)M_U.$$ (7) This tiny modification of Eq. (1) allows interesting effects for certain forms of $`M_R`$. What we find is that if the off-diagonal elements in the first row and first column of $`M_R`$ are small or zero, then the small-angle MSW solution results as in , whereas if these elements are important, the large-angle solution of either the “just-so” vacuum or large-angle MSW oscillation type can result. Instead of looking at the most general forms for $`M_R`$, we will illustrate this in two representative cases: $$\begin{array}{cccc}(I)\hfill & M_R^I=\left(\begin{array}{ccc}B& 0& 0\\ 0& 0& A\\ 0& A& 1\end{array}\right)\mathrm{\Lambda }_R,\hfill & (II)\hfill & M_R^{II}=\left(\begin{array}{ccc}0& A& 0\\ A& 0& 0\\ 0& 0& 1\end{array}\right)\mathrm{\Lambda }_R.\hfill \end{array}$$ (8) Small-angle MSW Solution Form (I): $$M_\nu ^I=N^{0T}(M_R^I)^1N^0=\left(\begin{array}{ccc}\eta ^2/B& 0& 0\\ 0& 0& ϵ^2/A\\ 0& ϵ^2/A& 2ϵ/Aϵ^2/A^2\end{array}\right)\frac{M_U^2}{\mathrm{\Lambda }_R}.$$ (9) This light neutrino mass matrix is diagonalized by an orthogonal transformation of the type given in Eq. (3) for which Eq. (4) applies with $`\theta _{23}^\nu ϵ`$. The presence of the new $`\eta `$ contribution to the 11 element of $`N^0`$ thus does not modify the small-angle MSW solution obtained earlier, if $`M_R`$ is of type (I). The ratio of the mass differences required for this solution, $`\delta m_{23}^23.5\times 10^3\mathrm{eV}^2`$, $`\delta m_{12}^27\times 10^6\mathrm{eV}^2`$, can easily be obtained with $`Aϵ`$, while the absolute mass scale follows with $`M_U100`$ GeV and $`\mathrm{\Lambda }_R1.5\times 10^{14}`$ GeV. Moreover, the mixing results, $`\mathrm{sin}^22\theta _{\mu \tau }1.0`$ and $`\mathrm{sin}^22\theta _{e\mu }0.008`$, lie in the desired ranges for the atmospheric and small-angle MSW solutions . Vacuum Oscillation Solution Form (II): $$M_\nu ^{II}=N^{0T}(M_R^{II})^1N^0=\left(\begin{array}{ccc}0& 0& (\eta /A)ϵ\\ 0& ϵ^2& ϵ\\ (\eta /A)ϵ& ϵ& 1\end{array}\right)\frac{M_U^2}{\mathrm{\Lambda }_R}.$$ (10) A rotation in the 2-3 plane by an angle $`\mathrm{tan}\theta _{23}^\nu =ϵ`$ will eliminate the 22, 23, and 32 entries and induce 12 and 21 entries that are equal to $`(\eta /A)ϵ^2`$ (neglecting terms higher order in $`ϵ^2`$). Following this with a rotation in the 1-3 plane by an angle $`\theta _{13}^\nu (\eta /A)ϵ`$ brings the matrix to the form $$M_\nu ^{II}\left(\begin{array}{ccc}(\eta /A)^2ϵ^2& (\eta /A)ϵ^2& 0\\ (\eta /A)ϵ^2& 0& 0\\ 0& 0& 1\end{array}\right)\frac{M_U^2}{\mathrm{\Lambda }_R}.$$ (11) For $`\left|\eta /A\right|1`$, a pseudo-Dirac form for the mass matrix of $`\nu _e`$ and $`\nu _\mu `$ obtains with nearly-degenerate neutrinos. One finds $$\begin{array}{cc}\hfill m_{\nu _e}m_{\nu _\mu }& \left|\eta /A\right|ϵ^2(M_U^2/\mathrm{\Lambda }_R),\hfill \\ \hfill \delta m_{12}^2& 2\left|\eta /A\right|^3ϵ^4(M_U^2/\mathrm{\Lambda }_R)^2,\hfill \\ \hfill \delta m_{23}^2& m_{\nu _\tau }^2(M_U^2/\mathrm{\Lambda }_R)^2\hfill \end{array}$$ (12) If one takes $`\delta m_{12}^24\times 10^{10}eV^2`$, corresponding to the vacuum oscillation solution of the solar neutrino problem , and $`\delta m_{23}^23.5\times 10^3eV^2`$ , then $`\left|\eta /A\right|0.05`$ and the pseudo-Dirac condition is satisfied. This means that $`\left|\theta _{13}^\nu \right|7\times 10^3`$, which we shall ignore, and $`\left|\theta _{12}^\nu \pi /4\right|0.05`$. Thus, to a good approximation we can write $`\theta _{23}^\nu =ϵ`$, $`\theta _{13}^\nu =0`$, and $`\theta _{12}^\nu =\pi /4`$, or $$U_\nu =\left(\begin{array}{ccc}1/\sqrt{2}& 1/\sqrt{2}& 0\\ 1/\sqrt{2}& 1/\sqrt{2}& ϵ\\ ϵ/\sqrt{2}& ϵ/\sqrt{2}& 1\end{array}\right).$$ (13) With the use of Eqs. (2) and (5), this gives for the lepton mixing matrix $$V_{lepton}\left(\begin{array}{ccc}1/\sqrt{2}& 1/\sqrt{2}& s_{12}s_{23}^{}\\ c_{23}^{}/\sqrt{2}& c_{23}^{}/\sqrt{2}& s_{23}^{}\\ s_{23}^{}/\sqrt{2}& s_{23}^{}/\sqrt{2}& c_{23}^{}\end{array}\right).$$ (14) with $`s_{23}^{}\mathrm{sin}\theta _{\mu \tau }`$, where $`\theta _{\mu \tau }\theta _{23}^Lϵ63^{}8^{}=55^{}`$; hence $`\mathrm{sin}^22\theta _{\mu \tau }0.9`$ while $`\mathrm{sin}^22\theta _{e\mu }1.0`$, which are consistent with the experimental limits. It is interesting that the very small value of $`\delta m_{12}^2`$ needed for the vacuum oscillation solution to the solar neutrino problem has a natural explanation in this approach. From Eq. (12) it is apparent that $`m_{\nu _\mu }m_{\nu _e}\eta `$. In other words, the very tiny parameter $`\eta `$ required to fit $`m_u/m_t`$ — a quantity pertaining to the first family — actually ends up controlling the masses of both the muon- and electron-neutrino which are nearly degenerate. Large-angle MSW Solution The question arises whether form (II) can also give the large-angle MSW solution to the solar neutrino problem . The answer is yes, though in this case $`\mathrm{sin}^22\theta _{e\mu }`$ departs significantly from maximality. For this solution one needs to have $`\mathrm{\Delta }m_{12}^22\times 10^5\mathrm{eV}^2`$ along with $`\mathrm{\Delta }m_{23}^23.5\times 10^3\mathrm{eV}^2`$. Although Eq. (11) still applies, the condition for pseudo-Dirac neutrinos is no longer satisfied as one finds $`\eta /A1.8`$. Three Majorana neutrinos emerge for which $$\begin{array}{cc}\hfill m_3& M_U^2/\mathrm{\Lambda }_R,\hfill \\ \hfill m_2& \frac{1}{2}(\eta /A)ϵ^2\left[\eta /A+\sqrt{4+(\eta /A)^2}\right]M_U^2/\mathrm{\Lambda }_R,\hfill \\ \hfill m_1& \frac{1}{2}(\eta /A)ϵ^2\left[\eta /A+\sqrt{4+(\eta /A)^2}\right]M_U^2/\mathrm{\Lambda }_R\hfill \end{array}$$ (15) With $`M_U100`$ GeV and $`\mathrm{\Lambda }_R1.7\times 10^{14}`$ GeV, the three neutrino masses are given numerically by $`5.9\times 10^2\mathrm{eV},4.7\times 10^3`$ eV and $`9.3\times 10^4`$ eV. Making use of the above value for $`\eta /A`$, we can write down the equivalent of Eq. (13) for the neutrino mixing matrix $$U_\nu =\left(\begin{array}{ccc}0.389& 0.878& 1.73ϵ\\ 0.922& 0.392& 0.961ϵ\\ 0.18ϵ& 2ϵ& 0.961\end{array}\right).$$ (16) The lepton mixing matrix is then found numerically to be $$V_{lepton}\left(\begin{array}{ccc}0.360& 0.908& 0.186\\ 0.469& 0.367& 0.812\\ 0.811& 0.222& 0.556\end{array}\right).$$ (17) from which one obtains $`\mathrm{sin}^22\theta _{\mu \tau }=0.82`$ and $`\mathrm{sin}^22\theta _{e\mu }=0.43.`$ These values for the mixing angles are on the low side of the allowed experimental region for this large-angle MSW type of solution. Moreover, the $`A3\times 10^6`$ parameter required is some thirty times smaller than that found earlier with a pair of pseudo-Dirac neutrinos; thus a considerably larger hierarchy is required in the right-handed Majorana neutrino matrix to reproduce the large-angle MSW mixing than for the vacuum solution. These features have also appeared in other forms we have assumed for $`M_R`$. In summary, we have shown that, through the introduction of a small correction which gives mass to the up quark and hence also modifies the related Dirac neutrino matrix, a bimaximal solution to the solar and atmospheric neutrino oscillations can be achieved, provided the right-handed Majorana neutrino matrix mixes the first family with the other two. The large-angle vacuum solution is somewhat preferred over the large-angle MSW solution for the solar neutrino problem, since a smaller hierarchy is required in the Majorana matrix and the mixing angles are more nearly maximal as suggested by present experimental data for those type of solutions. But as presently understood, the model does not suggest a preference for the large-angle solutions over the small-angle MSW solution. One of us (CHA) thanks the Fermilab Theoretical Physics Department for its kind hospitality where much of his work was carried out. The research of SMB was supported in part by the Department of Energy under contract No. DE-FG02-91ER-40626. Fermilab is operated by Universities Research Association Inc. under contract No. DE-AC02-76CH03000 with the Department of Energy.
no-problem/9906/cond-mat9906044.html
ar5iv
text
# Comparison of Power Dependence of Microwave Surface Resistance of Unpatterned and Patterned YBCO Thin Film ## I Introduction The nonlinearity (power dependence) of the microwave (rf) surface impedance $`Z_S`$ of the high-$`T_c`$ superconductors is important for both practical applications and fundamental materials understanding. Nonlinearity in the surface impedance not only degrades the power handling ability of devices but also causes intermodulation and harmonic generation problems at moderate power levels. There has been a large amount of effort to investigate the nonlinear $`Z_S`$, but the origins of these phenomena are still not understood. Since most of the measurements have been performed on photolithographically patterned stripline or coplanar waveguide devices, one of the suggested explanations for the nonlinearity is defects and damage caused by patterning. All patterned devices have some current crowding near their edges. For example, stripline and microstrip devices have current distributions that peak sharply near their edges, where the patterning processes have a relatively large influence. A SEM picture focusing on the edge of a patterned YBCO thin film is shown in Fig. 1. The edge is not straight as can be seen in the picture and the roughness of the edge is of the order of a superconducting penetration depth. This edge roughness results from the initial imperfections of the photoresist pattern and from nonuniformities of the etching process. The edge roughness led to the speculation that the condition of the YBCO line edge was influencing the power dependence, and this was the initial motivation for this work. The bright particles on the surface of the YBCO film are the well-known copper oxide outgrowths that are present in all films whose stoichiometry is not exact. We have found that outgrowths do not adversely affect the microwave properties and therefore it is unnecessary to make large efforts to deposit films with exactly the correct stoichiometry. We have also observed that improvement of the surface morphology to make smoother films than that shown in Fig. 1 does not improve their microwave properties either with regard to their residual $`R_S`$ or their power dependence. In the present work, we used both a sapphire dielectric resonator and a stripline resonator to directly compare the power dependence of the microwave surface resistance $`R_S`$ of the same YBCO thin films, first unpatterned in the dielectric resonator and then patterned in the stripline resonator. Experimental procedures and results are shown in the following sections. To help understand our experimental results better, we have also modeled the $`R_S`$ as measured in the dielectric resonator and in the stripline resonator using the following simple phenomenological assumption, $$\rho (\stackrel{}{r})=\rho _0+\rho _2J_{rf}(\stackrel{}{r})^2,$$ (1) $$\lambda (\stackrel{}{r})=\lambda _0+\lambda _2J_{rf}(\stackrel{}{r})^2$$ (2) where $`\rho (\stackrel{}{r})`$ and $`\lambda (\stackrel{}{r})`$ are, respectively, the real part of the complex local resistivity and penetration depth, $`J_{rf}(\stackrel{}{r})`$ is the local current density in the film, and $`\rho _2`$ and $`\lambda _2`$ are the corresponding nonlinear coefficients. ## II Experimental Method In order to study the effects of defects and damage introduced in the patterning process on the power dependence of the microwave surface resistance of YBCO thin films, we measured the nonlinear microwave frequency surface resistance at 10.7 GHz of unpatterned, 2-inch-diameter, YBCO films using a specially designed sapphire dielectric resonator. The unpatterned films were then patterned using standard photolithography and wet chemical etching with 0.1% phosphoric acid to make stripline resonators with resonant frequencies $`f_0`$ = $`n`$$`f_1`$, where $`n=1,2`$, … and $`f_1`$ = 1.5 GHz. The patterning and dicing procedures are described in Fig. 2. The microwave surface resistance was then measured as a function of rf power for the patterned striplines. Comparing our results “before” and “after” the patterning process, directly clarifies whether the patterning contributes to the power dependence of the microwave surface resistance observed for YBCO thin films. ### A Sapphire Dielectric Resonator We have designed a sapphire dielectric resonator shown in Fig. 3 to measure the microwave surface resistance and especially the power dependence of unpatterned YBCO thin films. The resonator incorporates some unique features described below that allow high-power measurements to be carried out without heating the sample via dissipated power. Figure 4 is a cut-away view of the dielectric resonator. The resonator was made from a $`\frac{1}{2}`$-inch-diameter and $`\frac{1}{4}`$-inch-high cylindrical sapphire puck that is grounded by two 2-inch-diameter YBCO wafers (see left hand side of Fig. 2). The resonator has fixed output coupling and in-situ-adjustable input coupling to ensure critical input coupling (no reflection at the input end) for high-power measurements. The resonator is packaged in an oxygen-free copper case, as shown in Fig. 3. Special thermal contacts were designed to avoid heating problems for high-power experiments. A copper pressure plate was employed on top of the upper YBCO wafer. Springs were used (as shown in Fig. 3) to ensure that the pressure on the YBCO is independent of temperature. A gold-plated copper foil was soldered onto the copper pressure plate to enhance thermal conduction to the copper case from the top YBCO film which was more likely to be affected by heating than the bottom wafer that was anchored to the copper base plate. Thermal-conducting grease was applied between the sapphire puck and the two grounding wafers. The whole package was operated in a vacuum environment in a cryocooler with the bottom plate bolted to the cold finger with an indium foil for thermal contact. A temperature sensor and a 25-$`\mathrm{\Omega }`$ heater were mounted on the copper case of the resonator with indium foil to monitor and control the temperature. The sapphire dielectric resonator was operated at the $`TE_{011}`$ mode for which the center frequency was 10.7 GHz. The loaded quality factor $`Q_L`$, reflection ($`S_{11}`$, $`S_{22}`$) and transmission ($`S_{21}`$) coefficients were measured as a function of input power and temperature. The input coupling loop was tuned in-situ to maintain critical coupling at each temperature and input power level. Critical coupling on the input maximizes the circulating power and rf magnetic field in the resonator. The output coupling loop was fixed and weakly coupled during each measurement. The unloaded quality factor $`Q_0`$ was obtained from the measured $`S`$ parameters and $`Q_L`$, and the surface resistance was deduced from $`R_S`$ = $`G/Q_0`$, where $`G`$ is the calculated geometric factor of the resonator. The results for $`R_S`$ are presented as a function of peak magnetic field in the resonator. The $`R_S`$ measured by this method is the average of $`R_S`$ in the top and bottom YBCO films. The $`Q_L`$ was measured using a standard frequency-domain 3-dB bandwidth method at a low input power where the $`Q`$ is independent of power. At higher input power levels, a time-domain method was employed to reduce heating effects. In this method, a CW microwave pulse with enough duration to fully charge the resonator was sent into the resonator. After the input pulse is turned off, we can calculate $`Q_L`$ from the decay rate of the output signal that is proportional to $`exp[\omega t/Q_L]`$. By use of the time-domain method with this special thermal design, the $`R_S`$ could be measured without heating problems for input powers up to 40 dBm, which corresponds to a rf peak magnetic field of $``$100 Oe. Surface resistance measurements on two pairs of unpatterned YBCO films were carried out for temperatures in the range 30 to 80 K. In the analysis of the experimental data, we have, as is usual, treated the losses of the YBCO superconducting ground planes as a perturbation that does not affect the field or current distribution, while ignoring the losses due to the sapphire puck itself and to the copper case. The assumption that only the YBCO films contribute to the power loss is expected to be valid for the following reasons. A good-quality sapphire crystal is almost lossless at the cryogenic temperature where we operated, since the $`\mathrm{tan}\delta `$(loss tangent) is estimated to be on the order of $`10^9`$. For YBCO films, a $`Q_0`$ of about $`10^6`$ is observed at low temperatures. A pair of 2-inch-diameter Nb films have yielded $`Q_0`$ values greater than $`5\times 10^6`$. That the Nb films gave a higher $`Q`$ than those of YBCO indicates that the loss in the sapphire puck is negligible compared to that of the YBCO films. We estimated that the loss in the copper case is very small and yields a $`Q`$ greater than $`2.5\times 10^8`$, much higher than that of the YBCO films. ### B Stripline Resonator After the microwave surface resistance of the unpatterned YBCO films was measured with the sapphire dielectric resonator, the films were patterned and stripline resonators were made. A standard patterning process with wet chemical etching was employed. As shown in Fig. 2, each 2-inch wafer was made into four striplines and four ground planes. One piece from each of the top and the bottom YBCO wafers used with the dielectric resonator in Fig. 3 was measured. A description of the stripline resonator used here can be found in detail elsewhere. Similar to the dielectric resonator method, the loaded quality factor $`Q_L`$ was measured for the stripline as a function of microwave input power up to 35 dBm, corresponding to a rf peak magnetic field of 1000 Oe in the resonator. One advantage of the stripline resonator is that the $`R_S`$ at different overtone frequencies can be easily measured. By measuring overtones, we could measure the $`Q`$ at frequencies close to that of the dielectric resonator even though the fundamental resonance of the stripline resonator is 1.5 GHz. ## III Experimental Results The $`R_S`$ as a function of rf peak magnetic field $`H_{\mathrm{max}}`$ at different temperatures was measured for two pairs of unpatterned YBCO films, denoted as pair 1 and 2 (see Fig. 3). The results are shown in Fig. 5. For pair 1, no nonlinearity in $`R_S`$ of the unpatterned wafers was observed up to 80 K even at +40 dBm, the maximum available power, and corresponding to $`H_{rf}`$ values up to 200 Oe. As mentioned above, $`R_S`$ obtained for the unpatterned wafers is an average surface resistance of the pair of films. Figure 6(a) compares $`R_S`$ of the unpatterned and patterned films at 35 K for pair 1. The open circles represent the $`R_S`$ of the patterned film from the top wafer measured in the dielectric resonator, and the crosses denote the $`R_S`$ of the patterned film from the bottom wafer. The average of those two curves (dashed line) agrees with the $`R_S`$ of the unpatterned wafers very well and suggests that there is no nonlinearity for $`H_{\mathrm{max}}`$ up to 500 Oe. Figure 6(b) shows the results at 75 K for pair 1. For the case of stripline resonators, because of package modes interfering with some of the overtone modes, the $`R_S`$ could not be measured at the mode closest in frequency to that of the dielectric resonator. Therefore $`R_S`$ was measured at the closest frequencies possible. The results for $`R_S`$ were then scaled assuming $`R_S`$ $``$ $`\omega ^2`$. For pair 1 the $`R_S`$ of the unpatterned films was measured at 10.7 GHz, and the $`R_S`$ of the two patterned films was measured at 12 GHz and 7.5 GHz. From the results in Fig. 6, we conclude that for pair 1, no degradation in the residual surface resistance was observed in the patterned stripline films up to a $`H_{\mathrm{max}}`$ of $``$ 200 Oe. No power dependence was observed up to the maximum input power available for the unpatterned films, and no effects of patterning on the power handling of YBCO films could be found either at 35 K or at 75 K for pair 1. We grew another pair of 2-inch-diameter YBCO wafers pair 2, under a different deposition temperature with the intention to decrease the power handling of the microwave surface resistance. For pair 2 at 35 K no power dependence of the $`R_S`$ was observed in the dielectric resonator up to the maximum input power, similar to the results obtained for pair 1. However, at $`T=`$ 75 K, a nonlinearity in the $`R_S`$ was observed above $`H_{\mathrm{max}}`$ $``$ 100 Oe. Results for the patterned films are shown together with that of the unpatterned films in Fig. 7. In this case, $`R_S`$ was measured at 7.5 GHz for both patterned films in the stripline resonator. The measured residual surface resistance of the unpatterned and patterned films agrees to within 10$`\%`$. The $`R_S(H_{\mathrm{max}})`$ for the unpatterned films actually turns up at a somewhat lower $`H_{\mathrm{max}}`$ than for the patterned films in the stripline resonator. The apparently better power handling of the stripline resonator can be explained using the same surface impedance by a simple model introduced below. The $`H_{\mathrm{max}}`$ is the rf peak magnetic field in the two types of resonators, but the two types of resonators have very different field distributions. In the next section, we use the different current/field profiles for both the sapphire dielectric and stripline resonators to model the measured power dependence of the unpatterned and patterned films. ## IV Model The current profiles in the films for the dielectric and stripline resonators are different, as shown in Fig. 8. For the dielectric resonator with a film diameter much larger than the diameter of the sapphire puck so that the edge effects can be ignored, the current density $`\stackrel{}{J}`$ is in the $`\varphi `$ direction and has a dependence on radius given by $$J(r)=\frac{H(r)}{\lambda _0}=\{\begin{array}{cc}A\frac{\beta }{\xi _1\lambda _0}J_1(\xi _1r)\hfill & \text{for }0<r<a\hfill \\ A\frac{\beta }{\lambda _0\xi _2}\frac{J_0(\xi _1a)}{K_0(\xi _2a)}K_1(\xi _2r)\hfill & \text{for }a<r<R\hfill \end{array}$$ (3) where $`J_0`$, $`J_1`$, $`K_0`$, $`K_1`$ are various Bessel functions, $`\beta `$, $`\xi _1`$ and $`\xi _2`$ are constants determined by the geometry of the resonator, a is the radius of the sapphire puck and $`R`$ is the radius of the superconducting films. The current density thus peaks near the center of the film as shown in Fig. 8. For the stripline resonator, the current density $`\stackrel{}{J}`$ has been calculated numerically in Ref. For the purpose of simplicity in the calculation, we approximated $`J(x)`$ with the following analytical form, $$J(x)=\{\begin{array}{cc}J_s(0)[1(\frac{2x}{w})^2]^{\frac{1}{2}}\hfill & xw\hfill \\ J_s(0)(\frac{1.165}{\lambda _0})(wb)^{\frac{1}{2}}\mathrm{exp}(\frac{(w/2|x|)b}{\lambda _{0}^{}{}_{}{}^{2}})\hfill & xw\hfill \end{array}$$ (4) where, $`w`$ and $`b`$ are, respectively, the width and thickness of the film. This analytical approximation agrees with the numerically calculated current distribution to within 5$`\%`$ in our case. Thus $`J(x)`$ peaks sharply at the edges of the patterned film for the stripline resonator. For the dielectric resonator, as is obvious in Eq. (3), $`J(x)`$ is related directly to the rf peak magnetic field $`H_{\mathrm{max}}`$. For the stripline resonator, $`H_{\mathrm{max}}`$ can be calculated numerically from the current distribution in the stripline. For the same maximum field $`H_{\mathrm{max}}`$ in the dielectric and stripline resonators, the portion of the film carrying a high current density is larger for the dielectric resonator due to the broader peak in the $`J(r)`$ distribution of the dielectric resonator. Therefore an apparently better power handling in the stripline resonator is not surprising. The quality factors $`Q_0`$($`H_{\mathrm{max}}`$) and in turn $`R_S`$($`H_{\mathrm{max}}`$) of both resonators have been modeled as a function of rf peak magnetic field using the current distributions given above, with the assumption for the impedance is given by Eqs. (1) and (2), and $$Q_0=\omega \frac{W_{\mathrm{stored}}}{P_{\mathrm{diss}}},$$ (5) where $`W_{\mathrm{stored}}`$ is the total energy stored and $`P_{\mathrm{diss}}`$ is the power dissipated in the resonator. The parameters $`\rho _0`$ and $`\lambda _0`$ were taken from experimentally measured values, $`\rho _0=7.2\times 10^{11}`$$`\mathrm{\Omega }`$m, $`\lambda _0`$ = 0.2 $`\mu `$m, that fit the low-field, linear part of $`R_S`$ and the temperature dependence of the resonant frequency. The nonlinear parameters of resistivity and penetration depth used in the calculations in the superconducting phase were $`\rho _2=1.0\times 10^{35}`$$`\mathrm{\Omega }`$$`\mathrm{m}^5/\mathrm{A}^2`$ and $`\lambda _2=2.5\times 10^{29}`$$`\mathrm{m}^5/\mathrm{A}^2`$. These parameters were taken to fit the $`R_S`$ obtained from the dielectric resonator measurements. The unloaded quality factor for the stripline resonator was then calculated with the same set of parameters. At low rf field, calculated values of $`Q_0`$ of $`7\times 10^5`$ and 6000, respectively, were obtained for the dielectric and stripline resonators, and these values are consistent with the measured values. The modeled results also show that the power dependence of $`R_S`$ for both the dielectric and stripline resonators is quantitatively consistent with the measured power dependence for both resonators. The modeled $`R_S`$ of the dielectric resonator turns up at a smaller value of $`H_{\mathrm{max}}`$ than that of the stripline resonator, which agrees with the experimental result well as shown in Fig. 9. ## V Conclusions and Discussion We have measured the microwave surface resistance of the same YBCO thin films, before and after patterning, using a sapphire dielectric resonator and a stripline resonator. A phenomenological model of the $`R_S`$$`(H_{\mathrm{max}})`$ in the two cases is presented. Based on this model, the calculated results fit the experimental data well in both cases using the same materials parameters. Therefore, we conclude that our results are consistent with no damage or degradation of the power dependence of the microwave surface resistance due to the patterning of the films. The procedure described in this paper can also be used to test other patterning processes such as ion beam etching, etc. ###### Acknowledgements. This work was supported by the Air Force Office of Scientific Research. The authors express their gratitude to Bob Koneizcka and Earle Macedo for applying their tireless efforts and effective skills to the preparation of the devices used in this study, and to Peter Murphy for providing the SEM pictures used in this study.
no-problem/9906/physics9906037.html
ar5iv
text
# Accurate ab initio anharmonic force field and heat of formation for silane ## I Introduction The spectroscopy and thermochemistry of the silane (SiH<sub>4</sub>) molecule have aroused interest from a number of perspectives. Its importance as a precursor for the chemical vapor deposition (CVD) of silicon layers has been discussed at length by Allen and Schaefer, who also review early theoretical work on the molecule. The spectroscopy of the tetrahedral group IV hydrides AH<sub>4</sub> (A=C, Si, Ge, Sn, Pb) has been extensively studied. For a review of early work on AH<sub>4</sub> (A=Si, Ge, Sn) the reader is referred to Ref.. A complete bibliography on experimental work on methane and its isotopomers would be beyond the scope of this work (see Refs. for detailed references): we do note that an accurate ab initio force field was computed by a team involving two of us. Based on this force field, a number of theoretical spectroscopic studies of the excited vibrational states of CH<sub>4</sub> were recently studied: we note in particular a full-dimenstional variational study by Carter et al., a low-order perturbation theoretical/resonance polyad study by Venuti et al., and a high-order canonical Van Vleck perturbation theory study by Wang and Sibert. We also note an accurate anharmonic force field on the isoelectronic NH$`{}_{}{}^{+}{}_{4}{}^{}`$ molecule by two of us. The infrared spectrum of silane, SiH<sub>4</sub>, was first studied in 1935 by Steward and Nielsen and a set of fundamental frequencies for the most abundant isotopomer was first obtained in 1942 by Nielsen and coworkers. The isotopomers of SiH<sub>4</sub> have been the subject of considerable high-resolution experimental work; for instance, we note for <sup>28</sup>SiH<sub>4</sub>, <sup>29</sup>SiH<sub>4</sub>, <sup>30</sup>SiH<sub>4</sub>, for <sup>28</sup>SiH<sub>3</sub>D, for <sup>28</sup>SiHD<sub>3</sub>, and for <sup>28</sup>SiD<sub>4</sub>. The molecule is of considerable astrophysical interest, having been detected spectroscopically in the atmospheres of Jupiter and Saturn and in the interstellar gas cloud surrounding the carbon star IRC+10 216 Until most recently, only fairly low-resolution data were available for SiH<sub>2</sub>D<sub>2</sub>; as the present paper was being prepared for publication, a high-resolution study of the $`\{\nu _3,\nu _4,\nu _5,\nu _7,\nu _9\}`$ Coriolis resonance polyad appeared, in which assignments were facilitated by mixed basis set CCSD(T) and MP2 calculations of the quartic force field. One of the interesting features of the infrared spectra of silane is their pronounced local-mode character (e.g. ), leading to complex resonance polyads. The strongly ‘local’ character also inspired a study of the SiH<sub>4</sub> spectrum up to seven quanta using algebraic methods. In the present work, we shall report a high-quality quartic force field that is of constant quality for all the isotopomers of silane. A theoretical spectroscopy study by Wang and Sibert is currently in progress on excited states and vibrational resonance polyads of SiH<sub>4</sub> and isotopomers, using high-order (6th and 8th) canonical Van Vleck perturbation theory and the force field reported in the present work. Since this can be done at very little additional computational expense, we shall also report a benchmark atomization energy and heat of formation of SiH<sub>4</sub>. The thermodynamic properties of silane are linked to a controversy concerning the heat of vaporization of silicon, which is of fundamental importance to computational chemists since it is required every time one attempts to directly compute the heat of formation of any silicon compound, be it ab initio or semiempirically. $`\mathrm{\Delta }H_{f,0}^{}`$\[Si(g)\] is given in the JANAF tables as 106.6$`\pm `$1.9 kcal/mol. Desai reviewed the available data and recommended the JANAF value, but with a reduced uncertainty of $`\pm `$1.0 kcal/mol. Recently, Grev and Schaefer (GS) found that their ab initio calculation of the TAE of SiH<sub>4</sub>, despite basis set incompleteness, was actually larger than the value derived from the experimental heats of formation of Si($`g`$), H($`g`$), and SiH<sub>4</sub>($`g`$). They concluded that the heat of vaporization of silicon should be revised upwards to $`\mathrm{\Delta }H_{f,0}^{}`$\[Si($`g`$)\]=108.07(50) kcal/mol, a suggestion supported by Ochterski et al.. Very recently, however, Collins and Grev (CG) considered the scalar relativistic contribution to the binding energy of silane using relativistic coupled cluster techniques within the Douglas-Kroll (no-pair) approximation, and found a contribution of -0.67 kcal/mol. This would suggest a downward revision of the GS value of $`\mathrm{\Delta }H_{f,0}^{}`$\[Si(g)\] to 107.4 kcal/mol, which is in excellent agreement with a recent redetermination by Martin and Taylor of 107.15$`\pm `$0.39 kcal/mol. (This latter value was derived by combining a benchmark ab initio calculation of the total atomization energy of tetrafluorosilane, TAE<sub>0</sub>\[SiF<sub>4</sub>\], with a very precise fluorine bomb calorimetric measurement of $`\mathrm{\Delta }H_f^{}`$\[SiF<sub>4</sub>(g)\].) In addition, it was pointed out that the JANAF value of $`\mathrm{\Delta }H_{f,0}^{}`$\[SiH<sub>4</sub>($`g`$)\]=10.5$`\pm `$0.5 kcal/mol is in fact the Gunn and Green value of 9.5$`\pm `$0.5 kcal/mol increased by a correction of +1 kcal/mol for the phase transition Si(amorphous)$``$Si(cr). (Gunn and Green considered this correction to be an artifact of the method of preparation and ignored it.) Clearly, a calibration calculation of TAE<sub>0</sub>\[SiH<sub>4</sub>\] might be desirable, and is the secondary purpose of the present study. Accurate thermochemical parameters of SiH<sub>4</sub> (and other silicon compounds) are of practical importance for the thermodynamic and kinetic modeling of such processes as laser-induced chemical vapor deposition of silicon films from silane, the chemical vapor deposition of tungsten contacts for ULSI (ultralarge scale integrated circuit) chips by SiH<sub>4</sub> reduction of WF<sub>6</sub> (e.g. ) and the generation of SiOxNy films by low-pressure chemical vapor deposition from mixtures of SiH<sub>4</sub> with N<sub>2</sub>O and/or NH<sub>3</sub> (e.g. as antireflective coatings and for ultrathin capacitors). (We also mention in passing the use of silane compounds in dentistry.) While GS’s work was definitely state of the art in its time, the attainable accuracy for this type of compound may well have gone up an order of magnitude in the seven years since it was published: in a recent systematic study of total atomization energies of a variety of first-and second-row molecules for which they are precisely known, procedures like the ones used in the present work achieved a mean absolute error of 0.23 kcal/mol, which dropped to 0.18 kcal/mol if only systems well described by a single reference determinant (as is the case with SiH<sub>4</sub>) were considered. In order to ascertain the utmost accuracy for hydrides, a zero-point energy including anharmonic corrections was found to be desirable: this is obtained as a by-product of the accurate anharmonic force field which is the primary subject of the present contribution. ## II Computational methods All electronic structure calculations were carried out using MOLPRO 97 running on DEC Alpha and SGI Origin computers at the Weizmann Institute of Science. The CCSD(T) \[coupled cluster with all single and double substitutions (CCSD) supplemented with a quasiperturbative estimate of the contribution of connected triple excitations\] method, as implemented in MOLPRO, was used throughout for the electronic structure calculations on SiH<sub>4</sub>. For the Si($`{}_{}{}^{3}P`$) atom, we employed the definition of Ref. for the open-shell CCSD(T) energy. The calculations including only valence correlation employed the standard Dunning cc-pV$`n`$Z (correlation consistent valence $`n`$-tuple zeta) basis sets on hydrogen and two different variants of the cc-pV$`n`$Z or aug-cc-pV$`n`$Z (augmented cc-pV$`n`$Z) basis sets on Si. The first variant, cc-pV$`n`$Z+1, was used in the force field calculations, and includes an additional high-exponent $`d`$ function to accommodate the greater part of the inner-shell polarization effect, which is known to be important for both energetic and geometric properties of second-row molecules. The second variant, aug-cc-pV$`n`$Z+2d1f, includes two high-exponent $`d`$ functions and a high-exponent $`f`$ function, with exponents determined by successively multiplying the highest exponent already present for that angular momentum by a factor of 2.5. Such a set should give an exhaustive account of the energetic effects of inner-shell polarization. Calculations including inner-shell correlation (not to be confused with inner-shell polarization, which is an SCF-level effect) were carried out using the Martin-Taylor core correlation basis set. Relativistic effects were determined with the same basis set and as ACPF (averaged coupled pair functional) expectation values of the first-order Darwin and mass-velocity operators. Optimizations were carried out by univariate polynomial interpolation. Force constants in symmetry coordinates were determined by recursive application of the central finite difference formula: the symmetry coordinates are defined in the same way as in previous studies on the isovalent CH<sub>4</sub> and NH$`{}_{}{}^{+}{}_{4}{}^{}`$ molecules. The vibrational analyses were performed using a modified version of the SPECTRO program running on an IBM RS6000 workstation at NASA Ames and the DEC Alpha at the Weizmann institute. The alignment conventions for the anharmonic constants of a spherical top follow the work of Hecht and general formulae for these constants were taken from the paper by Hodgkinson et al.. Similar to previous work on the spherical tops Be<sub>4</sub> and CH<sub>4</sub>, the accuracy of the various spectroscopic constants was verified by applying opposite mass perturbations of $`\pm `$0.00001 a.m.u. to two of the hydrogen atoms, then repeating the analysis in the asymmetric top formalism. Finally, the reported zero-point energies include the $`E_0`$ term (which is the polyatomic equivalent of the $`a_0`$ Dunham coefficient in diatomics). ## III Results and discussion ### A Vibrational frequencies and anharmonic force field An overview of the basis set convergence of the computed bond distance, harmonic frequencies, and vibrational anharmonic corrections is given in Table 1. The effect of adding inner-shell polarization functions to the cc-pVTZ basis set is modest but significant (0.006 Å) on the bond distance: the Si–H stretching frequencies, however, are affected by 20–25 cm<sup>-1</sup>. The bending frequencies are not seriously affected: somewhat surprising are the fairly strong effects on the vibrational anharmonicities (including, to a lesser extent, the bending frequencies). The overall behavior is in contrast to previous observations for SO<sub>2</sub> in which the inner-polarization effects on lower-order properties like geometry and harmonic frequencies are very noticeable but those on anharmonicities next to nonexistent, but is consistent with the very strong basis set sensitivity noted for the first three anharmonic corrections of the first-row diatomic hydrides by Martin. Likewise, a rather strong sensitivity with respect to basis set improvement from VDZ+1 over VTZ+1 to VQZ+1 is seen for the Si–H stretching frequencies and all the anharmonicities, even as the harmonic bending frequencies appear to be close to converged with the VTZ+1 basis set. It appears that in general, basis set sensitivity of anharmonicities of A–H stretches is much more pronounced than that of A–B stretches. The effect of inner-shell correlation, while nontrivial for the purpose of accurate calculations, is quite a bit more modest than that of inner-shell polarization (as measured by comparing the cc-pVTZ and cc-pVTZ+1 results), and in fact is not dissimilar to what one would expect for a first-row molecule (e.g. CH<sub>4</sub> ). We will now consider computed fundamentals for the various isotopomers of silane with our best force field, CCSD(T)/cc-pVQZ+1. All relevant data are collected in Table 2. For <sup>28</sup>SiH<sub>4</sub>, <sup>29</sup>SiH<sub>4</sub>, and <sup>30</sup>SiH<sub>4</sub>, agreement between the computed and observed fundamentals can only be described as excellent, with a mean absolute deviation of 2.5 cm<sup>-1</sup>. Agreement for the completely deuterated isotopomer <sup>28</sup>SiD<sub>4</sub> is even better, with a mean absolute deviation of 1.9 cm<sup>-1</sup>. For the <sup>28</sup>SiH<sub>3</sub>D isotopomer, agreement is likewise excellent, with a mean absolute deviation of 2.1 cm<sup>-1</sup>. It would appear that the force field is certainly of good enough quality to permit assignments for the less well known isotopomers. For <sup>28</sup>SiHD<sub>3</sub>, the only precisely known bands are the Si–H stretch, $`\nu _1`$=2187.2070(10) cm<sup>-1</sup> , and the $`\nu _5`$ degenerate bend, 850.680823(10) cm<sup>-1</sup> . Meal and Wilson , in their 1956 low-resolution study, assigned absorptions at 1573, 1598, and 683 cm<sup>-1</sup> to $`\nu _2`$, $`\nu _4`$, and $`\nu _6`$, respectively. Our calculations confirm this assignment and are on average within about 2 cm<sup>-1</sup> of all the above bands. $`\nu _3`$ was not observed by Meal and Wilson, and these authors speculated that it coincide with the 683 cm<sup>-1</sup> ($`\nu _6`$) peak. Our own calculations predict a splitting of about 5.7 cm<sup>-1</sup> between $`\nu _3`$ and $`\nu _6`$; B3LYP/VTZ+1 infrared intensity calculations suggest that both bands should be observable. Inspection of the relevant spectrum (Fig. 3 in Ref.) revealed that, at the resolution afforded by the equipment used, meaningful resolution between $`\nu _3`$ and $`\nu _6`$ becomes essentially impossible, especially given contamination (noted by Meal and Wilson) from a SiH<sub>2</sub>D<sub>2</sub> impurity with $`\nu _4`$=682.5 cm<sup>-1</sup>. Until most recently, the only available information for SiH<sub>2</sub>D<sub>2</sub> was the Meal and Wilson work. Our calculations, like those of Rötger et al., unambiguously suggest assignment of the 1601 and 1587 cm<sup>-1</sup> bands to $`\nu _2`$ and $`\nu _6`$, respectively, rather than the opposite assignment proposed by Meal and Wilson. We note that $`\nu _6`$ is in a very close Fermi resonance with $`\nu _5+\nu _9`$ (the unperturbed levels being only about 10 cm<sup>-1</sup> apart), despite the fairly small interaction constant $`k_{569}`$=-20.88 cm<sup>-1</sup>. Our calculations confirm the assignments for all other bands aside from $`\nu _1`$ and $`\nu _8`$, which are calculated to be within 1 cm<sup>-1</sup> of each other such that a meaningful decision on whether or not to exchange $`\nu _1`$ and $`\nu _8`$ is impossible. The Meal-Wilson empirical force field value of 844 cm<sup>-1</sup> for $`\nu _5`$ (which they were unable to observe) agrees well with our calculation as well as with the high-resolution value of 842.38121(9) cm<sup>-1</sup>. Of the very recent measurements by Rötger et al., all five bands in the Coriolis pentad ($`\nu _3`$, $`\nu _4`$, $`\nu _5`$, $`\nu _7`$, and $`\nu _9`$) are in excellent agreement with the present calculation (mean absolute deviation 1.1 cm<sup>-1</sup>). Among the sources of residual error in the quartic force field, neglect of inner-shell correlation and imperfections to CCSD(T) appear to be the potentially largest. As seen in Table 1, inclusion of core correlation increases harmonic frequencies by as much as 7 cm<sup>-1</sup> in this case. The effect of correlation beyond CCSD(T) was seen to work in the opposite direction for the first-row diatomic hydrides; in the present work, we have compared FCI/VDZ+1 and CCSD(T)/VDZ+1 harmonic frequencies for the SiH diatomic in the $`X{}_{}{}^{2}\mathrm{\Pi }`$ and $`a{}_{}{}^{4}\mathrm{\Sigma }_{}^{}`$ states, and found a reduction in $`\omega _e`$ of 4 and 10 cm<sup>-1</sup>, respectively. (The FCI–CCSD(T) difference for $`\omega _e`$ was found in Ref. to converge very rapidly with the basis set.) Since FCI frequency calculations in a reasonable-sized basis set for SiH<sub>4</sub> are simply not a realistic option, we have taken another track. We have assumed that the computed CCSD(T)/VQZ+1 force field is fundamentally sound, and that any residual error would mostly affect the equilibrium bond distance and the diagonal quadratic force constants. We have then taken our quartic force field in symmetry coordinates, substituted the computed CCSD(T)/MTcore bond distance (which agrees to four decimal places with the best experimental value), and have iteratively refined the four diagonal quadratic force constants such that the four experimental fundamentals of <sup>28</sup>SiH<sub>4</sub> are exactly reproduced by our calculation. The final adjusted force field is given in Table 3 and is available in machine-readable format from the corresponding author. As seen in Table 2, our computed fundamentals for the other isotopomers with the adjusted force field are in essentially perfect agreement with experiment where accurate values are available. Discrepancies arise for some modes of SiH<sub>2</sub>D<sub>2</sub>, SiHD<sub>3</sub>, and SiD<sub>4</sub> where only low-resolution data are available. Particularly the discrepancy for $`\nu _2`$ of SiD<sub>4</sub> is completely out of character: the experimental difficulties involved in its determination suggest that perhaps the experimental value may be in error. (A discrepancy of 1.2 cm<sup>-1</sup> for $`\nu _2`$ in SiH<sub>3</sub>D is halved upon accounting for a Fermi resonance $`2\nu _8\nu _2`$.) We hope that our computed force field will stimulate further spectroscopic work on SiH<sub>4</sub> and may serve as a basis for studies employing more sophisticated vibrational treatments, such as the variational techniques variational techniques very recently applied to methane or high-order canonical Van Vleck perturbation theory. As noted in the Introduction, a study of the latter type is already in progress. ### B Geometry At the CCSD(T)/MTcore level, we compute a bond distance of 1.4734 Å, which we know from experience should be very close to the true value. Ohno, Matsuura, Endo, and Hirota (OMEH1) estimate an experimental $`r_e`$ bond distance of 1.4741 Å without supplying an error bar; in a subsequent study (OMEH2), the same authors, using two different methods, obtain 1.4734(10) Å (“method I”) and 1.4707(6) Å (“method II”), respectively, where uncertainties in parentheses are three standard deviations. The deviation between the (diatomic approximation) “method II” value and our present calculation is more than an order of magnitude greater than usual for this level of ab initio theory, while the “method I” value agrees to four decimal places with our calculation. (Normally, because of neglect of correlation effects beyond CCSD(T) which have the tendency to lengthen bonds by 0.0002–0.0006 Å, we expect our computed bond distance to be slightly short, rather than too long.) The computed bond distance of Rötger et al., 1.4735 Å at the CCSD(T)\[all electron\] level in a mixed basis set which does not contain any core correlation functions, is likewise in excellent agreement with the OMEH2 “method I” value. ### C Atomization energy of SiH<sub>4</sub> Using a 3-point geometric extrapolation $`A+B.C^n`$ from the SCF/AV$`n`$Z+2d1f ($`n`$=T,Q,5) atomization energies, we find an SCF limit component of the total atomization energy of 259.83 kcal/mol, only marginally different from the directly computed SCF/AV5Z+2d1f value of 259.82 kcal/mol and only 0.05 kcal/mol larger than the GS result. The CCSD valence correlation component was extrapolated using the 2-point formula $`A+B/n^3`$ from AV$`n`$Z+2d1f ($`n`$=Q,5) results; thus we obtain a CCSD limit of 64.26 kcal/mol, which is 0.8 kcal/mol larger than the largest basis set value (63.45 kcal/mol) and 1.4 kcal/mol larger than the largest basis set value of GS (62.86 kcal/mol). Using the alternative 3-point extrapolation $`A+B/(l+1/2)^C`$ from AV$`n`$Z+2d1f ($`n`$=T,Q,5) we obtain a somewhat smaller basis set limit of 63.92 kcal/mol; however, as discussed in Ref., this procedure appears to systematically underestimate basis set limits and was found to yield excellent agreement with experiment largely due to an error compensation with neglect of scalar relativistic effects. At 0.81 kcal/mol, the extrapolated basis set limit contribution of connected triple excitations is quite modest, and differs by only 0.02 kcal/mol from the largest basis set value of 0.79 kcal/mol. In fact, it is largely immaterial whether the extrapolation is done from AV$`n`$Z+2d1f ($`n`$=T,Q) or from AV$`n`$Z+2d1f ($`n`$=Q,5), and we obtain essentially the same result for the (T) contribution as GS (0.82 kcal/mol). This is an illustration of the fact that connected triple excitations generally converge more rapidly with basis set than the CCSD correlation energy. Adding up the two basis set limit values, we find a valence correlation component to TAE of 65.05 kcal/mol; given the essentially purely single-reference character of the SiH<sub>4</sub> wave function there is little doubt that the CCSD(T) limit is very close to the full CI limit as well. As noted by GS, the contribution of inner-shell correlation of SiH<sub>4</sub> is negative: we find -0.365 kcal/mol compared to their -0.31 kcal/mol. The spin-orbit contribution is trivially obtained from the Si($`{}_{}{}^{3}P`$) atomic fine structure as -0.43 kcal/mol, while our computed scalar relativistic contribution, -0.70 kcal/mol, is essentially identical to the CG value. Finally, we obtain TAE<sub>e</sub>=323.39 kcal/mol. The anharmonic zero-point vibrational energy (ZPVE) from our best force field (including $`E_0`$) is 19.59 kcal/mol. This is very close to the value of 19.69 kcal/mol obtained by GS as an average of estimated fundamentals and CISD/TZ2P harmonic frequencies: the computational effort involved in improving this estimate by a mere 0.1 kcal/mol would therefore have been hard to justify if the anharmonic force field would not have been required for another purpose. Also, from past experience, we know that such good agreement between rigorous anharmonic ZPVEs and estimates cannot be taken for granted for hydrides. Our best TAE<sub>e</sub> and ZPVE finally lead to TAE<sub>0</sub>=303.80 kcal/mol, to which we attach an error bar of about 0.18 kcal/mol based on previous experience. This should be compared with the GS largest basis set result of 303.03 kcal/mol (or 302.36 kcal/mol after applying the CG scalar relativistic contributions) or the value derived from JANAF heats of formation of Si(g), H(g), and SiH<sub>4</sub>(g), 302.62 kcal/mol. If we consider alternative values for $`\mathrm{\Delta }H_{f,0}^{}`$\[Si(g)\] of 108.1$`\pm `$0.5 kcal/mol (GS), 107.4$`\pm `$0.5 kcal/mol (applying CG to the latter value), or 107.15$`\pm `$0.38 kcal/mol (Martin & Taylor), we would obtain from our calculation $`\mathrm{\Delta }H_{f,0}^{}`$\[SiH<sub>4</sub>($`g`$)\] values of 10.8$`\pm `$0.5, 10.1$`\pm `$0.5, and 9.9$`\pm `$0.4 kcal/mol, respectively. Only the first of these values cannot be reconciled with Gunn and Green; the very similar values derived from the Collins-Grev-Schaefer and Martin-Taylor $`\mathrm{\Delta }H_{f,0}^{}`$\[Si(g)\] agree to within accumulated error bars with both the JANAF and Gunn-Green values for the heat of formation of silane. While our best value of 9.9$`\pm `$0.4 kcal/mol at first sight slightly favors the Gunn-Green value (in which the Si($`cr`$)$``$Si($`amorph`$) transition enthalpy was considered an artifact of the manner of preparation), the difference is “too close to call”. We contend that our calculated value is more reliable than either experiment. ## IV Conclusions From accurate ab initio calculations and a minor empirical adjustment, a quartic force field for silane has been derived that is consistently of spectroscopic quality ($`\pm `$1 cm<sup>-1</sup> on vibrational fundamentals) for all isotopomers of silane studied here (<sup>28</sup>SiH<sub>4</sub>, <sup>29</sup>SiH<sub>4</sub>, <sup>30</sup>SiH<sub>4</sub>, <sup>28</sup>SiH<sub>3</sub>D, <sup>28</sup>SiH<sub>2</sub>D<sub>2</sub>, <sup>28</sup>SiHD<sub>3</sub>, and <sup>28</sup>SiD<sub>4</sub>). As in previous studies on second-row molecules, we found that inner-shell polarization functions have an appreciable effect on computed properties, and for hydrides this apparently includes the vibrational anharmonicities. From large basis set coupled cluster calculations and extrapolations to the infinite-basis set limit, we obtain TAE<sub>0</sub>=303.80$`\pm `$0.18 kcal/mol, which includes an anharmonic zero-point energy (19.59 kcal/mol), inner-shell correlation ($``$0.36 kcal/mol), scalar relativistic corrections ($``$0.70 kcal/mol), and atomic spin-orbit corrections ($``$0.43 kcal/mol). In combination with the recently revised $`\mathrm{\Delta }H_{f,0}^{}`$\[Si(g)\], 107.15$`\pm `$0.39 kcal/mol, we obtain $`\mathrm{\Delta }H_{f,0}^{}`$\[SiH<sub>4</sub>(g)\]=9.9$`\pm `$0.4 kcal/mol, intermediate between the JANAF and Gunn-Green values of 10.5$`\pm `$0.5 and 9.5$`\pm `$0.5 kcal/mol, respectively. ###### Acknowledgements. JM is a Yigal Allon Fellow, the incumbent of the Helen and Milton A. Kimmelman Career Development Chair, and an Honorary Research Associate (“Onderzoeksleider in eremandaat”) of the National Science Foundation of Belgium (NFWO/FNRS). KKB was a Fulbright Visiting Scholar at the Weizmann Institute of Science (on leave of absence from SDSC) during the course of this work. This research was partially supported by the Minerva Foundation, Munich, Germany. We thanks Drs. X.-G. Wang and E. L. Sibert III (U. of Wisconsin, Madison) for their encouragement.
no-problem/9906/cond-mat9906404.html
ar5iv
text
# Deviation from Maxwell Distribution in Granular Gases with Constant Restitution Coefficient ## Abstract We analyze the velocity distribution function of force-free granular gases in the regime of homogeneous cooling when deviations from the Maxwellian distribution may be accounted only by leading term in the Sonine polynomial expansion. These are quantified by the magnitude of the coefficient $`a_2`$ of the second term of the expansion. In our study we go beyond the linear approximation for $`a_2`$ and observe that there are three different values (three roots) for this coefficient which correspond to a scaling solution to the Boltzmann equation. The stability analysis performed showed, however, that among these three roots only one corresponds to a stable scaling solution. This is very close to $`a_2`$, obtained in previous studies in a linear with respect to $`a_2`$ approximation. PACS numbers: 81.05.Rm, 36.40.Sx, 51.20.+d, 66.30.Hs The granular gases, i.e., rarefied systems composed of inelastically colliding particles have been of particular interest during the last decade (e.g. ). Compared to gases of elastically colliding particles, the dissipation of energy at inelastic collisions leads to some novel phenomena in these systems. One can mention clustering (e.g. ), formation of vortex patterns (e.g.), etc. Before clustering starts, the granular gas being initially homogeneous, keeps for some time its homogeneity, although its temperature permanently decreases. This regime is called the homogeneous cooling regime (HC). In the present study we address the properties of the velocity distribution of granular particles in the regime of HC, such as the deviation from the Maxwellian distribution and the stability of the distribution function. We assume that the restitution coefficient $`ϵ`$ does not depend on the impact velocity, i.e. that $`ϵ=\mathrm{const}`$. The properties of the velocity distribution for the system with the impact-velocity dependent coefficient of restitution (e.g. ) will be addressed elsewhere . It is well known that granular gases in the HC regime do not reveal Maxwellian distribution (e.g.). The high-velocity tail is overpopulated , while the main part of the distribution is described by the sum of the Maxwellian one and the correction to it, written in terms of the Sonine polynomial expansion (e.g. ). Usually only the leading, second term, in this expansion is taken into account , moreover in previous studies only linear analysis with respect to the coefficient $`a_2`$, which refers to this second term has been performed. Finding that $`a_2`$, obtained within the linear approximation, is small, the authors of Ref. conclude a posteriori that the the linear approximation is valid. In our approach we also assume that one can restrict oneself to the leading term in the Sonine polynomial expansion and ignore the other. However we go beyond the linear approximation with respect to the coefficient $`a_2`$ and perform complete analysis within this level of the system description. We observed that there are three different values of $`a_2`$ exist which correspond to the scaling solution of the Boltzmann equation. The stability analysis for the velocity distribution function shows, however, that only one value of $`a_2`$ corresponds to a physically acceptable stable scaling solution. The stable solution is close to the result previously obtained within the linear analysis . To introduce notations and specify the problem we briefly sketch the derivation of the coefficient $`a_2`$ , in accordance with the approach developed in Ref.. So far we introduce the (time-dependent) temperature $`T(t)`$, and thermal velocity $`v_0(t)`$, which are related to the velocity distribution function $`f(𝐯,t)`$ as $$\frac{3}{2}nT(t)=𝑑𝐯\frac{v^2}{2}f(𝐯,t)=\frac{3}{2}nv_0^2(t)$$ (1) here $`n`$ is the number density of the granular gas, the particles are assumed to be of a unit mass ($`m=1`$), and (1) is written for 3D-systems. The inelasticity of collisions is characterized by the coefficient of the normal restitution $`ϵ`$ (here we consider smooth particles), which relates after-collisional velocities $`𝐯_1^{}`$, $`𝐯_2^{}`$ to the pre-collisional ones, $`𝐯_1`$, $`𝐯_2`$ as: $$𝐯_{1/2}^{}=𝐯_{1/2}\frac{1}{2}(1+ϵ)(𝐯_{12}𝐞)𝐞$$ (2) where $`𝐯_{12}=𝐯_1𝐯_2`$ is the relative velocity, the unit vector $`𝐞=𝐫_{12}/|𝐫_{12}|`$ gives the direction of the vector $`𝐫_{12}=𝐫_1𝐫_2`$ at the instant of the collision. The time-evolution of the velocity distribution function is subjected to the Enskog-Boltzmann equation, which for the force-free case reads : $`{\displaystyle \frac{}{t}}f(𝐯,t)=g_2(\sigma )\sigma ^2{\displaystyle 𝑑𝐯_2𝑑𝐞\mathrm{\Theta }(𝐯_{12}𝐞)|𝐯_{12}𝐞|}`$ (3) $`\times \left\{{\displaystyle \frac{1}{ϵ^2}}f(𝐯_1^{},t)f(𝐯_2^{},t)f(𝐯_1,t)f(𝐯_2,t)\right\}`$ (4) where $`\sigma `$ is the diameter of particles, $`g_2(\sigma )=(2\eta )/2(1\eta )^3`$ ($`\eta =\frac{1}{6}\pi n\sigma ^3`$ is packing fraction) denotes the contact value of the two-particle correlation function , which accounts for the increasing collision frequency due to the excluded volume effects; $`\mathrm{\Theta }(x)`$ is the Heaviside function. The velocities $`𝐯_1^{}`$ and $`𝐯_2^{}`$ refer to the precollisional velocities of the so-called inverse collision, which results with $`𝐯_1`$ and $`𝐯_2`$ as the after-collisional velocities. The factor $`1/ϵ^2`$ in the gain term appears respectively from the Jacobian of the transformation $`d𝐯_1^{}d𝐯_2^{}d𝐯_1d𝐯_2`$ and from the relation between the lengths of the collisional cylinders $`ϵ|𝐯_{12}^{}𝐞|dt=|𝐯_{12}𝐞|dt`$ . Assuming that the velocity distribution function is of a scaling form: $$f(𝐯,t)=\frac{n}{v_0^3(t)}\stackrel{~}{f}(𝐜)$$ (5) one can show, that the scaling function satisfies the time-independent equation : $$\frac{\mu _2}{3}\left(3+c_1\frac{}{c_1}\right)\stackrel{~}{f}(𝐜)=\stackrel{~}{I}(\stackrel{~}{f},\stackrel{~}{f})$$ (6) with the dimensionless collision integral: $`\stackrel{~}{I}(\stackrel{~}{f},\stackrel{~}{f})={\displaystyle 𝑑𝐜_2𝑑𝐞\mathrm{\Theta }(𝐜_{12}𝐞)|𝐜_{12}𝐞|}`$ (7) $`\times \left\{ϵ^2\stackrel{~}{f}(𝐜_1^{})\stackrel{~}{f}(𝐜_2^{})\stackrel{~}{f}(𝐜_1)\stackrel{~}{f}(𝐜_2)\right\}`$ (8) and with the moments of the dimensionless collision integral : $$\mu _p𝑑𝐜_1c_1^p\stackrel{~}{I}(\stackrel{~}{f},\stackrel{~}{f}),$$ (9) while the time-evolution of temperature reads: $$dT/dt=(2/3)BT\mu _2$$ (10) where $`B=B(t)v_0(t)g_2(\sigma )\sigma ^2n`$. To proceed we use the Sonine polynomial expansion for the velocity distribution function $$\stackrel{~}{f}(𝐜)=\varphi (c)\left\{1+\underset{p=1}{\overset{\mathrm{}}{}}a_pS_p(c^2)\right\}$$ (11) where $`\varphi (c)\pi ^{d/2}\mathrm{exp}(c^2)`$ is the Maxwellian distribution and the first few Sonine polynomials read: $`S_0(x)=1`$, $`S_1(x)=x^2+\frac{3}{2}`$, $`S_2(x)=\frac{x^2}{2}\frac{5x}{2}+\frac{15}{8}`$, etc. Multiplying both sides of Eq. (6) with $`c_1^p`$ and integrating over $`d𝐜_1`$, we obtain : $$\frac{\mu _2}{3}pc^p=\mu _p$$ (12) where integration by parts has been performed and where we define $$c^pc^p\stackrel{~}{f}(𝐜,t)𝑑𝐜.$$ (13) The odd moments $`c^{2n+1}`$ are zero, while the even ones, $`c^{2n}`$, may be expressed in terms of $`a_k`$ with $`0kn`$. Calculations show that $`c^2=\frac{3}{2}`$, implying $`a_1=0`$, according to the definition of the temperature (1) (e.g. ), and that $`c^4=\frac{15}{4}\left(1+a_2\right)`$. Now we assume, that the dissipation is not large, so that the deviation from the Maxwellian distribution may be accurately described only by the second term in the expansion (11) with all high-order terms with $`p>2`$ discarded. Then (12) is an equation for the coefficient $`a_2`$. Using the above results for $`c^2`$ and $`c^4`$ it is easy to show that Eq. (12) converts for $`p=2`$ into identity, while for $`p=4`$ it reads: $$5\mu _2\left(1+a_2\right)\mu _4=0$$ (14) The coefficients $`\mu _p`$ may be expressed in terms of $`a_2`$ due to the definition (9) and the assumption $`\stackrel{~}{f}=\varphi (c)[1+a_2S_2(c^2)]`$. Using the properties of the collision integral one obtains for $`\mu _p`$ : $`\mu _p={\displaystyle \frac{1}{2}}{\displaystyle 𝑑𝐜_1𝑑𝐜_2𝑑𝐞\mathrm{\Theta }(𝐜_{12}𝐞)|𝐜_{12}𝐞|\varphi (c_1)\varphi (c_2)}`$ (15) $`\left\{1+a_2\left[S_2(c_1^2)+S_2(c_2^2)\right]+a_2^2S_2(c_1^2)S_2(c_2^2)\right\}\mathrm{\Delta }(c_1^p+c_2^p)`$ (16) where $`\mathrm{\Delta }\psi (𝐜_i)\left[\psi (𝐜_i^{})\psi (𝐜_i)\right]`$ denotes change of some function $`\psi (𝐜_i)`$ in a direct collision. Calculations, similar to that, described in , yield the following result (some detail are given in ): $$\mu _2=\sqrt{2\pi }(1ϵ^2)\left(1+\frac{3}{16}a_2+\frac{9}{1024}a_2^2\right)$$ (17) and $$\mu _4=4\sqrt{2\pi }\left\{T_1+a_2T_2+a_2^2T_3\right\}$$ (18) with $`T_1={\displaystyle \frac{1}{4}}(1ϵ^2)\left({\displaystyle \frac{9}{2}}+ϵ^2\right)`$ (19) $`T_2={\displaystyle \frac{3}{128}}(1ϵ^2)(69+10ϵ^2)+{\displaystyle \frac{1}{2}}(1+ϵ)`$ (20) $`T_3={\displaystyle \frac{1}{64}}(1+ϵ)+{\displaystyle \frac{1}{8192}}(1ϵ^2)(930ϵ^2)`$ (21) The coefficients $`\mu _2`$ and $`\mu _4`$ were provided in Ref. up to terms of the order of $`𝒪(a_2)`$. One obtains the coefficient $`a_2`$ in the Sonine polynomial expansion in this approximation by substituting (17,18) into (14) and discarding in Eqs. (17,18) all terms of the order of $`𝒪(a_2^2)`$: $$a_2^{\mathrm{NE}}=\frac{16(1ϵ)(12ϵ^2)}{8117ϵ+30ϵ^2(1ϵ)}$$ (22) Calculations including the next order terms $`𝒪(a_2^2)`$ in the coefficients $`\mu _2`$ and $`\mu _4`$ show that Eq. (14) is a cubic equation, which for physical values of $`ϵ`$, $`0ϵ1`$, has three different real roots, as it shown on Fig. 1. FIG. 1.: The left hand side of Eq. 14 over $`a_2`$ for $`ϵ=0.8`$. Obviously Eq. 14 has three real solutions. Although the cubic equation may be generally solved, the resultant expressions for the roots are too cumbersome to be written explicitly. However, one of the roots (the middle one) is rather small and close to that given by Eq. (22), obtained within the linear approximation. This suggests the perturbative solution of the cubic equation near this root: $$a_2=a_2^{\mathrm{NE}}\left[1\frac{1005(1ϵ^2)4096T_3}{6080(1ϵ^2)4096T_2}a_2^{\mathrm{NE}}+\mathrm{}\right]$$ (23) where we do not write explicitly terms of the order $`𝒪\left([a_2^{\mathrm{NE}}]^3\right)`$ and high-order terms. In Fig. 2 the dependence of $`a_2^{\mathrm{NE}}`$ and of the corresponding improved value $`a_2`$ are shown as a function of the restitution coefficient $`ϵ`$. As one can see from Fig. 2 the maximal deviation between these is less than $`10\%`$ at small $`ϵ`$ and decreases as $`ϵ`$ tends to $`1`$. The other two roots, shown on Fig. 3 are of the order of 1 or 10, i.e. are not small. Physically, this means that one can not cut the Sonine polynomial expansion in this case at the second term and next order terms are not negligible to be discarded. Taking into account the next order terms, i.e., releasing the assumption that $`a_p0`$ for $`p>2`$, breaks down the above analysis, since the coefficients $`\mu _2`$, $`\mu _4`$ occur to be dependent not only on $`a_2`$, but on $`a_3`$, $`a_4,\mathrm{}`$ as well. Thus the occurrence of several roots for the $`a_2`$, found within the above approach, which satisfy the conditions required by the scaling ansatz (5) does not imply the existence of several different scaling solutions. Nevertheless such possibility may not be completely excluded. If one assumes that few scaling distributions of the velocity may realize, depending on the initial conditions at which the HC state has been prepared, a natural question arises: Whether the particular scaling solution is stable with respect to small perturbations, and what is the domain of attraction of this particular scaling solution in some parametric space. FIG. 3.: The other two solutions for second Sonine coefficient $`a_2`$ of Eq. 14 over the coefficient of restitution $`ϵ`$. Certainly, the stability problem is very complicated to be solved in general. Therefore, we restrict ourselves to the stability analysis of the scaling distribution (5) where the scaling function $`\stackrel{~}{f}(𝐜)`$ has nonzero value of the coefficient $`a_2`$, while the other coefficients $`a_p`$ with $`p>2`$ are negligibly small. (For this scaling solution our above results for the coefficients $`\mu _2`$, $`\mu _4`$ are valid). Moreover, we assume, that small perturbations of the (vanishingly small) coefficients $`a_p`$ with $`p>2`$ do not influence the stability of the distribution, and analyze the stability only with respect to variation of the coefficient $`a_2`$. To analyze the stability of the velocity distribution we write it in a more general form: $$f(𝐯,t)=\frac{n}{v_0^3(t)}\stackrel{~}{f}(𝐜,t)$$ (24) which leads, as it easy to show, to the following generalization of Eq. (6: $$\frac{\mu _2}{3}\left(3+c_1\frac{}{c_1}\right)\stackrel{~}{f}(𝐜,t)+B^1\frac{}{t}\stackrel{~}{f}(𝐜,t)=\stackrel{~}{I}(\stackrel{~}{f},\stackrel{~}{f})$$ (25) with the collisional integral and coefficients $`\mu _p`$ being now time-dependent. The quantities $`c^p`$ also depend now on time. The temperature, however, evolves still according to (10). Using $`\stackrel{~}{f}=\varphi (c)[1+a_2(t)S_2(c^2)]`$ and performing essentially the same manipulations which led before to Eq. (14), we arrive at the following equation for the coefficient $`a_2(t)`$: $$\dot{a}_2(4/3)B\mu _2\left(1+a_2\right)+(4/15)B\mu _4=0$$ (26) with $`\mu _2`$, $`\mu _4`$ still given by (17,18), but with the time-dependent coefficient $`a_2(t)`$. Writing the above value $`B(t)`$ as $`B(t)`$ $`=`$ $`(8\pi )^{1/2}\tau _c(0)^1u(t)^{1/2}`$ (27) $`\tau _c(0)^1`$ $``$ $`4\pi ^{1/2}g_2(\sigma )\sigma ^2nT_0^{1/2},`$ (28) where $`\tau _c(0)`$ is related to the initial mean-collision time at the initial temperature $`T_0`$, and $`u(t)T(t)/T_0`$ is the reduced temperature, we recast Eq. (26) into the form: $$\frac{da_2}{d\widehat{t}}=\frac{\sqrt{2/\pi }}{15}u^{1/2}F(a_2)$$ (29) where $`\widehat{t}`$ is the reduced time, measured in units of $`\tau _c(0)`$, and where we define a function: $$F(a_2)5\mu _2(1+a_2)\mu _4.$$ (30) The form of the function $`F(a_2)`$ for some particular value of $`ϵ`$ is shown on Fig. 1. This form of $`F(a_2)`$ persists for all physical values of the restitution coefficient, $`0ϵ1`$. This has three different roots, $`F(a_2^{(i)})=0`$, $`i=1,2,3`$, which makes $`da_2/dt`$ vanish yielding the scaling form for the solution of the Enskog-Boltzmann equation. The stability of the scaling solution, corresponding to $`a_2^{(i)}`$ requires for the derivative $`dF/da_2`$, taken at $`a_2^{(i)}`$ to be negative, since only in this case a small deviation $`a_2a_2^{(i)}`$ from $`a_2^{(i)}`$, corresponding to a scaling solution will decay with time. As one can see from Fig. 1 only the middle root, which corresponds to small values of $`a_2`$, and is close to $`a_2^{\mathrm{NE}}`$, predicted by linear theory , has negative $`dF/da_2`$, and thus is stable. We also observed that for any $`0ϵ1`$ the point $`a_2=0`$ belongs to the attractive interval of this stable root. Naturally, this means that initial Maxwellian distribution will relax to the non-Maxwellian with $`a_2a_2^{\mathrm{NE}}`$. Note that relaxation of any (small) perturbation to this value of $`a_2`$ occurs, as it follows from Eq. (29), on the collision time-scale, i.e., practically “immediately” on the time-scale which describes the evolution of the temperature. Therefore we conclude, that the scaling solution of the Enskog-Boltzmann equation with $`a_2`$ corresponding to the middle root of the function $`F(a_2)`$, given with a high accuracy by Eqs. (23,22), and with negligibly small other coefficients $`a_3`$, $`a_4,\mathrm{}`$ of the Sonine polynomial expansion is a stable one with respect to (relatively) small perturbations. In conclusion, we analyzed the velocity distribution function of the granular gas with a constant restitution coefficient at the regime of the homogeneous cooling. We assume that the deviations from the Maxwellian distribution may be described using only the leading term in the Sonine polynomial expansion, with all other high-order terms discarded. In this approach the deviations from the Maxwellian distribution are completely characterized by the magnitude of the coefficient $`a_2`$ of the leading term. We go beyond previous linear theories and perform a complete analysis (on the level of the description chosen), without discarding any nonlinear with respect to $`a_2`$ terms. Performing the stability analysis of the scaling solution of the Enskog-Boltzmann equation we observe, that only one value of $`a_2`$, obtained within our nonlinear analysis corresponds to a stable scaling solution. We also report a corrections for this value of $`a_2`$ with respect to the previous result of the linear theory. This corrections are small (less than $`10\%`$) for all values of the restitution coefficient $`ϵ`$ and vanishes as $`ϵ`$ tends to unity in the elastic limit.
no-problem/9906/cond-mat9906337.html
ar5iv
text
# Levy-Nearest-Neighbors Bak-Sneppen Model \[ ## Abstract We study a random neighbor version of the Bak-Sneppen model, where ”nearest neighbors” are chosen according to a probability distribution decaying as a power-law of the distance from the active site, $`P(x)|xx_{ac}|^\omega `$. All the exponents characterizing the self-organized critical state of this model depend on the exponent $`\omega `$. As $`\omega 1`$ we recover the usual random nearest neighbor version of the model. The pattern of results obtained for a range of values of $`\omega `$ is also compatible with the results of simulations of the original BS model in high dimensions. Moreover, our results suggest a critical dimension $`d_c=6`$ for the Bak-Sneppen model, in contrast with previous claims. \] Since its introduction, the Bak-Sneppen (BS) model has had much success as perhaps the simplest and yet non-trivial self-organized critical (SOC) extremal model. Understanding its behavior is therefore very important to get an insight in the behavior of other SOC extremal models . The BS model is easily defined: To each site $`i`$ on a hypercubic lattice in $`d`$ dimensions is assigned a random variable $`f_i`$ taken from a probability distribution $`p(f)`$, say, uniform in $`[0,1]`$. Then at each time-step the site $`i`$ with the smallest $`f_i`$ is chosen (it is called the active site), and its variable and the variables of its $`2d`$ nearest-neighbors are updated taking them from $`p(f)`$. As a result of this dynamics, the system organizes in a stationary state where almost all the variables $`f_i`$ are above a threshold $`f_c`$. Moreover, in this state, the dynamics of the model has self-similar features: each update of the minimum variable triggers a local avalanche of updates; the time durations of the avalanches obey a power-law distributions characterized by an exponent $`\tau `$. The number of sites touched by an avalanche of duration $`t`$ grows like $`t^\mu `$. Also the first return times (defined as the times between two successive returns of the activity to the same site) and the all return times (the times between the first passage of the activity on a site and any successive return to the same site) are power-law distributed, with exponents $`\tau _f`$ and $`\tau _a`$ respectively. Not all of the above exponents are independent. Indeed it is possible to show, from renewal theory, that $`\tau _a+\tau _f=2`$ if $`\tau _a<1`$, $`\tau _a=\tau _f`$ if $`\tau _a>1`$ . Recently, a non trivial relation between $`\tau `$ and $`\mu `$ has been unveiled in exploiting the hierarchical structure of the update avalanches (each avalanche is made up of smaller avalanches, and so on down to the microscopic scale). Both relations are satisfied for $`d=1`$ with $`\tau 1.07`$ and $`\mu 0.42`$, $`\tau _a0.42`$ and $`\tau _f1.58`$ . The only known exactly solved version of the BS model is the Random Nearest-Neighbor (RNN) model: there ”nearest neighbors” are chosen at random over the lattice. As a result geometric correlations typical of low dimensions are lost, and the RNN can be considered as a Mean Field version of the BS model. The exponents of the RNN model are known to be $`\tau =\tau _a=\tau _f=3/2`$ and $`\mu =1`$. In particular these exponents satisfy both $`\tau _a=\tau _f`$ and the relation between $`\tau `$ and $`\mu `$ proposed in . In this relation has been carefully studied, and it has been ”graphically” explicited (see Fig.1). The knowledge of the two exponent relations has given confidence in high-dimensional simulations whose main result is that the upper critical dimension of the model is $`d_{uc}=8`$ (the upper critical dimension is the dimension above which the exponents should take the RNN values). This conclusion is at odds with previous claims, based on analogies of the BS model with directed percolation, that $`d_{uc}`$ should be $`4`$. A further result from is the presence of two regimes as the dimension $`d`$ of the system increases: for $`d3`$ the model is recurrent ($`\tau _a<1`$; in random walk theory recurrence means that every site of the lattice is touched an infinite number of times with probability one), whereas for $`d4`$ the model is transient ($`\tau _a>1`$; transience means that there is a finite probability, smaller than $`1`$, that a site will be touched by the process), yet non trivial (i.e., different from the RNN model) as long as $`d<8`$. Actually, before , a further exponent relation was believed to hold, namely $`\mu =\tau _a`$. In this relation is shown not to hold for $`d>2`$. Indeed, since $`\mu 1`$ always and $`\tau _a(RNN)=3/2`$, it is straightforward to conclude that at least as soon as $`\tau _a>1`$, $`\tau _a\mu `$. The dimensionality $`d=3`$, with $`\tau _a<1`$ and $`\mu \tau _a`$, can be therefore considered as representative of a further regime within the recurrent one. The BS model shows therefore an extremely rich behavior changing the dimensionality $`d`$ of the system. Yet, high-dimensional simulations are always susceptible of strong finite-size corrections, and the good convergence of the results is difficult to prove. In this Communication we propose a new way to interpolate between the $`d=1`$ and the RNN models: The ”nearest neighbors” of the active site $`x_{ac}`$ are chosen at random over the lattice, but with a probability that is a power-law decreasing function of the distance from it $$P(x)|xx_{ac}|^\omega .$$ (1) As we will show, varying $`\omega `$ we find the same behavioral pattern as found in varying $`d`$. We name this model the Levy-Random-Nearest-Neighbor model (LRNN; here the use of the word Levy is somehow an abuse since we use also $`\omega >3`$). As a loose analogy, we recall that the same idea has been applied also to the $`d=1`$ Ising model with interactions decaying as (1), and indeed it has been found that, varying $`\omega `$ it is possible to go from the $`d=1`$ model to mean-field like results . Simulations are performed over $`1d`$ lattices of up to $`2^{19}`$ sites, with growing sizes showing stability of the exponents. In Fig.1 we show the $`\tau `$ avalanche exponents for different values of $`\omega `$ plotted against the corresponding values of $`\mu `$. All the $`\mu /\tau `$ pairs nicely satisfy the exponent relation between the two exponents obtained in . This is a first important check of the consistency of our simulations and of the exponent relation. In Fig.2 we plot the $`\tau _a`$, $`\tau _f`$ and $`\mu `$ exponents for different values of $`\omega `$ from $`\omega =3`$ to $`\omega =1^+`$ (this extreme value is not shown since simulations become extremely difficult due to the non normalizability of the distribution if $`\omega =1`$). Many important aspects of the model can be discussed looking at Fig.2. We find that the exponent relation between $`\tau _a`$ and $`\tau _f`$ is satisfied both when $`\tau _a<1`$ and when $`\tau _a>1`$. Therefore this result and Fig.1 confirm the validity of the two already known exponent relations. Moreover, we see that the exponents tend to their RNN values as $`\omega 1`$. This result should have been expected. Indeed the probability distribution (1) is normalizable in the thermodynamic limit only as long as $`\omega >1`$; when $`\omega <1`$ then the normalization is ruled by the length of the lattice $$_1^Lx^\omega L^{1\omega }$$ (2) that diverges in the limit of infinite lattice size $`L`$. Therefore the distribution $`P(x)`$, properly normalized, degenerates to $`0`$, just as in the RNN case, where the normalization is $`1/L`$ (case $`\omega =0`$). The opposite case, $`\omega \mathrm{}`$, is also intriguing. Indeed we could naively expect the $`d=1`$ limit to be recovered when $`\omega 3`$. In that case the average distance of the ”neighbors” and its variance are finite, just as for nearest neighbors. We would thus expect the $`\omega >3`$ case to belong to the same universality class as the original BS model, but this conclusion is not correct. In order to shed light on this problem, we also performed simulations taking the neighbors according to a distribution exponentially decreasing with respect to the distance from the active site $`x_{ac}`$. In this case, instead, we nicely recover the known $`d=1`$ exponents of the BS model. We conclude therefore that the presence of diverging moments of order higher than two drives the system out of its nearest neighbor fixed point. The latter holds instead whenever the distribution of the random neighbors has all its moments finite. This result, although non trivial, is not new in (annealed or quenched) disordered systems. For example, it has been shown that diverging moments of order higher than two in the disorder distribution can change the universality class of directed polymers in random environments, and of the related Kardar-Parisi-Zhang surface growth equation. It is well known in classical random walk theory that random walks with a jump probability distribution with finite variance belong to the Gaussian universality class. Random walks with infinite higher moments have a microscopic structure that is different from a Gaussian one, mainly made of clusters of points seldom separated by long jumps. On larger and larger length scales, this cluster structure disappears, and the walks ”renormalize” to Gaussian ones. Yet, a main difference between a simple random walk and the BS model is the presence of memory effects. Indeed, any time a site is chosen (either as the active one or as one of its neighbors), any memory of its previous updates is lost. Therefore, the interaction between the geometric structure given by the choice of the neighbors according to $`P(x)`$ and the updates of the corresponding variables can give rise to non-trivial effects. As already mentioned above, the geometrical fractal dimension $`D_f`$ of the avalanches is another quantity of interest. It is possible to relate $`D_f`$ to the other exponents of the model. Indeed, we recall that the fractal dimension relates the volume $`N`$ of the avalanche (that is, the number of sites touched by the avalanche) with the typical size $`R`$ of the avalanche, as $`NR^{D_f}`$. On the other hand, $`N`$ scales with the duration of the avalanche as $`Nt^\mu `$. The relation between the typical size of the avalanche $`R`$, and its duration $`t`$ is given by $`tR^z`$, $`z`$ being the dynamical exponent of the model. Then we find $`D_f=z\mu `$. In principle, in order to determine $`z`$ it is possible to use the all return time distribution $`P_a(t)`$. At long times on a $`d`$dimensional lattice of $`N=L^d`$ sites, $`P_a(t)`$ flattens. W e can write a scaling form for $`P_a(t)`$ as $$P_a(t,L)=t^{\tau _a}f\left(\frac{t}{L^z}\right)$$ (3) with $`f(x)const`$ when $`x0`$ and $`f(x)x^{\tau _a}`$ when $`x\mathrm{}`$. In this second case we find that $`P_a(t,L)=L^d`$ (roughly speaking, as soon as every site of the lattice has been touched by an avalanche, they have all the same probability to be chosen), from which we have $`z=d/\tau _a`$. Then we can write an expression for the fractal dimension $$D_f=d\frac{\mu }{\tau _a}.$$ (4) This expression holds for the high dimensional simulations of : when avalanches are compact objects ($`d=1,2`$), $`\mu =\tau _a`$. Then, $`\mu \tau _a`$ and avalanches become fractal objects. Indeed (4) approximates well the data given in for $`d=3`$ ($`\tau _a=0.92`$, $`\mu =0.85`$, $`D_f=2.63\mu /\tau _a=2.77\mathrm{}`$) and for $`d=4`$ ($`\tau _a=1.15`$, $`\mu =0.92`$, $`D_f=3.34\mu /\tau _a=3.2`$). As a byproduct, we find that (4) suggests an upper critical dimension for the Bak-Sneppen model $`d_{uc}=6`$: indeed with $`\mu =1`$ and $`\tau =3/2`$ (the ”mean-field”, RNN, values of the exponents) we find $`D_f(d=6)=4`$, that is indeed the predicted avalanche fractal dimension in the RNN limit. $`d_{uc}=6`$ is at odds with what stated in , where $`d_{uc}=8`$ was suggested by numerical simulations, but also with where $`d_{uc}=4`$ was claimed based on analogies with directed percolation. Although (4) seems to hold in the high dimensional case, it does not fit the numerical results in the present LRNN approach. The reason is that, whereas in the high dimensional case there is a single relation between time and space, namely $`tL^{d/\tau _a}`$, in the LRNN case there is a further relation, the usual Levy random walk law $`tR^{\omega 1}`$. We measure the fractal dimension of avalanches using the distance between the rightmost and leftmost touched sites as a measure of $`R`$, and this corresponds to $`z=\omega 1`$. Indeed, as it can be seen from Fig.3, $`D_f=\mu (\omega 1)`$ approximates very well the measured fractal dimensions. As noted above about the fractal dimension of high-dimensional avalanches, we observed that $`\mu =\tau _a`$ corresponds to compact avalanches. In the LRNN case we see from Fig.2 that indeed $`\mu \tau _a`$ for $`\omega <2`$, even if the fractal dimension $`D_f<1`$ already for $`\omega <3`$. We can try to understand this result remembering that $`D_f`$ is related to the random walk exponent $`z=\omega 1`$: although $`z`$ is different from its Gaussian value $`z=2`$ as soon as $`\omega <3`$, a Levy random walk with $`\omega >2`$ is still compact, and so is the structure built by a choice of neighbors according to (1). Only when $`\omega <2`$ such a structure becomes genuinely fractal, and $`\mu \tau _a`$. The possibility to obtain the exponent $`D_f`$ from $`\mu `$ and $`\tau _a`$ in high dimensions, and from $`\mu `$ and $`\omega `$ in the LRNN version, suggests that indeed there are at most two independent exponents in the model, namely $`\mu `$ and $`\tau _a`$. The LRNN model suggests the presence of a further (although non trivial) exponent relation. Indeed, in Fig.4 we show the values of the $`\tau _a`$ exponent as a function of the corresponding $`\mu `$ exponent for different values of $`\omega `$ and for different dimensions. As it can be seen, the agreement is good, suggesting that the knowledge of $`\mu `$ (or of $`\tau _a`$) is sufficient to know all the other exponents through relations that, as for the $`\mu /\tau `$ one, could be highly non-trivial. In conclusion, we have introduced a modification of the Bak-Sneppen model where the neighbors of the active site are chosen at random over the lattice with a probability that decreases like a power-law of the distance from the active site, with an exponent $`\omega `$. As a result we find that the characteristic exponents of the model interpolate between the $`d=1`$ limit ($`\omega =\mathrm{}`$) and the mean field (RNN) limit ($`\omega 1`$). In particular, we verify that the known exponent relations hold for this model too. Moreover, we find and verify an exponent relation for the fractal dimension of the avalanches, $`D_f=(\omega 1)\mu `$. As a byproduct we obtain a relation between $`D_f`$ and $`\mu `$ and $`\tau _a`$ also in high dimensions, fitting well the present numerical results up to $`d=4`$, although it suggests an upper critical dimension $`d_{uc}=6`$ (and not $`d_{uc}=4`$ or $`d_{uc}=8`$ as previously believed). More accurate numerical simulations in high dimensions are therefore needed. The relevance of the results reported in this Communication is manifold: they can be looked at as an interesting modification of the Bak-Sneppen model, but their full importance emerges when compared to the high dimensional results of . Indeed, they lead us to propose a new value of the upper critical dimension of the model, namely $`d_{uc}=6`$, and to conjecture the existence of a still undiscovered exponent relation between $`\mu `$ and $`\tau _a`$, reducing therefore the number of independent exponents to one in any dimension. This results show therefore that there is still some way to go before a full and satisfying understanding of the BS model is achieved. The authors thank F. Slanina for useful discussions. R. Cafiero and P. De Los Rios aknowledge financial support under the European network project FMRXCT980183.
no-problem/9906/astro-ph9906226.html
ar5iv
text
# Formation of Large-Scale Obscuring Wall and AGN Evolution Regulated by Circumnuclear Starbursts ## 1 Introduction Recently, intriguing evidences regarding the host galaxies of active galactic nuclei (AGNs) and quasars (QSOs) have been accumulated. First, it has been reported that host galaxies of Seyferts are intrinsically unlike between type 1 (Sy1) and type 2 (Sy2) (Heckmann et al. 1989; Maiolino et al. 1995, 1997, 1998, 1999; Pérez-Olea & Colina 1996; Hunt et al. 1997; Malkan et al. 1998; Storchi-Bergmann & Schmitt 1998). Sy1’s inhabit earlier type quiescent hosts, while Sy2’s are frequently associated with circumnuclear starbursts, which often lie in barred galaxies. The observations may indicate that Sy2’s are in an earlier evolutionary stage than Sy1’s (Radovich et al. 1998), whereas in the unified model (Antonucci 1993, for a review) this dichotomy is simply accounted for with the orientation of the nucleus with an obscuring torus of subparsec scale. Second, the recent HST images of nearby QSOs have shown that luminous QSO phenomena occur preferentially in luminous host galaxies, often being ellipticals (McLeod & Rieke 1995b; Bahcall et al. 1997; Hooper et al. 1997). Also, at high redshifts, towards the QSO H1413+117 at $`z=2.546`$ (“cloverleaf”, a gravitationally lensed quasar) and QSO BR1202-0725 at $`z=4.69`$, a large amount of dust have been detected, i.e., $`10^9M_{}`$ for H1413+117 and $`10^8M_{}`$ for BR1202-0725 (Barvainis et al. 1992; Omont et al. 1996). Molecular gas of at least $`10^{11}M_{}`$ are also found for BR1202-0725 (Ohta et al. 1996a). These suggest that active star formation are on-going around the quasars. QSOs, however, are mostly identified as type 1 and only a few type 2 QSOs are discovered so far (Almaini et al. 1995; Ohta et al. 1996b; Brandt et al. 1997). Hence, a circumnuclear starburst does not seem to urge type 2 as far as quasars are concerned. QSOs are distinctive from Seyferts in that the host galaxy is in general fainter than the AGN itself (McLeod & Rieke 1995b; Bahcall et al. 1997; Hooper et al 1997). These facts on the host properties for Seyferts and QSOs suggest a possibility that the AGN type has a close relation to circumnuclear starburst events and the relative luminosity of starburst to the AGN. We consider a physical mechanism which may connect circumnuclear star-forming activities with AGN type. Here, attention is concentrated on the radiative force by the circumnuclear starburst. In ultraluminous IR galaxies, the observed IR luminosities (Scoville et al. 1986; Soifer et al. 1986) are comparable to or greater than the Eddington luminosity for dust opacity (Umemura et al. 1998, 1999). Hence, the radiative force is very likely to play an important role on the circumnuclear structure of $`100`$ pc. In this Letter, supposing the mass distribution in circumnuclear regions, we analyze the equilibrium configuration and the stability of dusty gas which is supported by radiative force by a starburst and an AGN. In addition, taking the stellar evolution in the starburst regions into consideration, we investigate the time evolution of gas distributions, and attempt to relate the luminosity of the circumnuclear starburst to the evolution of the AGN type. ## 2 Radiatively-Supported Obscuring Wall The circumnuclear starburst regions frequently exhibit ring-like features and have radial extension of $`10`$ pc up to kpc (Wilson et al. 1991; Forbes et al. 1994; Mauder et al. 1994; Buta et al. 1995; Barth et al. 1995; Maoz et al. 1996; Leitherer et al. 1996; Storchi-Bergman et al. 1996). Thus, here we consider a ring of starburst. (If the starburst regions are anisotropic or clumpy, the effects are expected to be smeared out as discussed later.) We calculate the radiation force and the gravity which are exerted on the dusty gas. Here, the gravitational potential is determined by four components, the galactic bulge, the central black hole, the gas disk, and the starburst ring. We assume the galactic bulge to be an uniform sphere whose mass and radius are $`M_{\mathrm{bul}}`$ and $`R_{\mathrm{bul}}`$, the mass of the central black hole to be $`M_{\mathrm{BH}}`$, the gas disk to be a Mestel disk whose mass and radius are $`M_{\mathrm{disk}}`$ and $`R_{\mathrm{disk}}`$, and the starburst ring to be an uniform torus whose mass, curvature radius, thickness, and bolometric luminosity are $`M_{\mathrm{SB}}`$, $`R_{\mathrm{SB}}`$, $`a_{\mathrm{SB}}`$, and $`L_{\mathrm{SB}}`$, respectively (see Figure 1). The observations of IRAS galaxies by Scoville et al. (1991) show that the central regions within several hundred parsecs possess the gas of $`<10^{10}\mathrm{M}_{}`$. By taking this fact into account, we adopt the mass ratio as $`M_{\mathrm{bul}}:M_{\mathrm{BH}}:M_{\mathrm{disk}}:M_{\mathrm{SB}}=1:0.01:0.1:1`$. Also, it is found that the starburst ring often consists of compact star clusters of $`<10`$ pc (Barth et al. 1995; Maoz et al. 1996; Leitherer et al. 1996). The size of these clusters is about a tenth of radial extension of the starburst ring. Therefore, we assume $`R_{\mathrm{bul}}:R_{\mathrm{disk}}:R_{\mathrm{SB}}:a_{\mathrm{SB}}=10:1:1:0.1`$. In addition, the nucleus is postulated to be a point source whose bolometric luminosity is $`L_{\mathrm{nuc}}`$. Here, the material is assumed to be subject to the radiative force directly by the starburst radiation. The radiation flux by an infinitesimal volume element, $`dV`$, of the starburst ring is given by $`dF^i(r,z)=(\rho _{\mathrm{SB}}/4\pi ł^2)n^idV`$, at a point of $`(r,z)`$ in cylindrical coordinates, where $`i`$ denotes $`r`$ or $`z`$, $`\rho _{\mathrm{SB}}`$ ($`=L_{\mathrm{SB}}/2\pi ^2a_{\mathrm{SB}}^2R_{\mathrm{SB}}`$) is the luminosity density of the starburst ring, $`l`$ is the distance from $`(r,z)`$ to this element, and $`n^i`$ is a directional cosine. Hence, the radiation flux force by the starburst ring and the nucleus at $`(r,z)`$ is given by $$f_{\mathrm{rad}}^i=\frac{\chi }{c}\frac{\rho _{\mathrm{SB}}}{4\pi l^2}n^i𝑑V+\frac{\chi }{c}\frac{iL_{\mathrm{nuc}}}{4\pi \left(r^2+z^2\right)^{3/2}},$$ (1) where $`\chi `$ is the mass extinction coefficient for the dusty gas (Umemura et al. 1998) and $`c`$ is the light speed. Using the above equation, the equilibrium between the radiation force and the gravity is written as $$f_{\mathrm{rad}}^z+f_{\mathrm{grav}}^z=0,$$ (2) in the vertical directions and $$\frac{j^3}{r^2}+f_{\mathrm{rad}}^r+f_{\mathrm{grav}}^r=0,$$ (3) in the radial directions, where $`j`$ is the specific angular momentum of the dusty gas and $`f_{\mathrm{grav}}^i`$ is the gravitational force. In figure 2, in the case that the starburst ring is a dominant radiation source (Case A), the resultant equilibrium branches are shown in the $`r`$-$`z`$ space. Here, $`\mathrm{\Gamma }_{\mathrm{SB}}`$ and $`\mathrm{\Gamma }_{\mathrm{nuc}}`$ are the Eddington measures defined by $`\mathrm{\Gamma }_{\mathrm{SB}}=L_{\mathrm{SB}}/(4\pi cGM_{\mathrm{total}}/\chi )`$ and $`\mathrm{\Gamma }_{\mathrm{nuc}}=L_{\mathrm{nuc}}/(4\pi cGM_{\mathrm{total}}/\chi )`$ respectively with $`M_{\mathrm{total}}=M_{\mathrm{bul}}+M_{\mathrm{BH}}+M_{\mathrm{disk}}+M_{\mathrm{SB}}`$. The solid and dashed curves represent the equilibrium branches which are stable in vertical directions. Above the curves, the vertical component of the gravity which works to lower the gas is stronger than the radiation force, while below the curves the radiation force lifts the gas towards the curves. The dotted curves show the vertically unstable branches. The gas is accelerated upwards by the radiative force above the curves or falls downwards by the gravity below the curves. In order to get the configuration of finally stable equilibrium, the stability in radial directions must be taken also into consideration. On the dashed curves, the effective potential in the radial directions turns out to be locally maximal. Thus, the dashed branches are unstable points of saddle type. Finally, only solid curves are stable branches. Figure 2 shows that stable branches emerge only for $`\mathrm{\Gamma }_{\mathrm{SB}}1`$. This agrees with a naive expectation. If $`\mathrm{\Gamma }_{\mathrm{SB}}>1`$, the radiation force blows out the dusty gas in most regions. (e.g. see a dotted curve of $`\mathrm{\Gamma }_{\mathrm{SB}}=2`$). When $`\mathrm{\Gamma }_{\mathrm{SB}}1`$, the covering factor of the stable wall is a function of $`\mathrm{\Gamma }_{\mathrm{SB}}`$. If $`\mathrm{\Gamma }_{\mathrm{SB}}1`$, the wall surrounds both the nucleus and the starburst ring. When the starburst luminosity is smaller than $`\mathrm{\Gamma }_{\mathrm{SB}}=0.55`$, the wall forms only in the vicinity of the starburst ring and exhibits a torus-like configuration. The $`A_V`$ of the wall is expected to be at least several, because radiative force directly from a starburst can be exerted on such a wall. Then, this large-scale wall of dusty gas would work obscure the nucleus. When the flux force of scattered diffuse radiation operates efficiently in the wall, the wall of larger optical depth may be supported and therefore $`A_V`$ could be much larger. (The detail should be investigated by multi-dimensional radiation hydrodynamics, which will be performed in the future analysis.) If $`A_V`$ of the wall is only several magnitudes, the AGN would be changed to an intermediate type between type 1 and 2, e.g. type 1.3, 1.5 and so on, while the $`A_V`$ greater than ten magnitudes would result in the perfect shift from type 1 to type 2. Next, we examine the case that the nucleus is brighter than the circumnuclear starburst (Case B). In this case, only vertically stable branches emerge even if $`\mathrm{\Gamma }_{\mathrm{SB}}+\mathrm{\Gamma }_{\mathrm{nuc}}<1`$ (see Fig. 3). On the dashed curves in Fig. 3, contrastively to the Case A, there is no solution for the radial equilibrium, regardless of the value of starburst luminosity. The gas around the dashed curves is swung away due to the cooperation of radiative force and angular momentum. Resultantly, the formation of the stable wall is precluded in the Case B. This implies that the luminous nuclei like QSOs are not likely to be obscured, which are therefore mostly identified as type 1. Further, for the Case A, we consider the effects of stellar evolution in the starburst regions on the stable equilibrium branches. We assume a Salpeter-type initial mass function (IMF), $`\varphi =A(m_{}/M_{})^{1.35}`$, the mass-luminosity relation, $`(l_{}/L_{})=(m_{}/M_{})^{3.7}`$, and the mass-age relation, $`\tau =1.1\times 10^{10}\mathrm{yr}(m_{}/M_{})^{2.7}`$, where $`m_{}`$ and $`l_{}`$ are respectively the stellar mass and luminosity. Recently, it has been revealed that in starburst regions the IMF is deficient in low-mass stars, with the cutoff of about $`2M_{}`$, and the upper mass limit is inferred to be around $`40M_{}`$ (Doyon et al. 1992; Charlot et al. 1993; Doane & Mathews 1993; Hill et al. 1994; Brandl et al. 1996). Using the IMF for a mass range of $`[2M_{},40M_{}]`$ and the above relations, the total stellar luminosity of starburst regions is given by a function of time as $`L_{}=1.5\times 10^{10}\left(87t_7^{0.87}1\right)\left(M_{SB}/10^{10}M_{}\right)L_{}`$, where $`t_7`$ is the elapsed time after the coeval starburst in units of $`10^7\mathrm{yr}`$. Also, if we postulate that stars of $`>8M_{}`$ are destined to undergo supernova explosions and release the energy radiatively with the efficiency of $`\epsilon `$ to the rest mass energy, the total supernova luminosity is $`L_{\mathrm{SN}}=1.7\times 10^{11}t_7^{0.87}\left(M_{\mathrm{SB}}/10^{10}M_{}\right)\left(\epsilon /10^4\right)L_{}`$ until $`t_7=4.0`$ and $`L_{\mathrm{SN}}=0`$ when $`t_7>4.0`$. Hence, the total luminosity of the starburst ring is given by $$L_{\mathrm{SB}}(t_7)=L_{}+L_{\mathrm{SN}}.$$ (4) Using this dependence on time, the luminosity can be translated into the age of the starburst regions. Therefore, the values of $`\mathrm{\Gamma }_{\mathrm{SB}}`$ in figure 2 represent the evolutionary stage of the circumnuclear starburst. For instance, if we adopt $`M_{\mathrm{SB}}=10^{10}M_{}`$, $`\mathrm{\Gamma }_{\mathrm{SB}}=1`$ and $`0.55`$ correspond to $`4.2\times 10^7\mathrm{yr}`$ and $`8.1\times 10^7\mathrm{yr}`$, respectively. To summarize, if $`\mathrm{\Gamma }_{\mathrm{SB}}>1`$ in the early evolutionary stage, the dusty gas is blown away by radiative acceleration. Since the blown-out dusty gas would emit the strong IR radiation, we may recognize the objects as ultraluminous infrared galaxies. When $`\mathrm{\Gamma }_{\mathrm{SB}}`$ becomes just below unity, both the nucleus and the starburst ring are surrounded by the dusty wall. Then, the AGN is likely to be type 2. In the later stages, the dusty gas forms a torus-like obscuring wall, which shrinks on a time-scale of several $`10^7`$ yr. Then, the AGN tends to be identified as type 1 for a wide viewing angle. This implies that the type of AGN evolves from higher to lower in several $`10^7`$ yr according as the circumnuclear starburst becomes dimmer. ## 3 Discussion Here, we have assumed that the nuclear activity and the circumnuclear starburst are the simultaneous events. A solution, for instance, which links the two events, is the radiatively-driven mass accretion onto a central black hole due to the radiation drag (Umemura et al. 1997, 1998; Ohsuga et al. 1999). However, in very early luminous phases of the starburst, the mass accretion onto the black hole is prevented due to the super-Eddington radiative force. It result in a radiative blizzard in nuclear regions. Thus, the nucleus is just identified as an ultraluminous infrared galaxy without being accompanied by an AGN. We predict in the present model that ultraluminous infrared galaxies evolve into Seyferts or QSOs in later less luminous phases of the starburst. The circumnuclear starburst could not be axisymmetric but clumpy. However, the rotational time scale of the starburst ring is shorter than the shrinking time-scale of the obscuring wall. In the present case, the former is around $`3.0\times 10^6\mathrm{yr}`$ and the latter is several $`10^7\mathrm{yr}`$. Therefore, the anisotropies of the starburst are expected to be smeared out by a ’wheel effect’. The stable obscuring wall might be subject to the other local instabilities, i.e., Rayleigh-Taylor or self-gravitational instabilities. The density gradient of the dusty wall is positive inside the equilibrium surface and negative outside the surface. They are in the same directions as the effective acceleration. Thus, the wall would not be subject to Rayleigh-Taylor instabilities. As for the self-gravitational instability, the time-scale of the instability could be as short as $`<10^6`$ yr. So, the wall may fragment on a time-scale shorter than the evolutionary time-scale of the wall. Then, numerous compact gas clouds would form in the wall. They would emit the narrow emission lines because they have the velocity dispersion of several 100 km s<sup>-1</sup>. Also, if the compact clouds are optically thick, the radiative force is less effective for them, so that they fall into the central regions. They also may partially obscure the nucleus. In this letter, we do not argue that a conventional obscuring torus of subparsec scale is dispensable. Even if the inner obscuring torus may operate to intrinsically differentiate the type of AGNs, the present large-scale wall can work also to raise further the type index. In particular, the present mechanism may provide a physical solution to account for the tendency that Sy2’s are more frequently associated with circumnuclear starbursts than Sy1’s, whereas quasars are mostly observed as type 1 regardless of star-forming activity in the host galaxies. If we adopt the size and mass that conform to realistic values, e.g., $`R_{\mathrm{SB}}`$ 100 pc and $`M_{\mathrm{SB}}10^{10}M_{}`$, the obscuring wall is extended to several 100 pc. Interestingly, it is recently reported that the spectra of a sample of AGNs are more consistent with obscuring material extended up to $`>`$$``$ 100 pc around the nuclei (Rudy et al. 1988; Miller et al. 1991; Scarrott et al. 1991; Goodrich 1995; McLeod & Rieke 1995a; Maiolino et al. 1995; Maiolino & Rieke 1995) and the hosts of Sy’2 possess more frequently extended dust lanes (Malkan et al. 1998). Also, the covering factor of a dusty torus around a QSO, MG 0414+0534, is fairly small (Oya et al. 1999). These observations are quite intriguing in the light of the present picture. ###### Acknowledgements. We are grateful to T. Nakamoto, H. Susa, and S. Oya for helpful discussion. The calculations were carried out at the Center for Computational Physics in University of Tsukuba. This work is supported in part by Research Fellowships of the Japan Society for the Promotion of Science for Young Scientists, 6957 (KO) and the Grants-in Aid of the Ministry of Education, Science, Culture, and Sport, 09874055 (MU).
no-problem/9906/hep-th9906113.html
ar5iv
text
# Lamé Instantons ## Abstract We perform a precise analytic test of the instanton approximation by comparing the exact band spectrum of the periodic Lamé potential to the tight-binding, instanton and WKB approximations. The instanton result gives the correct leading behavior in the semiclassical limit, while the tight-binding approximation does even better. WKB is off by an overall factor of $`\sqrt{e/\pi }`$. Periodic quantum systems arise in many areas of physics, from crystal structures in solid state physics, to optical lattices in AMO physics, to solitons in polymer physics, and to the vacuum structure of QCD. An important general feature of these systems is the phenomenon of quantum tunneling that broadens discrete energy spectra into bands. Various semiclassical techniques have been developed to analyze the spectra of such systems: the tight-binding, WKB, and instanton approximations. In quantum field theory, the instanton approach provides important insights into symmetry breaking and the QCD vacuum . In this paper we compare these approximations for an exactly solvable periodic system – the Lamé model – with particular emphasis on the instanton approximation for the width of the lowest energy band. Two well-studied quantum mechanical instanton models are the double-well, $`V(\varphi )=\frac{1}{2}\varphi ^2(1\sqrt{g}\varphi )^2`$, and the ‘Sine-Gordon’, $`V(\varphi )=\frac{1}{g}(1\mathrm{cos}(\sqrt{g}\varphi ))`$. The Lamé model, $`V(\varphi )=\frac{1}{2g}\mathrm{sn}^2(\sqrt{g}\varphi |\nu )`$, provides a new example, and has the advantage that it can be solved exactly without the instanton approximation, while the instanton calculation can also be done analytically. We find that there is a difference between the semiclassical and tight-binding limits, limits that are often regarded as synonymous. We also highlight a new connection between the non-perturbative instanton approach and the algebraic approach to spectra . Consider the Schrödinger-like Lamé equation $`{\displaystyle \frac{d^2}{d\varphi ^2}}\mathrm{\Psi }(\varphi )+j(j+1)\nu \mathrm{sn}^2(\varphi |\nu )\mathrm{\Psi }(\varphi )=\mathrm{\Psi }(\varphi )`$ (1) where $`j=1,2,3,\mathrm{}`$ is an integer, $`\mathrm{sn}(\varphi |\nu )`$ is one of the Jacobi elliptic functions, and $`0\nu 1`$ is the elliptic parameter. This equation is known to be exactly solvable in terms of theta functions, and is known to have a spectrum of $`j`$ bound bands . The function $`\mathrm{sn}^2(\varphi |\nu )`$ is periodic, with period $`2K(\nu )`$, where $`K(\nu )=_0^{\pi /2}𝑑\theta /\sqrt{1\nu \mathrm{sin}^2\theta }`$ is the elliptic quarter period. For small $`\nu `$ (say $`\nu <0.2`$), $`\mathrm{sn}^2(\varphi |\nu )`$ looks to the eye like $`\mathrm{sin}^2(\varphi )`$. For large $`\nu `$ (say $`\nu >0.9`$), $`\mathrm{sn}^2(\varphi |\nu )`$ looks like a sequence of periodically displaced Pöschl-Teller $`\mathrm{sech}^2(\varphi )`$ potentials (see Fig. 1), displaced with a period that diverges logarithmically, $`2K(\nu )\mathrm{log}(\frac{16}{1\nu })`$, as $`\nu 1`$. While the Lamé equation is exactly solvable for any $`j`$ and $`\nu `$, explicit expressions for the wavefunctions and energies are cumbersome for $`j2`$ . A dramatic simplification is that the band-edge energies of (1) are given by the $`2j+1`$ eigenvalues of the finite dimensional matrix $`J_x^2+\nu J_y^2`$, where $`J_x`$ and $`J_y`$ are $`su(2)`$ generators in a spin $`j`$ representation . Thus we can evaluate the width $`\mathrm{\Delta }`$ of the lowest band as the difference between the two smallest eigenvalues of the matrix $`J_x^2+\nu J_y^2`$. For example, for $`j=1`$, $`\mathrm{\Delta }^{\mathrm{exact}}=1\nu `$; for $`j=2`$, $`\mathrm{\Delta }^{\mathrm{exact}}=1\nu +2\sqrt{\nu ^2\nu +1}`$; and for $`j=3`$, $`\mathrm{\Delta }^{\mathrm{exact}}=3(1\nu )+2\sqrt{1\nu +4\nu ^2}2\sqrt{4\nu +\nu ^2}`$. We will be interested later in large values of $`j`$, in which case such explicit expressions, as functions of $`\nu `$, become more difficult to derive. Instead, we have shown algebraically that for any $`j`$, the exact energy splitting of the lowest band, as $`\nu 1`$, is $`\mathrm{\Delta }^{\mathrm{exact}}={\displaystyle \frac{8j\mathrm{\Gamma }(j+1/2)}{4^j\sqrt{\pi }\mathrm{\Gamma }(j)}}(1\nu )^j\left(1+{\displaystyle \frac{j1}{2}}(1\nu )+\mathrm{}\right)`$ (2) As $`\nu 1`$ the period becomes infinite, which suppresses tunneling, and we therefore expect this to be relevant for the tight-binding and semiclassical approximations. Note that the band-width (2) vanishes as $`\nu 1`$, and vanishes more rapidly for larger values of $`j`$. We now compare this exact result (2) with the tight-binding, instanton and WKB approximations. In the tight-binding approximation of solid state physics the separation $`L`$ between neighboring wells becomes large and we treat the potential as a sum of periodically displaced ‘atomic’ potential wells: $`V(\varphi )=_nU(\varphi nL)`$. In the Lamé case, this physical approximation is explicitly realized by the remarkable mathematical identity : $`\nu \mathrm{sn}^2(\varphi |\nu )={\displaystyle \frac{E^{}(\nu )}{K^{}(\nu )}}({\displaystyle \frac{\pi }{2K^{}(\nu )}})^2{\displaystyle \underset{n=\mathrm{}}{\overset{\mathrm{}}{}}}\mathrm{sech}^2\left({\displaystyle \frac{\pi }{2K^{}(\nu )}}(\varphi 2nK(\nu ))\right)`$ (3) As expected, each ‘atomic well’ has the form of a Pöschl-Teller well (but the rescaling factor $`\frac{\pi }{2K^{}}`$ is non-obvious). Including the $`j(j+1)`$ factor from (1), each atomic well has $`j`$ discrete bound states, and the effect of the periodic sum is to broaden these states into the $`j`$ bound bands of the Lamé potential. For the lowest band we use the ground state $`\mathrm{\Psi }_0(\varphi )=\sqrt{\frac{\sqrt{\pi }\mathrm{\Gamma }(j+1/2)}{2K^{}\mathrm{\Gamma }(j)}}\mathrm{sech}^j(\frac{\pi }{2K^{}}\varphi )`$ of the atomic well. The width of this band can be calculated using standard solid state techniques : $`\mathrm{\Delta }^{\mathrm{tight}\mathrm{binding}}`$ $`=`$ $`4j(j+1)({\displaystyle \frac{\pi }{2K^{}}})^2{\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\varphi {\displaystyle \underset{n0}{}}\mathrm{sech}^2\left({\displaystyle \frac{\pi }{2K^{}}}(\varphi 2nK)\right)\mathrm{\Psi }_0(\varphi )\mathrm{\Psi }_0(\varphi 2K)`$ (4) $``$ $`{\displaystyle \frac{8j\mathrm{\Gamma }(j+1/2)}{4^j\sqrt{\pi }\mathrm{\Gamma }(j)}}(1\nu )^j\left(1+{\displaystyle \frac{j1}{2}}(1\nu )+\mathrm{}\right)`$ (5) where in the second line we have kept dominant terms as $`\nu 1`$, and used the fact that $`\mathrm{exp}[\pi K(\nu )/K^{}(\nu )]\frac{1\nu }{16}(1+\frac{1}{2}(1\nu )+\mathrm{})`$. This result (5) agrees precisely with the exact result (2) to this order in $`1\nu `$. Therefore, for any $`j`$, the tight-binding approximation is good as $`\nu 1`$; i.e. as the separation between atomic wells becomes large. We now turn to an instanton evaluation of the width of the lowest band, expected to be good in the semiclassical limit in which tunneling effects are small. Naively, one might expect that this is also just the $`\nu 1`$ limit in which the wells become infinitely separated, but it is actually more interesting than this. To make contact with the standard instanton approach we define $`V(\varphi )={\displaystyle \frac{1}{2g}}\mathrm{sn}^2(\sqrt{g}\varphi |\nu )={\displaystyle \frac{1}{2}}\varphi ^2g{\displaystyle \frac{(\nu +1)}{6}}\varphi ^4+\mathrm{}`$ (6) where $`g`$ is some coupling (which we will relate to $`j`$ and $`\nu `$ below), and we have chosen units so that the perturbative mass in a given well is $`1`$. The width of the lowest band can be found by considering the Euclidean path integral connecting two neighboring classical minima of the potential; here $`\varphi =0`$ and $`\varphi =2K/\sqrt{g}`$ for example. Rescaling the field variable to $`\chi =\sqrt{g}\varphi `$, $`\mathrm{exp}\left({\displaystyle \frac{1}{\mathrm{}}}S[\varphi ]\right)=\mathrm{exp}\left({\displaystyle \frac{1}{\mathrm{}g}}{\displaystyle 𝑑t[\frac{1}{2}(\frac{d\chi }{dt})^2+\frac{1}{2}\mathrm{sn}^2(\chi |\nu )]}\right)`$ (7) Thus the semiclassical limit is $`\mathrm{}g1`$, or $`\frac{1}{2}\mathrm{}V_{\mathrm{peak}}`$ : the ground state in each well is far below the barrier height. There is a standard technique for computing the instanton approximation for the width of the lowest band . In the semiclassical limit, the Euclidean path integral is dominated by ‘instanton’ solutions to the Euclidean equations of motion, satisfying the first-order equation $`\dot{\chi }_{\mathrm{inst}}=\sqrt{2V(\chi _{\mathrm{inst}})}`$. Using the rescaled potential $`V(\chi )=\frac{1}{2}\mathrm{sn}^2(\chi |\nu )`$, we find the Lamé instanton: $`\chi _{\mathrm{inst}}(t)=K(\nu )+\mathrm{sn}^1(\mathrm{tanh}(t)|\nu )`$ (8) Here the integration constant has been chosen so that the instanton is centered at $`t=0`$, and $`\mathrm{sn}^1`$ means the inverse function (a standard function in Mathematica). This instanton interpolates between $`\chi =0`$ at $`t=\mathrm{}`$, and $`\chi =2K(\nu )`$ at $`t=+\mathrm{}`$, as shown in Fig. 2. The corresponding Euclidean action is $`S_0={\displaystyle _0^{2K}}𝑑\chi \sqrt{2V(\chi )}={\displaystyle \frac{1}{\sqrt{\nu }}}\mathrm{log}\left({\displaystyle \frac{1+\sqrt{\nu }}{1\sqrt{\nu }}}\right)`$ (9) The leading exponential factor in the instanton expression for the band width is $`\mathrm{exp}[S_0/(\mathrm{}g)]`$. But there is also a prefactor that is related to the determinant of the fluctuation operator $`\frac{d^2}{dt^2}+V^{\prime \prime }(\chi _{\mathrm{inst}}(t))`$. This prefactor is physically significant, as it encapsulates the collective coordinate effects of fluctuations about the instantons. Here, $`V^{\prime \prime }(\chi _{\mathrm{inst}}(t))=(1\nu )\left[{\displaystyle \frac{\nu \mathrm{tanh}^4(t)+2(1\nu )\mathrm{tanh}^2(t)1}{(1\nu \mathrm{tanh}^2(t))^2}}\right]`$ (10) which is plotted in Fig. 3. We see that $`V^{\prime \prime }(\chi _{\mathrm{inst}}(t))`$ is highly localized in the vicinity of the instanton, and tends to $`1`$ (the square of the mass) as $`t\pm \mathrm{}`$, as expected. Coleman has given a very simple method for evaluating the fluctuation determinant (with zero mode removed), resulting in the formula (for a periodic potential) $`\mathrm{\Delta }E^{\mathrm{inst}}=c\mathrm{\hspace{0.17em}4}\mathrm{}{\displaystyle \frac{1}{\sqrt{\pi \mathrm{}g}}}\mathrm{exp}[{\displaystyle \frac{1}{\mathrm{}g}}S_0]`$ (11) where the constant $`c`$ is simply determined by the asymptotic behavior of the zero mode: $`\dot{\chi }_{\mathrm{inst}}(t)ce^{|t|}`$ as $`|t|\mathrm{}`$. Here, the zero mode is $`\dot{\chi }_{\mathrm{inst}}(t)=\mathrm{sech}(t)/\sqrt{1\nu \mathrm{tanh}^2(t)}\frac{2}{\sqrt{1\nu }}e^{|t|}`$. Thus we obtain the Lamé instanton result $`\mathrm{\Delta }E^{\mathrm{inst}}={\displaystyle \frac{8\mathrm{}}{\sqrt{\pi \mathrm{}g}}}{\displaystyle \frac{1}{\sqrt{1\nu }}}\mathrm{exp}[{\displaystyle \frac{1}{\mathrm{}g\sqrt{\nu }}}\mathrm{log}\left({\displaystyle \frac{1+\sqrt{\nu }}{1\sqrt{\nu }}}\right)]`$ (12) To relate this to the eigenvalues $``$ of the Lamé equation (1), we compare (1) with the Schrödinger equation, $`\frac{\mathrm{}^2}{2}\frac{d^2}{d\varphi ^2}\mathrm{\Psi }+\frac{1}{2g}\mathrm{sn}^2(\sqrt{g}\varphi )\mathrm{\Psi }=E\mathrm{\Psi }`$, for the potential $`V(\varphi )`$ in (6). We therefore identify $`\nu j(j+1)={\displaystyle \frac{1}{\mathrm{}^2g^2}},={\displaystyle \frac{2E}{\mathrm{}^2g}}`$ (13) Thus the instanton approximation for the band-width of the eigenvalue $``$ in (1) is $`\mathrm{\Delta }^{\mathrm{inst}}={\displaystyle \frac{16}{\sqrt{\pi }}}\left(\nu j(j+1)\right)^{3/4}\left(1+\sqrt{\nu }\right)^{2\sqrt{j(j+1)}}\left(1\nu \right)^{\sqrt{j(j+1)}1/2}`$ (14) As a function of $`j`$ this differs from the exact and tight-binding expressions (2) and (5), even in the large period limit $`\nu 1`$. However, from (7) and (13), the semiclassical limit $`\mathrm{}g1`$ means $`\nu j(j+1)1`$. So, to compare the instanton formula (14) with the results (2) and (5) it is not enough to take $`\nu 1`$; we also need to take $`j`$ to be large. Physically, it is not enough to take far-separated wells; they must also be deep wells. For large $`j`$ and $`\nu `$ near 1, $`\mathrm{\Delta }^{\mathrm{inst}}{\displaystyle \frac{8j^{3/2}}{\sqrt{\pi }\mathrm{\hspace{0.17em}4}^j}}\left(1\nu \right)^j\left(1+{\displaystyle \frac{j1}{2}}(1\nu )+\mathrm{}\right)`$ (15) which agrees perfectly with the large $`j`$ limit (using Stirling’s formula) of the exact result (2). As another test of our instanton formula (14) we can fix $`\nu `$ to any value (not necessarily near 1) and take $`j`$ large (in order to be in the semiclassical regime), and compare to the band-width obtained numerically from the two lowest eigenvalues of the matrix $`J_x^2+\nu J_y^2`$. The results are shown in Fig. 4, showing $`10\%`$ agreement for $`j10`$. In terms of the algebraic spectral program , we find that non-perturbative instantons play an interesting role in the semiclassical limit via eigenvalue differences for finite dimensional, but very large matrices. As a final calculation, we have computed the WKB approximation for the energy splitting, and we find that in the semiclassical limit $`\mathrm{\Delta }E^{\mathrm{WKB}}=\frac{2\mathrm{}}{\pi }\mathrm{exp}[\frac{1}{\mathrm{}g}_{\mathrm{TP}}𝑑\chi \sqrt{\mathrm{sn}^2(\chi )\mathrm{}g}]=\sqrt{\frac{e}{\pi }}\mathrm{\Delta }E^{\mathrm{inst}}`$, confirming that WKB gets the correct leading exponential but the WKB prefactor normalization is incorrect . To conclude, the instanton approximation gives the correct leading semiclassical result for the lowest band-width. But the Lamé model (1) has two independent parameters, $`\nu `$ and $`j`$, that allow us to probe separately the ‘far-separated-well’ and ‘deep-well’ limits respectively. The tight-binding approximation (something of a misnomer) is good for large period, for any well-depth; while the instanton approximation requires deep wells for a given period so that the combination $`\nu j(j+1)`$ is large. It would be interesting to study instanton - anti-instanton interactions and correlation functions in the Lamé system, and to compare to lattice simulations in the spirit of . Also, Lamé solitons with spatial profile of the form (8) will arise in the corresponding $`1+1`$ dimensional model. This work was supported in part by DOE grant DE-FG02-92ER40716.00.
no-problem/9906/astro-ph9906434.html
ar5iv
text
# Optical observations of the black hole candidate XTE J1550-564 during the September/October 1998 outburst ## 1 Introduction Low Mass X-ray Binaries (LMXBs) are systems formed by a low-mass companion and a compact object. A subclass of LMXBs are the Soft X-ray Transients (SXTs, so called X-ray Novae, although the physics is quite different from the classical novae). In these systems, sporadic outbursts are produced due to some poorly understood mechanism for which some mass is sporadically transferred onto the compact primary via an accretion disk. There are two types of SXTs. In Type I, the compact object is a weakly magnetized neutron star, whereas in Type II, the compact object is likely to be a black hole. Normally they brighten in the course of a few days to become one of the brightest sources in the X-ray sky, then declining in brightness over the next few months. The X-ray spectra are often dominated by an ultrasoft component and a hard X-ray tail. The most recent Type II SXT, XTE J1550-564 was first detected on Sep 7.09 UT by the All-Sky Monitor on the Rossi X-Ray Timing Explorer. First detection yielded an intensity of $``$ 70 mCrab (2-12 keV; 5-$`\sigma `$ significance) (Smith and Remillard 1998). The source showed a steady rise to 1.7 Crab on Sep 15. Thereafter, there was increased variability, with the intensity reaching 3.2 Crab on Sep 18.7 UT. On Sep 19 and 20, a large flare peaked at 6.8 Crab, and the intensity fell back to the range of 2.7-3.6 Crab on Sep 20 and 21 (Remillard et al. 1998). The source was also detected in hard X-rays (20-100 keV) by the Burst and Transient Source Experiment on the Compton Gamma-Ray Observatory (Wilson et al. 1998). An optical counterpart with V $``$ 16 mag. was proposed by Orosz et al. (1998) by means of V-band images obtained on Sep 8.99 UT. It is located at R.A. = 15<sup>h</sup>50<sup>m</sup>58$`\stackrel{s}{.}`$78, Decl. = $`56\mathrm{°}`$ $`28\mathrm{}`$ $`35\stackrel{}{.}0`$ (equinox 2000.0)( $`l^{II}`$ = 326.2$`\mathrm{°}`$, $`b^{II}`$ = -2.3$`\mathrm{°}`$). A likely radio counterpart to the X-ray transient, coincident with the optical one, was detected on Sep 9 and 10 (Campbell-Wilson et al. 1998). In this paper we present a series of optical observations of the XTE J1550-564 counterpart taken with the 0.9m Dutch telescope at the European Southern Observatory (ESO), La Silla between Sep 11 and Oct 23, 1998. ## 2 Observations Observations of the optical counterpart of XTE J1550-564 were carried out with the 0.9m Dutch telescope at the ESO La Silla Observatory. Observations were taken between Sep 10 and Oct 23, 1998. The apparent proximity of the source to the Sun after Oct 23 prevented additional optical observations. The CCD used was a TEK CCD (512 x 512 pixels) that yielded a $`3\stackrel{}{.}77`$ x $`3\stackrel{}{.}77`$ square field (see Figure 1). Typical exposure times were 600 s for the Johnson-U filter, 180 s for the Johnson-V filter, and 120 s for the Gunn-i filter per observing night. On Sep 11-16, a series of exposures were taken in the UVi-bands during 3 hr/night in order to search for short-term variations. The number of frames were 54 (U), 1 (B), 69 (V), 1(R) and 88 (i) during the observing period. The data were reduced using IRAF. The images were processed to eliminate the electronic bias and flat field corrected to remove the pixel-to-pixel sensitivity variations. The optical light curve was obtained when using the differences in magnitude between the object and 9 field stars, given on Table 1. The typical uncertainties for the quoted coordinates and UBVRi magnitudes are 1<sup>′′</sup> and 0.01 mag respectively. The magnitudes were calculated using the SExtractor software package (Bertin and Arnouts 1996). A 300-s spectrum of the optical counterpart (range 3500-7400 Å, resolution 2.0 Å) was obtained with the 3.6-m ESO telescope (equipped with EFOSC2 and a B300 grism) on Sep 15.98 UT. He-Ar lamps were used for wavelength calibration and the standard LTT 7379 for the flux calibration. The reduction of the spectroscopic data was also carried out with IRAF. We corrected the spectrum for interstellar reddening following Cardelli et al. (1989), using E(B-V) = 0.7, as discussed in the next section. ## 3 Results and discussion The first CCD images were taken on Sep 11 (MJD 51066), and showed XTE J1550-564 with magnitudes V = 16.66 $`\pm `$ 0.01, U = 18.14 $`\pm `$ 0.01 and i= 14.98 $`\pm `$ 0.01. The object was monitored on an almost daily basis in the UVi filters. The flux dropped by $``$ 1.5 mag during our observing period. It is striking that at the early epochs (MJD 50165 to 50175), the optical light curve seems to remain approximately on a constant level, with some stochastic flaring activity (Figure 2), whereas the X-ray spectrum is very hard, although is gradually softening. Figure 3 shows the complete optical light curve based on our measurements. X-ray light curves from RXTE and BATSE have been included in the figure for comparison. Just after the flaring activity, a large brightening, with an amplitude of 1 mag in the V-band is detected, coincident with the X-ray flare during which the V flux of the system reaches V = 15.6 mag. We do not find support for the delay of $``$ 1 day claimed by Jain et al. (1999). Unfortunately, we could not get observations in the U filter during the maximum, so we cannot confirm that this peak is also present in the U band. However, there is no signature of this peak in the i band. After this flaring episode, the light curve shows a constant decrease in brightness, that extends at least until our last data taken on MJD 51110. We measure the following decay rates: dU/dt= 0.041 $`\pm `$ 0.006 mag/day, dV/dt= 0.048 $`\pm `$ 0.004 mag/day and di/dt=0.048 $`\pm `$ 0.005 mag/day. We have searched for a superhump periodicity as seen in other SXTs (see Della Valle, Masetti and Bianchini 1998 and references therein) by means of Fourier analysis of the data obtained during Sep 11-16 following the prescription of Horne and Baliunas (1986) for unequally spaced time series. No convincing periodicity was found in the 0.05-2 day interval. The optical spectrum shown in Figure 4 shows a strong and broad H-$`\alpha `$ emission line, and broad and weaker emission lines from H-$`\beta `$ and He II (4686 Å), as initially reported by Castro-Tirado et al. (1998). He I (5876 Å) is strong and N III (4640 Å) is also marginally detected. Line characteristics are given in Table 2. These lines are typical for soft X-ray transients in outburst (Bradt and McClintock 1983). The emission lines and several interstellar absorption lines are superposed on a red continuum. The main interstellar features are the diffuse interstellar band at 4430 Å (EW = 1.9 Å), the blend (EW = 2.4 Å) due to the Na D lines at 5890 and 5898 Å and the blend due to the 6269 and 6282 Å lines (EW = 1.7 Å). The ‘warp’ feature around 5000 Å has been observed in other SXTs at maximum or during the decline phase (Della Valle et al. 1991, Bianchini et al. 1997, Masetti et al. 1997) where the Balmer lines (especially H<sub>β</sub>) were normally seen in emission filling in shallow absorptions. The colour excess can be estimated on several ways. From the equivalent width (EW) of the 4430 Å band (EW = 1.9 Å), we estimate a colour excess of E(B-V) = 0.80 (Herbig 1975). From the empirical relation between the width of the Na D lines given by Barbon et al. (1990) and considering EW = 2.4 Å for the Na D lines, E(B-V) = 0.60. Hereafter we will adopt $`<`$E(B-V)$`>`$ = 0.70 $`\pm `$ 0.10, i.e. A<sub>V</sub> = 2.2 following the relationship given by Savage and Mathis (1979). The distance to RXTE J1550-564 could be estimated on the basis of the linear relation between the equivalent width of the Na D lines and the distance (following Charles et al. 1989). We get D $``$ 2.5 kpc. From the work of Neckel and Klare (1980) it can be roughly seen how the absorption A<sub>V</sub> varies with the distance. On two fields 2$`\mathrm{°}`$ away from the RXTE J1550-564 position, A<sub>V</sub> apparently remains constant up to 2-3 kpc so no firm conclusion can be drawn. Hererafter we consider D $``$ 2.5 kpc, that is significantly different from D $``$ 6 kpc (Sobczak et al. 1999), a value obtained on the basis of the similarities of the optical and X-ray brightness of XTE J1550-564 with respect to the X-ray Nova Oph 1977. Given the large uncertainty in all the parameters, this distance estimate cannot be excluded. The peak luminosity during the flare on MJD 51075 is, following Sobczak et al. (1999), L = 2.0 $`\times `$ 10<sup>38</sup> (D/2.5 kpc)<sup>2</sup> erg s<sup>-1</sup>, which corresponds to the Eddington luminosity for M = 1.5 M at 2.5 kpc. The optical to X–ray ratio during the flare is L<sub>(V-band)</sub>/L$`_{(210\mathrm{keV})}`$ $``$ 2200, higher than the average value of $``$ 500 found by van Paradijs & McClintock (1995). For a quiescent magnitude of B $``$ 22 (Jain et al. 1999), i.e. B $``$ 19 after the dereddening correction and a distance D $``$ 2.5 kpc, we derive M<sub>B</sub> $``$ +7 for the progenitor, assuming that there is no contribution from the disk in quiescence. This value is consistent with the spectral type of a low-mass $``$ K0–K5 main-sequence companion (Allen 1976), similar to other SXTs for which radial velocities studies have been performed. The low-mass companion is also supported by the large magnitude range from quiescence to the outburst ($``$ 5 mag). ## 4 Conclusions Both the overall optical and X-ray light curves for XTE J1550-564 during the first two months since the onset of the source resemble the light curves of other type II SXTs (Tanaka and Shibazaki 1996). The spectrum shows emission lines arising from H-$`\alpha `$, H-$`\beta `$, He II and N III, typical of soft X-ray transients in outburst. From the interstellar absortion lines, we derive E(B-V) = 0.70 $`\pm `$ 0.10 and D $``$ 2.5 kpc, and M<sub>B</sub> $``$ +7 for the progenitor, which is consistent with a low-mass $``$ K0–K5 main-sequence companion. Only spectroscopic observations, to be performed when the system returns to quiescence, may lead to determine of the mass function of the system. This will make possible to discern whether the compact object in the XTE J1550-564 system is a neutron star or a black hole. ###### Acknowledgements. We thank the referee, R. A. Remillard, for useful suggestions. One of us (CSF) is very grateful to J. Gorosabel for his valuable help. This work has been partially supported by the spanish INTA grant Rafael Calvo Rodés and the spanish CICYT grant ESP95-0389-C02-02.
no-problem/9906/cond-mat9906387.html
ar5iv
text
# Classical orbit bifurcation and quantum interference in mesoscopic magnetoconductance ## Abstract We study the magnetoconductance of electrons through a mesoscopic channel with antidots. Through quantum interference effects, the conductance maxima as functions of the magnetic field strength and the antidot radius (regulated by the applied gate voltage) exhibit characteristic dislocations that have been observed experimentally. Using the semiclassical periodic orbit theory, we relate these dislocations directly to bifurcations of the leading classes of periodic orbits. TPR-99-07 Since it has become feasible to laterally confine a two-dimensional electron gas (2DEG) on length scales considerably smaller than the mean-free path of the electrons, the connection between classical and quantum mechanics has gained increasing renewed interest. Therefore, many experimental and theoretical investigations have recently been focused on the onset of quantum interference effects in mesoscopic ballistic devices. A well-adapted theoretical tool for this regime is the semiclassical approach that approximates quantum mechanics to leading orders in $`\mathrm{}`$. It is conceptually quite remarkable because it links quantum interference effects to purely classical phase-space dynamics. The so-called trace formula, originally developed by Gutzwiller for the density of states of a system with only isolated orbits in phase space, has been extended to systems with continuous symmetries (see Ref. for further literature) and for other physical properties such as conductance or magnetic susceptibility . Many quantum interference effects observed in mesoscopic systems could successfully be explained by the interference of a few classical periodic orbits, such as the Shubnikov-de-Haas oscillations of the free 2DEG , the magnetoconductance oscillations of a 2DEG in an antidot superlattice or a large circular quantum dot , or the current oscillations in a resonant tunneling diode (RTD) . (See Ref. for examples from nuclear and metal cluster physics.) Real physical systems are usually neither integrable nor fully chaotic, but exhibit mixed phase-space dynamics. Upon variation of an external parameter (e.g., deformation, magnetic field strength, or energy), bifurcations of periodic orbits typically occur, whereby new orbits are born and/or old orbits vanish. In the RTD , period-doubling bifurcations were found to be responsible for a period doubling in the oscillations of the observed I-V curves. In superdeformed nuclei and in the elliptic billiard , period-doubling bifurcations dominate the quantum shell structure locally through new-born orbit families whose amplitudes are of relative order $`1/\mathrm{}`$. Here we discuss a different mechanism through which orbit bifurcations manifest themselves in the magnetoconductance of a mesoscopic device, a narrow channel with central antidots. We present a semiclassical interpretation of dislocations in the conductance maxima as functions of antidot diameter and magnetic field strength and relate them to bifurcations of the leading classes of periodic orbits. We will show that their effect is neither due to their leading order in $`\mathrm{}`$ nor to period doubling, but to a subtle interference of different orbit generations with comparable periods. For the semiclassical description of the conductance we follow the approach of Refs. . The smooth part of the conductance $`G_{xx}`$ (in the direction $`x`$ of the electric current) can be described by the classical Kubo formula, whereas its oscillating part $`\delta G_{xx}`$ is approximated in terms of periodic orbits (po): $`\delta G_{xx}=`$ (2) $`{\displaystyle \frac{1}{\mathrm{}^2}}{\displaystyle \frac{4e^2}{h}}{\displaystyle \underset{\mathrm{po}}{}}𝒞_{xx}{\displaystyle \frac{R_{\mathrm{po}}(\tau _\beta )F_{\mathrm{po}}(\tau _s)}{|\mathrm{Det}(\stackrel{~}{\mathrm{M}}_{po}\mathrm{𝟏})|^{1/2}}}\mathrm{cos}\left({\displaystyle \frac{S_{po}}{\mathrm{}}}\mu _{po}{\displaystyle \frac{\pi }{2}}\right).`$ Here $`S_{po}`$ is the action (evaluated at the Fermi energy $`E_F`$), $`\mu _{po}`$ the Maslov index, and $`\stackrel{~}{\mathrm{M}}_{po}`$ the stability matrix of each periodic orbit . The temperature $`T`$ is included in the factor $`R_{\mathrm{po}}(\tau _\beta )=(T_{po}/\tau _\beta )/\mathrm{sinh}(T_{po}/\tau _\beta )`$ involving the (primitive) time period $`T_{po}`$ and the scattering time $`\tau _\beta =\mathrm{}/(\pi kT)2.410^{11}\mathrm{s}`$. Damping due to a finite mean free path is given by $`F_{\mathrm{po}}(\tau _s)=e^{T_{po}/(2\tau _s)}`$, where $`\tau _s=m^{}\mu /e3.810^{11}\mathrm{s}`$ is the scattering time extracted from the experimental mobility $`\mu `$. $`\mathrm{}1\mu `$m is the characteristic length of the active region, and $`𝒞_{xx}`$ is the velocity-velocity correlation function of the periodic orbit, defined by $$𝒞_{xx}=_0^{\mathrm{}}\text{d}te^{t/\tau _s}_0^{T_{po}}\text{d}\tau v_x(\tau )v_x(t+\tau ).$$ (3) Eq. (2), as well as the standard trace formulae for the density of states, diverges at bifurcation points where two (or more) stationary points of the action coalesce and the stationary-phase approximation in the trace integral leads to $`\mathrm{Det}(\stackrel{~}{\mathrm{M}}_{po}\mathrm{𝟏})=0`$. This can be locally overcome by expanding the action into higher-order normal forms . The simultaneous requirement of asymptotically reaching the Gutzwiller amplitudes far from the bifurcation points leads to uniform approximations which were developed systematically by Sieber and Schomerus . At the critical points the amplitudes are increased by a factor $`\mathrm{}^\delta `$, where $`\delta `$ depends on the type of the bifurcation. In the classical limit $`\mathrm{}/S0`$, bifurcations therefore may dominate the quantum oscillations. In real systems $`\mathrm{}/S`$ is, however, finite so that the other prefactors in the trace formula may compensate the factor $`(S/\mathrm{})^\delta `$ to a degree that depends on the specific system. Note that the uniform approximations of Refs. are restricted to isolated bifurcations; a general treatment of bifurcations of higher codimension (i.e., bifurcations of bifurcations) is still lacking . We employ a slightly modified version of the uniform treatment of Ref. , incorporating the discrete symmetries of the present system. To include effects of complex ‘ghost orbits’ not available in our calculations, we use the local approximation of their contribution derived from the numerical information at the bifurcation points. (For the technical details we refer to a forthcoming extended publication.) The device investigated here consists of electrostatic gates confining a high-mobility 2DEG in a GaAs/GaAlAs heterostructure (see Fig. 1). The 2DEG was 82nm beneath the surface, its electron density was $`3.4710^{15}\mathrm{m}^2`$, and the mobility about $`100\mathrm{m}^2V^1\mathrm{s}^1`$. Four metallic gates are used to define a long, narrow channel ($`5\mu `$m$`\times 1\mu `$m). These and two circular antidot gates are contacted individually. Details about the device are presented in and the references cited therein. All measurements were taken at $`T100`$ mK using standard low-excitation AC techniques. The dots in Fig. 2(a) show the experimental maximum positions of $`\delta G_{xx}`$ as functions of magnetic field $`B`$ and antidot gate voltage $`V_g`$ . The nearly equally spaced maxima and their shift to higher $`B`$ for decreasing antidot diameter can be understood in analogy to the Aharonov-Bohm (AB) effect, if the AB ring is identified with cyclotron orbits around the antidots. Extracting the effective area from the experimental data yields a diameter between $`0.76\mu `$m and $`0.86\mu `$m, which is consistent with the device dimensions. The dislocations of the peak positions (see the boxes in Fig. 2), however, cannot be understood within this simple picture. They have been qualitatively reproduced in a quantum calculation by Kirczenov et al. . Our objective is to decide if these dislocations and the variation of the spacings between the maxima can be understood semiclassically (which was doubted in Refs. ). For the effective one-electron model potential we follow essentially Kirczenov et al. who assumed a parabolic shape $`V(r)=E_F\left[r/a_0(1+s)\right]^2`$ for $`r<a_0(1+s)`$ and $`V(r)=0`$ otherwise. Here $`r`$ denotes the distance to the gate, and $`a_0`$ is the length scale over which the potential falls from $`E_F`$ to 0, i.e., the diffuseness of the potential. $`s`$ is a dimensionless parameter modeling the depletion width around the gates. We use $`a_0=0.05\mu m`$ and $`s=s_c=1`$ for the gates defining the channel throughout this paper. The depletion width $`s_d`$ of the antidot gates is varied between 1.5 and 2.2. According to Ref. , this corresponds to an effective antidot diameter of $`0.35\mu `$m to $`0.42\mu `$m. Following Eckhardt and Wintgen , we numerically integrate simultaneously the classical equations of motion and the reduced (2D) stability matrix $`\stackrel{~}{\mathrm{M}}`$. The orbits are converged to periodicity using the information provided by $`\stackrel{~}{\mathrm{M}}`$. They are followed through varying $`B`$ fields and antidot diameters using an adaptive extrapolation scheme. We find a large variety of distinct periodic orbits, many of them breaking the symmetry of the potential. Some typical examples are shown in Figs. 1 and 3. We have included over 60 orbits (not counting their symmetry-related partners). Their actions, velocity-velocity correlation functions and periods were evaluated numerically. The Maslov index was determined similarly to Ref. . Since the orbit bifurcations are of leading order in $`\mathrm{}`$, we now want to check if they have an increased influence on the amplitude of the conductance oscillations. In Fig. 3(a) we show the quantity Tr$`\stackrel{~}{\mathrm{M}}`$ of four typical periodic orbits (shown to the right) taking part in two successive bifurcations (where Tr$`\stackrel{~}{\mathrm{M}}=2`$) under variation of the magnetic field strength $`B`$. The left one is a tangent bifurcation, the right one a pitchfork bifurcation. In Fig. 3(b), the contribution of these orbits to the conductance is plotted. The dotted line gives the result of the trace formula Eq. (2). The amplitudes are diverging (arrows!) at the bifurcations. The uniform approximation (solid line) removes the divergences. Fig. 3(c) represents the corresponding data for a system scaled to have 10 times larger actions, thus being closer to the semiclassical limit. It is important now to note that the amplitudes in the uniform approximation are nearly constant over the bifurcations. We have thus shown that the bifurcations have no locally dominant influence on the conductance of the present system . Having established this result, we can further simplify our semiclassical treatment. Whereas for individual orbits a uniform treatment of the bifurcations is vital, their influence becomes smaller if a larger number of orbits is included. This is demonstrated in Fig. 4, where the total $`\delta G_{xx}`$ has been calculated for $`s_d=1.86`$ including all relevant ($``$ 60) periodic orbits. The solid line shows the result of the uniform approximation, whereas the dotted line corresponds to the standard Gutzwiller approach. To remove the spurious divergences in Eq. (2) (and those due to bifurcations of higher codimension in the uniform approach), we have additionally convoluted $`\delta G_{xx}`$ over the magnetic field $`B`$ (cf. Ref. ). The results are very similar . In particular, the maximum positions are practically identical. In the following, we therefore use simply Eq. (2) with an additional convolution over $`B`$. The semiclassical result for the maximum positions in $`\delta G_{xx}`$ is shown in Fig. 2(b). We do not obtain a detailed quantitative agreement with the experimental data, since no effort has been made to optimize the model potential. Qualitatively, however, all features of the observed phase plot in Fig. 2(a) are reproduced. The spacing of the maxima will be analyzed in our extended publication, where we also compare our results to those of quantum calculations, optimize the potential $`V(r)`$ and discuss the scaling properties of our results. Presently we want to concentrate on the dislocations (see the boxes in Fig. 2), which are clearly reproduced in our approach. The semiclassical description even reproduces quantitatively the local behavior at the dislocations. This is shown in Fig. 2(c) that corresponds to the heavy boxes in Figs. 2(a) and (b). The points give the experimental maximum positions; the lines correspond to the semiclassical results (with slightly shifted but unscaled values of $`s_d`$ and $`B`$). To understand the semiclassical origin of these dislocations, we consider for the moment a model system with only 7 closely related orbits. The inserts in Fig. 5(b) show Tr$`\stackrel{~}{\mathrm{M}}`$ of these orbits versus $`B`$ for two different antidot diameters. With decreasing $`s_d`$, new orbits are born at the bifurcations. We classify the orbits in a grandparent, a parent and a child generation, depending if they are offsprings of orbit 1, 2, or 3, respectively. All members within a generation behave nearly identically, thus justifying our classification. The contribution of the grandparent and the child orbits to the conductance is shown in Figs. 5(a) and (b), respectively. The behavior of each generation is in complete agreement with the simple AB picture discussed above, but the effective areas and their dependence on $`B`$ are different. The children have a larger semiclassical amplitude than the grandparents, thus dominating the conductance. Therefore the maxima of the total $`\delta G_{xx}`$ closely follow the children’s wishes where they exist, and the grandparents’ will otherwise, as becomes clear from Fig. 5(d). The parents’ influence was found to be negligible throughout. The different orbit generations lead to slopes and spacings of the maxima that do not match along the generation boundaries. This is the origin of the observed dislocations which occur, indeed, close to the bifurcation lines. In the full calculation with over 60 orbits, the various families with their bifurcation structures (gray lines in Fig. 2b) are superimposed. Only those dislocations survive for which the above model scenario is locally dominating and no further orbits interfere. As a result, some of the dislocations disappear, some are slightly shifted in the $`(s_d,B)`$ plane, and no unique one-to-one relation between dislocations and bifurcations can be established. Nevertheless, the qualitative pattern remains the same. In summary, our semiclassical description successfully reproduces all main features observed experimentally in the magnetoconductance of a mesoscopic channel with antidots. We have analyzed especially the dislocations of the conductance maxima as functions of magnetic field $`B`$ and antidot diameter $`s_d`$, and show that these are related to bifurcations of the leading classical periodic orbits of the system. The dislocations are due to the fact that the bifurcations define the border lines between regimes of different predominant orbit generations, leading to different dependences of the conductance maxima on $`B`$ and $`s_d`$. This induces the observed dislocations of the maximum positions, analogously to lattice defects at interfaces. As the classical dynamics are not affected by a rescaling of the system, the scaling behavior of the dislocations can be easily understood in the semiclassical approach. The main mechanism for generating the dislocations has been demonstrated for 7 model orbits forming three generations related through two bifurcations, whereby the middle generation is least influential. For individual orbits, a uniform semiclassical treatment of the bifurcations is essential. For the total contribution of over 60 orbits, some cancellations take place and a convolution of the trace formula (2) over $`B`$ was found to be sufficient. The ways in which the orbit bifurcations affect the quantum oscillations here is quite different from those reported in Refs. . There the relevant bifurcations lead to period doublings, whereas here the periods of all relevant orbits are approximately constant. Furthermore, in the RTD only a few orbits were found to be important, whereas the present system is dominated by a much larger number of orbits with nearly identical actions, periods and amplitudes. It is not a local enhancement of the amplitudes of isolated bifurcating orbits, but the occasional mismatch of the slowly varying contributions from competing orbit generations under the variation of the system parameters that causes the dislocations in the phase plots of the conductance maxima. We thank A. Sachraida, C. Gould and P. J. Kelly for providing us with the experimental data and helpful comments, and S. Tomsovic for a critical discussion.
no-problem/9906/cond-mat9906023.html
ar5iv
text
# Observation of a Nanoscale Metallic Dot Self-Consistently Coupled to a Two-Level System ## Abstract We have observed anomalous transport properties for a $`50`$ nm Bi dot in the Coulomb-blockade regime. Over a range of gate voltages, Coulomb blockade peaks are suppressed at low bias, and dramatic structure appears in the current at higher bias. We propose that the state of the dot is determined self-consistently with the state of a nearby two-level system (TLS) to which it is electrostatically coupled. As a gate voltage is swept, the ground state alternates between states of the TLS, leading to skipped Coulomb-blockade peaks at low bias. At a fixed gate voltage and high bias, transport may occur through a cascade of excited states connected by the dynamic switching of the TLS. With the Coulomb blockade in single-electron transistors and quantum dots now firmly established, efforts have begun to turn towards more complex systems in which these artificial atoms form the building blocks. Significant progress has been made in understanding linear-response transport in what may be termed “artificial molecules” in studies of double quantum dots, as well as “artificial solids” in studies of arrays of quantum dots. Moreover, such multi-dot systems have been considered as the basis for novel computer memory and logic elements. In these applications the transport properties must be considered well beyond the regime of linear-response. In this Letter, we examine the transport properties of a quantum dot coupled to a single two-level system (TLS) as a model for nonlinear transport phenomena in systems containing artificial atoms. We report a set of anomalous transport properties of a small Bi dot in the Coulomb-blockade regime. The anomalies are explained by the presence of a TLS in close proximity to the dot. While TLS’s have been studied as a source of $`1/f`$ noise in MOSFET’s, quantum point contacts, and single-electron transistors (SET’s), exhibited as random telegraph signals, the coupling in our devices leads to non-stochastic configurations of the TLS. At low bias, reversible switching of the TLS leads to skipped Coulomb blockade peaks in a manner similar to peak suppression in double quantum dots. At higher bias, a novel effect emerges in which transport occurs through a cascade of excited states which are connected by the dynamic switching of the TLS. Similar effects may be expected in multi-dot systems. The devices, shown schematically in the inset to Fig. 1, are SET’s in the standard double-junction geometry with a capacitively coupled gate. They were fabricated using e-beam lithography to make a shadow-mask for use with the self-aligned double-angle evaporation technique. The leads, consisting of Cu wires $`20`$ nm in diameter, were deposited first. They were then oxidized in-situ for 30 minutes at an oxygen pressure of $`1\times 10^4`$ Torr before deposition of the Bi dot. The resultant junction resistances of $`R_J1`$ M$`\mathrm{\Omega }`$ satisfy the condition for the Coulomb blockade $`R_J>R_q=h/4e^26.5\mathrm{k}\mathrm{\Omega }`$. Voltage-biased DC measurements were performed in an Oxford dilution refrigerator held at the base temperature of 70 mK. In the measured devices, the diameter of the dot is roughly $`50`$ nm. As the size of the Bi dot is smaller than a typical grain in a film of Bi grown on the same surface, the dot is likely a single crystal. Furthermore, while quench-condensed amorphous Bi is superconducting, crystalline Bi is not. Bulk crystalline Bi is a semimetal, with a low Fermi energy, $`E_F=27.2`$ meV, and a low electron density, $`n=2.7\times 10^{17}/\mathrm{cm}^3`$ , about $`10^5`$ times lower than a typical metal. A small overlap of bands at the Fermi level leads to an equal number of electrons and holes. The highly anisotropic Fermi surface results in Fermi wavelengths $`\lambda _F=`$14–215 nm. These wavelengths are comparable to the dimensions of the dot, and quantum properties might be expected. However, the devices reported here exhibited only metallic behavior. Experiments were also conducted using Al as the dot material. The five Al devices studied exhibited only the standard Coulomb-blockade behavior. The anomalous behavior presented here occurred in all seventeen of the Bi dots studied. The most striking feature of the data is seen in $`IV`$ curves presented as a function of both bias and gate voltages, $`V_b`$ and $`V_g`$, respectively. In a series of $`IV_b`$ curves taken over a full period of the Coulomb blockade in $`V_g`$, an anomalous gap remains where the blockade would normally vanish (Fig. 1). This can also be seen in $`IV_g`$ sweeps in the same region (Fig. 2(a)), where at low bias ($`V_bE_c/e`$, where $`E_c`$ is the charging energy of the dot), the conductance peaks disappear. The range of gate voltages in which the suppression occurred varied from device to device, as did the number of missing peaks, with a maximum of about five. Each device exhibited only one region of peak suppression within the range of gate voltages studied, which corresponded to changing the number of electrons on the dot by up to 100. At higher biases, where $`V_b`$ is greater than the anomalous gap, but still smaller than $`E_c/e`$, the missing peaks emerge and typically split, as seen in Fig. 2(a). At even higher biases, the peaks are suppressed within an envelope, and the lineshapes become complex. An example of this from a different device is shown in Fig. 3. A notable feature of this device is the noise in the peaks, visible even at the measurement bandwidth of 1 Hz. The noise has a broad spectrum, as evidenced in measurements with a bandwidth of 1 kHz (Fig. 3, left inset). Before introducing the TLS, we first consider other possible origins of these anomalies. A gap similar in appearance to Fig. 1 may arise in superconductors. This would require that both the dot and the leads are superconducting. While Bi in its amorphous state may be superconducting, the Cu leads are not. Moreover, a magnetic field up to 9 T does not remove the anomalous gaps. The opening of a gap at the band overlap in the Bi semimetal due to size quantization may also be considered, but such a bandgap would not manifest itself as in Fig. 1. The result of a bandgap $`E_g`$ in the Bi dot would be to increase the spacing in gate voltage $`\mathrm{\Delta }V_g`$ between the conductance peaks which occur at the band edges: $$\mathrm{\Delta }V_g=\frac{C_\mathrm{\Sigma }}{eC_g}(E_c+E_g)=\frac{e}{C_g}+\frac{C_\mathrm{\Sigma }}{C_g}\frac{E_g}{e},$$ (1) where $`E_c=e^2/C_\mathrm{\Sigma }`$ is the charging energy, and $`C_g`$ and $`C_\mathrm{\Sigma }`$ are, respectively, the capacitance of the dot to the gate and the total capacitance of the dot. Finally, we may exclude migrating charges as the data are highly reproducible. In what follows, we show that the transport anomalies may be explained by postulating a TLS coupled electrostatically to the Bi dot. We imagine the TLS to consist of a double-well potential (Fig. 1 inset) in which a charged particle is free to move. Electrons on the dot interact with the charge in the TLS. Hence a change in the number of electrons on the dot may change the relative energies of the two states of the TLS. In turn, the switching of the TLS will act as an effective change in gate voltage for the dot. Thus, the charge state of the dot must be determined self-consistently with the state of the TLS. To model in detail the behavior of a dot electrostatically coupled to a two-level system, we consider the following classical Hamiltonian: $$H^{\mathrm{ON}/\mathrm{OFF}}=\frac{1}{2C_\mathrm{\Sigma }}(QQ_0Q_0^{\mathrm{ON}/\mathrm{OFF}})^2+V_gQ_{\mathrm{TLS}}^{\mathrm{ON}/\mathrm{OFF}}+E_{\mathrm{TLS}}^{\mathrm{ON}/\mathrm{OFF}},$$ (2) where ON/OFF denotes the state of the TLS. Without the terms due to the two-level system, this is just the ordinary Hamiltonian of a metal dot. The actual integer charge on the dot is given by $`Q`$, while the optimal charge is controlled by the gate voltage $`Q_0V_gC_g`$. The electrostatic coupling between the TLS and the dot can be expressed by a shift $`Q_0^{\mathrm{ON}/\mathrm{OFF}}`$ of the optimal charge on the dot. The TLS is also electrostatically coupled to the gate, yielding the linear term $`V_gQ_{\mathrm{TLS}}^{\mathrm{ON}/\mathrm{OFF}}`$. Lastly, there may be an internal energy difference between the states of the TLS, represented by $`E_{\mathrm{TLS}}^{\mathrm{ON}/\mathrm{OFF}}`$. The model is also applicable to multilevel systems, in which case the TLS parameters can take on several discrete values. By an appropriate choice of $`Q_0`$, we may set the OFF-state parameters to be zero. This choice for the Hamiltonian leads to the two families of parabolas shown in Fig. 2(b). The horizontal offset, vertical offset, and slope of the ON-state parabolas are determined by $`Q_0^{\mathrm{ON}}`$, $`E_{\mathrm{TLS}}^{\mathrm{ON}}`$, and $`Q_{\mathrm{TLS}}^{\mathrm{ON}}`$, respectively. As shown in Fig. 2(b), as the gate voltage is increased, the lowest energy configuration may switch back and forth many times between the TLS OFF and ON states. This provides a natural explanation for the missing Coulomb-blockade peaks at low bias voltages in Fig. 2(a). For example, the first switching ON of the TLS occurs together with a switch from $`N`$ to $`N+1`$ electrons on the dot (first dark circle in Fig 2(b)). For this double switch to occur, either the electron number must first change with the TLS fixed, or the TLS must first switch with the electron number fixed. Either of these processes alone requires activation energy. Hence the double switch has an activation barrier and is too slow to produce a measurable current . This process preempts the usual Coulomb-blockade peak which would have occurred at slightly higher gate voltage. The peaks continue to be suppressed at low bias until increasing $`V_g`$ causes the TLS to switch permanently ON, at which point conductance peaks resume. The number of missing peaks can be estimated from Eq. (2) as $$\frac{e}{|Q_{\mathrm{TLS}}^{\mathrm{ON}}|}\mathrm{Mod}\left(\frac{|Q_0^{\mathrm{ON}}|}{e}\right)[1\mathrm{Mod}\left(\frac{|Q_0^{\mathrm{ON}}|}{e}\right)],$$ (3) such that a small slope $`|Q_{\mathrm{TLS}}^{\mathrm{ON}}|`$ can lead to a large number of missing peaks. To understand the more complex behavior at higher bias voltages, we have simulated transport through the dot using the orthodox Coulomb-blockade model, but including the TLS. We consider the transport current between a right lead with voltage $`V_b/2`$ and a left lead with voltage $`V_b/2`$, both leads coupled via equal junction resistances $`R`$ to the dot. Assuming rapid thermalization of electrons on the dot, we solve the rate equations for transitions between different charge states of the dot. However, we also allow transitions to occur between the two states of the TLS. The transition rate is $`\mathrm{\Gamma }_{\mathrm{TLS}}`$ if the overall energy is thereby lowered, and $`\mathrm{\Gamma }_{\mathrm{TLS}}\mathrm{exp}(\mathrm{\Delta }E/k_BT)`$ if the overall energy is raised by $`\mathrm{\Delta }E`$. As shown in Figs. 2(a) and 3, the simulated current is similar to that observed experimentally, remarkably so in the latter. In both cases, the fitting parameters are those of the TLS (Eq. 2). In Fig. 2(a), the best fit was obtained in the saturated limit of $`\mathrm{\Gamma }_{TLS}`$ much faster than the electron tunneling rate, but the opposite limit was obtained in Fig. 3. Thus there appears to be a considerable variation of TLS tunneling rates from device to device. The interactions between the dot and the TLS may lead to a novel transport process in which the TLS is active, switching dynamically. Consider the inset of Fig. 2(b), which shows the states of the system and the simulated current. There are four accessible states A-D ($`V_b>0.2E_c`$), with $`N`$ or $`N+1`$ electrons on the dot, and with the TLS either ON or OFF. At this gate voltage, the bias voltage is sufficient to add an electron to the dot, going from state A to B, with the TLS OFF. The extra electron may escape into the other lead, causing relaxation from B back to A. However, it is also possible for the TLS to spontaneously switch from OFF to ON, since $`\mathrm{\Delta }E`$ is negative for this process, thereby taking the system from state B to state C. Now, the system may switch repeatedly between states C and D resulting in electron transport, with the TLS staying in the ON state. Alternatively, when the system is in state D with $`N`$ electrons, the TLS may switch back to OFF without changing the number of electrons on the dot, and the system is back where it started in state A. This process can then begin again, each cycle transferring a minimum of two electrons through the dot. If the TLS transition rate $`\mathrm{\Gamma }_{\mathrm{TLS}}`$ is large compared with the tunnelling rate, then these cycles, where each succesive electron moving through the dot is accompanied by a switch of the TLS, will dominate the transport. If the TLS is slow, then the system will switch telegraphically back and forth between transport with the TLS ON or OFF, each process exhibiting a different conductance. This may explain the noise seen in the device of Fig. 3. While it is evident that the TLS is associated with Bi, since these effects were not observed in our Al dots, the microscopic origin of the TLS is unknown. It may reside on the surface of the dot in the Bi oxide layer, or it may be due to an ancillary grain of Bi from the double-angle evaporation. If it were a bistable defect, the associated small charge displacement, on the order of a lattice spacing, would be an insignificant effect in our device geometry. A larger effective charge displacement may come from a trapping state which is either filled or unfilled. The dynamics of such a charge trap may then be responsible for the broad range of lifetimes observed. The phenomenology of the dot-TLS system may be important in a number of applications. In quantum dots, a system consisting of two dots connected in parallel will mimic a dot coupled to a two-level system if one of the dots is electrically coupled to only one lead. With perhaps broader implications, a single-electron transistor used as an electrometer may behave in a way similar to the dot-TLS system. Such an electrometer, which may be used to detect charge configurations, interacts with the system being measured. For example, the Bi dot in our devices may be said to measure the state of the TLS, but clearly this process sometimes changes the state of the TLS. Therefore, in using an electrometer to probe a system of mobile charges, additional movement of the charges due to the measurement process must also be considered. In conclusion, we have observed anomalous transport characteristics in a $`50\mathrm{nm}`$ Bi dot, including suppression of Coulomb-blockade conductance peaks at low bias voltages, and splitting of these peaks at higher bias. This behavior may be ascribed to an electrostatic coupling of the dot to a TLS. The interaction with the dot causes the TLS to switch back and forth multiple times as the gate voltage is increased. Furthermore, sufficient bias may result in a transport process where the TLS switches dynamically with transport of each electron through the dot. A similar effect is expected in some double-dot systems. It may be informative to study the dynamics of such dot-TLS systems, particularly if the switching rate of the TLS can be controlled. Finally, more complicated dynamics may be expected in dots coupled to multilevel systems.
no-problem/9906/hep-ph9906328.html
ar5iv
text
# 1 Introduction ## 1 Introduction Low-energy-threshold underground detectors with good angular and recoil electron energy resolution open a new window to probe the structure of the weak interaction and neutrino electromagnetic properties as a good alternative to what can be learned at reactor and accelerator experiments . We propose to do this using a number of artificial neutrino and anti-neutrino sources such as $`{}_{}{}^{51}Cr_{24}`$ and $`{}_{}{}^{90}SrY`$. The neutrino flux is known to within a one percent accuracy, in contrast to the reactor case and one can reach lower neutrino energies. Non-standard neutrino properties have been studied for several years, partly motivated by the solar neutrino problem. A possible explanation of this problem is related to a large neutrino magnetic moment . Present constraints for the electron neutrino magnetic moment coming from reactor experiments gives $`\mu _\nu =1.8\times 10^{10}\mu _B`$ . The improvement of this bound in a new reactor experiment is the goal of the MUNU experiment, now running. The idea of using an artificial neutrino source (ANS) to search for a neutrino magnetic moment was first put forward by Vogel and Engel in Ref. . This kind of sources have already been used to calibrate both GALLEX and SAGE experiments and, recently, this idea has been considered by several experimental groups working in underground physics. The ANS have as an advantage that the uncertainties in the neutrino flux intensity are lower than in the case of reactor neutrinos and they have a small size, which makes them suitable for a deep underground experiment; they could be even surrounded by the detector as is the plan for the LAMA collaboration that has as a goal the use of a $`{}_{}{}^{147}Pm`$ anti-neutrino source with a one ton $`NaI`$ detector in order to test for a neutrino magnetic moment in the region $`10^{11}\mu _B<\mu _\nu <10^{10}\mu _B`$. The BOREXINO collaboration has also the possibility of searching for a neutrino magnetic moment in such a region using an ANS located at a distance of the order of 10 m . Here analyse the potential of these sources in testing the neutrino magnetic moment in a detector with both angular and recoil electron energy resolution. Such a study could be interesting for a detector like that in the HELLAZ proposal that is planning to detect neutrinos through neutrino electron scattering with an energy threshold as low as 100 KeV and with and angular resolution of 35 mrad. These two characteristics could make HELLAZ proposal adequate for improving the limits on neutrino properties by using artificial neutrino sources. The only limitation HELLAZ could have in comparison with BOREXINO is the large mass the last experiment is planning (100 tones vs. 6 tones) although this might be solve with an adequate experimental set up; for example, if it were possible to surround the source with the detector to get the full $`4\pi `$ neutrino source of the source, although in this case oscillation could not be studied. ## 2 Artificial Sources and Magnetic Moment We will consider the next sources for being the most realistic and interesting ones in getting results: * $`{}_{}{}^{51}Cr_{24}`$ neutrino source. This is a neutrino source that has already been used for calibrating both SAGE and GALLEX experiments . The main neutrino line is $`E_\nu =746`$ KeV and the lifetime is 40 days. * $`{}_{}{}^{49}V_{23}^{}`$ This is a neutrino source that produces neutrinos with energy $`E_\nu =602`$ KeV. The lifetime is 1.3 years . * $`{}_{}{}^{145}Sm_{62}`$ This is a neutrino source that produces neutrinos with energy $`E_\nu =554`$ KeV. The lifetime is 1.34 years . * $`{}_{}{}^{37}Ar_{18}`$ This is a neutrino source that produces neutrinos with energy $`E_\nu =814`$ KeV. The lifetime is 35 days . * $`{}_{}{}^{90}Sr`$ anti-neutrino source. This source has been studied in Ref. and its potential for the BOREXINO case has been already discussed . The neutrino energy spectrum for such a source is shown in Fig 1. The half-life is 28 years. For the case of neutrino sources the detectors which are now being proposed would be able to measure the differential cross section for the $`\nu _ee`$ scattering. At leading order in the SM this is given by $`{\displaystyle \frac{d\sigma ^W}{dT}}={\displaystyle \frac{2m_eG_F^2}{\pi }}\left\{g_L^2+g_R^2(1{\displaystyle \frac{T}{E_\nu }})^2{\displaystyle \frac{m_e}{E_\nu }}g_Rg_L{\displaystyle \frac{T}{E_\nu }}\right\}`$ (1) where $`T`$ is the recoil electron energy, and $`E_\nu `$ is the neutrino energy; $`g_L=\frac{1}{2}+sin^2\theta _W`$ and $`g_R=sin^2\theta _W`$. For the case of an anti-neutrino source the process will be $`\overline{\nu _e}e`$ scattering and we just need to exchange $`g_L`$ with $`g_R`$ in order to get the corresponding differential cross section. On the other hand the differential cross section in the case of a neutrino magnetic moment is $`{\displaystyle \frac{d\sigma ^{mm}}{dT}}={\displaystyle \frac{\pi \alpha ^2\mu _\nu ^2}{m_e^2}}\left\{{\displaystyle \frac{1}{T}}{\displaystyle \frac{1}{E_\nu }}\right\}`$ (2) which adds incoherently to the weak cross section (neglecting neutrino mass). It is well known from this equation that for lower values of T the neutrino magnetic moment signal will drastically increase. This makes interesting the use of low threshold detectors such as the one has been proposed by the HELLAZ collaboration, and which could reach a threshold as low as 100 KeV. Besides the low energy threshold, HELLAZ will also have angular resolution. This could be useful not only to lower the systematic errors, but also to take advantage of the best regions in the ($`\theta `$,$`T`$) plane where the non-standard effects could be bigger. Although the restriction to a narrow window will limit the statistics, the enhancement of the neutrino magnetic moment (NMM) effect may over-compensate and one might have an overall gain. In it was shown that $`\frac{d\sigma ^W}{dT}`$ vanishes for forward electrons (which implies maximum recoil energy for $`e^{}`$) for a $`\overline{\nu }_e`$ energy given by: $$E_\nu =m_e\frac{g_Lg_R}{2g_R}=m_e/4sin^2\theta _W0.548MeV$$ (3) This kind of cancellation only takes place when considering scattering of $`\overline{\nu }_e`$ off $`e^{}`$ Of course, to be able to study this effect, experiments capable of measuring both the recoil (kinetic) energy of the electron (T) and its recoil angle ($`\theta `$) become necessary, so that we can select neutrino energies $`E_\nu `$ from a non-monochromatic source (a monochromatic source of $`\overline{\nu }_e`$ with $`E_\nu =m_e/4sin^2\theta _W`$ would be the ideal but there are not monochromatic anti-neutrino sources). The three variables $`E_\nu `$, $`T`$ and $`\theta `$ are related by the equation: $$cos\theta =\frac{T}{\sqrt{T^2+2m_eT}}\left(1+\frac{m_e}{E_\nu }\right)$$ (4) As discussed in the dynamical zero seems potentially interesting to measure $`\mu _\nu `$ since it opens a window in phase space where the weak cross section becomes small, so that the magnetic moment contribution could eventually become larger than the weak cross section. However, as discussed in in the context of reactor neutrino experiments, the fact that the statistics close to the dynamical zero is poor is, unfortunately, more important than the enhancement in $`d\sigma ^{mm}/d\sigma ^W`$. ### 2.1 Neutrino sources We will consider first the case of a $`Cr`$ neutrino source. As mentioned in the introduction this source has already been used for the calibration of SAGE and GALLEX experiments. We will consider the 746 KeV neutrino line that has the 81 % of the neutrino flux. For this case the differential cross section will be as given in Eq. (1) and we just need to substitute the corresponding value for the neutrino energy. The result is shown in Fig 2 both for the Standard Model case as well as for the case of a neutrino magnetic moment. Different values of the magnetic moment are shown. As we can see, for low values of $`T`$ the NMM signal is of the same order of the SM one for $`\mu _\nu .61\times 10^{10}\mu _B`$. The differential cross section $$\frac{d\sigma ^W}{dcos\theta }=\frac{d\sigma ^W}{dT}\frac{dT}{dcos\theta }$$ (5) can be easily obtained and it is shown in Fig. 3, where we can see a similar result: there is a region, for large electron recoil angle, for which the NMM signal is comparable to that of the SM. The similarities of these two figures are more evident if one notices that the recoil angle $`\theta `$ is maximum for lower T, as can be derived from Eq. (4). For other neutrino sources different that the $`Cr`$ source, the shape of the plots shown in Fig. 2 and Fig. 3 will have a slightly change, due to the fact that the neutrino energy is different. However the qualitatively result will be the same. ### 2.2 Anti-neutrino sources Now we consider the case of a $`{}_{}{}^{90}Sr^{90}Y`$ anti-neutrino source. This source has been studied by a Moscow group and its potential has been studied for the BOREXINO case . We can make a similar analysis as in the case of the neutrino sources in the previous section. The main difference here, beside the interchange of $`g_L`$ by $`g_R`$ in the differential cross section, is that we have an energy spectrum instead of a neutrino energy line. Therefore, we need integrate over the neutrino energy distribution. As we are interested in the angular distribution, it is more convenient to use Eq. (4) to make a change of variables in the integration. This has also been done in . The result can be express as $`{\displaystyle \frac{d\sigma }{dT}}_{E_\nu }={\displaystyle \frac{d^2\sigma }{dTd(cos\theta )}d(cos\theta )}={\displaystyle \mathrm{\Theta }_{\text{p. s.}}f(T,\theta )\frac{d\sigma (T,\theta )}{dT}\frac{m_epT}{(pcos\theta T)^2}d(cos\theta )}`$ (6) or, if we are interested in the angular distribution $`{\displaystyle \frac{d\sigma }{d(cos\theta )}}_{E_\nu }={\displaystyle \frac{d^2\sigma }{dTd(cos\theta )}𝑑T}={\displaystyle \mathrm{\Theta }_{\text{p. s.}}f(T,\theta )\frac{d\sigma (T,\theta )}{dT}\frac{m_epT}{(pcos\theta T)^2}𝑑T}`$ (7) In this equations $`\mathrm{\Theta }_{\text{p. s.}}`$ accounts for the allowed phase space and $`f(T,\theta )=f(E_\nu (T,\theta ))dn/dE_\nu `$ is the neutrino energy spectrum as a function of $`T`$ and $`\theta `$. We have computed the differential cross sections $`d\sigma /dT`$ and $`d\sigma /dcos\theta `$ for an anti-neutrino energy spectrum given as shown in Fig. 1, normalized to one. In the case of $`d\sigma /dT`$ we have integrated Eq. (6) in the whole allowed $`\theta `$ range. For the case of the differential cross section $`d\sigma /dcos\theta `$ we have integrated $`T`$ in the range $`.1MeV<T<0.5MeV`$, the energy range to which HELLAZ could be sensitive. The results are shown in Fig. 4 and Fig 5 both for the Standard Model case as well as for the case of a neutrino magnetic moment. Different values of the magnetic moment are shown in these figures. The kink in the distribution in Fig. 5 is due to the sharp decrease in the energy spectrum for $`E_\nu >0.5MeV`$ (fig. 1). Notice that, besides there is an additional contribution to the differential cross section, the shape is also different from that of the standard model. In particular the magnetic moment contribution is slightly bigger than the Standard Model one both for small angles and for big angles, meanwhile, in the intermediate region, the SM is bigger. However, the low values of the differential cross section in the small angle region, in comparison with other angles could make the analysis of this region more difficult. The case of the large angle region can be easily explained if we consider that, for minimum recoil electron energy the recoil angle is maximum, therefore, as the magnetic moment contribution increases at low $`T`$, it is natural to have a similar effect for large $`\theta `$. Besides the advantages that ANS have in general, this particular source seems to be interesting because the energy range of the spectrum that belongs to the $`{}_{}{}^{90}Sr`$ has a peak in an energy range that is close to the kinematical zero that has been already discussed in the previous section. In order to illustrate the potential of this energy region we show in Fig. 6 the result that would be obtained for the ideal hypothetical case of a pure $`{}_{}{}^{90}Sr`$ without any contaminant. ## 3 The NMM Signal and Recoil Angle Resolution We now come back to the real $`SrY`$ source. In this case one could try to optimise the best region in the $`\theta T`$ plane on which the non-standard effect is maximum. In order to do this analysis let us consider the curves in the $`(T,\theta )`$ plane given by the condition $$C=\frac{d\sigma ^{mm}/dT}{d\sigma ^W/dT}$$ (8) This gives us the curves shown in Fig. 7. These are characterized by a given ratio of the magnetic moment differential cross section to the SM one. Therefore, for C=1 we will get the curve where the magnetic moment signal is equal than the SM one, for C=2 the the magnetic moment signal is twice the SM one, and so on. The corresponding iso-curves are given in figure 7, for $`\mu _\nu =10^{10}\mu _B`$. In the figures the curves for $`c=1,2,4,8,32`$ are shown; of course, for different selection of magnetic moments, the values of $`c`$ in the same figure are scaled. For instance, taking $`\mu _\nu =10^{11}\mu _B`$, the curves shown in Fig. 4 would correspond to $`c=0.01,0.02,0.04,0.08,0.32`$. It is also important to notice that, for any couple ($`\theta `$,$`T`$) the neutrino energy is already fixed by the kinematics as can be seen from Eq. (4). The effect of the dynamical zero on the iso-curves can be noticed specially in the cases of ratios $`c=32,64`$, where curves surrounding the position of the dynamical zero appear. This effect does not appear in the case of a neutrino source as can be seen in Fig. 8 where similar curves are shown for the case of $`\nu _ee`$ scattering. Given that the iso-curves (Fig. 7) reflect the presence of a favoured region for searching for a magnetic moment, thanks to the dynamical zero, it seems interesting to integrate the cross section over regions in the $`(T,\theta )`$ plane limited by the iso-curves. In this way, we are optimising the region of integration to look for magnetic moment. Figure 9 shows the result of integrating the differential cross section given in Eq. (6) over $`T`$ and $`\theta `$ in regions such that $`d\sigma ^{mm}/d\sigma ^W>C`$. A neutrino magnetic moment $`\mu _\nu =6\times 10^{11}\mu _B`$ has been assumed. Of course, as the limiting ratio $`C`$ is taken larger, the magnetic moment signal becomes larger than the SM one. However, the integral in this case is small. Note that if one integrates over the whole region (C=0) one can probe the complete cross section, but the value of NMM relative to SM decreases. Therefore it is interesting to study intermediate regions such as the region limited by $`C=0.7`$, where the contribution of the neutrino magnetic moment is comparable with that of the weak interaction, although the statistics is a 30 % of the total one. ## 4 Discussion & Conclusions We have discussed the potential of investigating neutrino-electron scattering as a probe of neutrino magnetic moment of the order $`\mu _\nu 10^{11}\mu _B`$ at low-threshold underground detectors with good angular and recoil electron energy resolution. We propose to do this using a number of artificial neutrino and anti-neutrino sources such as $`{}_{}{}^{51}Cr_{24}`$ and $`{}_{}{}^{90}SrY`$. The neutrino flux is known to within a one percent accuracy, in contrast to the reactor case and one can reach lower neutrino energies. For the $`{}_{}{}^{90}SrY`$ source we have investigated the possible role of dynamical zeros in improving the sensitivity to the neutrino magnetic moment, with a negative result due to the poor statistics in this region. However, integrating over large kinematical regions, we estimated that the signal expected for a neutrino magnetic moment of $`\mu _\nu =6\times 10^{11}\mu _B`$ will be comparable to that expected in the Standard Model and corresponds to a 30% enhancement in the total number of expected events. In order provide a more reliable estimate of the sensitivities to the neutrino magnetic moment that can be reached in this kind of studies a dedicated experimental analysis will be necessary. ## Acknowledgements This work was supported by DGICYT grant PB95-1077, by Intas Project 96-0659 and by the EEC under the TMR contract ERBFMRX-CT96-0090. OGM was supported by SNI-México and by the grant CINVESTAV JIRA ’99/08. VBS was supported by the RFFR grant 97-02-16501 and by Generalitat valenciana. We thank Tom Ypsilantis for discussions.
no-problem/9906/cond-mat9906102.html
ar5iv
text
# Monte Carlo Simulation of Smectic Liquid Crystals and the Electroclinic Effect: the Role of the Molecular Shape ## I Introduction The response of smectic liquid crystals to applied electric fields has been extensively studied for both basic research and applications. One subject of particular interest is the electroclinic effect, which occurs in the smectic-A (SmA) phase of chiral molecules. In the electroclinic effect, an applied electric field in the smectic layer plane induces a tilt of the molecules relative to the layer normal, in a direction orthogonal to the field. The magnitude of the induced tilt scales linearly with the applied electric field for low fields, and then saturates at higher fields. This effect was predicted by Meyer on the basis of symmetry , and it was subsequently observed experimentally by Garoff and Meyer . It is now being exploited for electro-optic devices that display a continuous gray scale as a function of applied electric field, such as spatial light modulators . To optimize electroclinic liquid crystals for device development, one needs a theoretical understanding of how the electroclinic tilt depends on electric field, temperature, and molecular structure. So far, most theoretical work on the electroclinic effect has been through Landau theory, i.e. a minimization of the free energy expanded in powers of the molecular tilt and the electrostatic polarization . This work explains certain aspects of the electroclinic effect—in particular, it shows how the tilt and polarization depend on field for low fields, and it shows how the susceptibility to a field increases as the system approaches the second-order phase transition from the SmA to the smectic-C (SmC) phase. However, some important questions about the electroclinic effect are not addressed by Landau theory. The first and most general question is: How sensitive is the electroclinic effect to molecular shape? In other words, how much does the electroclinic susceptibility change with slight details of molecular structure? A second and more specific question is: How does the applied electric field change the distribution of molecular orientations? Does it make the molecules tilt as rigid rods from an initially untilted state to a tilted state? Or does it change a state of disordered tilt in random directions into a state of ordered tilt in one direction? The latter alternative is suggested by the de Vries description of the SmA phase . To address these questions, in this paper we present a series of Monte Carlo simulations of smectic liquid crystals. Simulation is an appropriate tool to address these questions for two reasons. First, in simulations we can begin with a microscopic model for the molecular structure and determine the large-scale order of the liquid-crystal system as a function of thermodynamic variables such as temperature, density, and applied field. We can then make small changes in the molecular shape and see how these changes affect the large-scale order of the system. Thus, we can determine how macroscopic properties such as the electroclinic susceptibility depend on details of the molecular shape. Second, in simulations we can take snapshots of the positions and orientations of all the molecules in the system, and hence can extract any correlation function to characterize the system. This information is not available in Landau theory, and is generally difficult to extract from experiments. Hence, simulations give us new information about the distribution of molecular orientations as a function of electric field, and about topological defects in the molecular orientations. In these simulations, we use a “bent-rod” rigid molecule with the oblique shape shown in Fig. 1. This shape is motivated by three considerations. First, the three-dimensional structure of many liquid-crystal molecules, such as the homologous series KNnm, has this general shape . In the center is a rigid molecular core, which defines the optical axis of the molecule, and on both ends are hydrocarbon chains, which extend out at an angle from the core. In the homologous series KNnm, the electroclinic tilt angle of the SmA phase can be increased by making the hydrocarbon chains longer, thus making the molecules more oblique. Second, the Boulder model for ferroelectric liquid crystals shows that molecules in the SmC phase typically take the shape of bent cylinders . For that reason, we can regard this shape as a generic feature of smectic liquid crystals. Third, density functional theory has been used to predict the phase diagram of parallel offset hard cylinders, a shape similar to bent rods . That work showed a high-density SmC phase for molecules with a higher offset ratio, i.e. the more oblique. These results confirm that the obliqueness of molecular shape is an important parameter to determine the phase behavior of smectic liquid crystals. To simulate a simple molecular structure with a bent-rod shape of variable obliqueness, we use a molecule composed of seven spheres arranged in the shape of the letter Z, as illustrated in Fig. 1. The spheres are “glued” rigidly together with no intra-molecular degrees of freedom, with a bend angle $`\theta `$ between the core and tail portions of the molecule. We consider the cases $`\theta =45^{}`$, which is quite oblique, and $`\theta =5^{}`$, which approaches the rod-shaped limit of $`\theta =0^{}`$. Each molecule also has a dipole moment that lies perpendicular to the molecular backbone, as shown, giving the molecule a chiral structure. The molecules interact through a soft repulsive sphere-sphere pair potential, and each molecular dipole interacts with the applied electric field. We neglect dipole-dipole interactions as an approximation to simplify the computations. These simulations provide clear evidence that steric repulsion alone can give rise to order in the molecular tilt, even without including intermolecular dipole-dipole interactions. Furthermore, they show that the bend angle $`\theta `$ plays a major role in determining phase behavior. For $`\theta =45^{}`$ the system has a phase transition directly from the isotropic phase to the SmC phase. By contrast, for $`\theta =5^{}`$, the system has nematic and SmA phases, each stable over a wide range of temperature. In the absence of an applied electric field, the molecules of the SmA phase are not aligned with the layer normal but rather are tilted in random directions, and the orientation of the tilt exhibits vortex-like point defects. When an electric field is applied, the magnitude of the molecular tilt increases and the direction of the tilt becomes more ordered, giving a strong electroclinic effect. At high fields, the electroclinic tilt angle saturates at approximately $`19^{}`$. The simulations also show that a high electric field applied to the nematic phase induces a transition into the SmA phase, showing another ordering effect of the field. The plan of this paper is as follows. In Sec. II we describe the details of the model and the computational method that we used. In Sec. III we present the results of the simulations for bend angle $`\theta =45^{}`$ and $`\theta =5^{}`$. In particular, we show the electroclinic effect in the SmA phase for $`\theta =5^{}`$. Finally, in Sec. IV, we discuss the significance of these results for experiments on smectic liquid crystals. ## II Simulation Method In our simulations, we consider molecules composed of seven soft spheres arranged in the rigid bent structure shown in Fig. 1. The molecular director is defined as the unit vector along the five-sphere core of the molecule. The interaction between molecules is reduced to an interaction between different spheres in different molecules. Intramolecular interactions and degrees of freedom are suppressed. The sphere-sphere interaction potential is the truncated Lennard-Jones potential, also known as the Weeks-Chandler-Anderson potential , cut off at its minimum so there is no attractive tail: $$U_{mn}^{\mathrm{int}}=\{\begin{array}{cc}4ϵ\left[\left(\frac{\sigma }{r_{mn}}\right)^{12}\left(\frac{\sigma }{r_{mn}}\right)^6\right]+ϵ,\hfill & \text{if }r_{mn}r_c=2^{1/6}\sigma \text{;}\hfill \\ 0,\hfill & \text{otherwise.}\hfill \end{array}$$ (1) where $`r_{mn}=|\stackrel{}{r}_m\stackrel{}{r}_n|`$ and $`m`$ and $`n`$ are the sphere indices in different molecules. We choose this short-range repulsive interaction to reduce required computation time and to focus on the role of steric effects without any contribution from attractive interactions. For the rest of this paper, we measure lengths in units of $`\sigma `$ and energies in units of $`ϵ`$. In addition, each molecule interacts with the applied electric field $`\stackrel{}{E}`$ through the coupling $$U_j^{\mathrm{dipole}}=\stackrel{}{E}\stackrel{}{p}_j,$$ (2) where $`\stackrel{}{p}_j`$ is the dipole moment of molecule $`j`$. The molecular dipole moment is defined to have unit magnitude, which gives a scale for the electric field. We simulate 500 molecules in a flexible three-dimensional box with periodic boundary conditions. We keep the system with constant volume density 0.75 Lennard-Jones particles per unit volume, and allow the aspect ratio of the simulation cell to adjust according to the Metropolis algorithm. We do not allow the cell to shear. The system is prepared by a procedure analogous to the experimental technique of cooling in a strong aligning field to avoid the formation of smectic domains. We begin the simulations at the high temperature $`k_BT=20.0`$, with the box size of $`11.5\times 11.5\times 35.0`$. This aspect ratio favors the formation of a five-layered smectic phase. In the initial state, the molecules have random positions but all the molecules are “double-aligned,” that is, both the directors and the dipole moments are aligned. During the preliminary cooling procedure, we suppress all orientational degrees of freedom and allow the molecules to diffuse while remaining double-aligned. The temperature of the system is reduced slowly at a rate $`10^4`$ per Monte Carlo step. The system comes to equilibrium quickly. In about 10,000 Monte Carlo steps, the molecules form five distinct layers. If the layer normal is not parallel to the $`z`$ axis, we measure the angle away from the $`z`$ axis, adjust the director of the molecules, choose a new random initial configuration, and repeat the simulation to get a layered system with the layer normal along the $`z`$ axis, with no defects in the layer structure. Once we reach this double-aligned smectic state, we reduce the temperature to about 1.5, still in the double-aligned state. Then, after the system is in equilibrium, we switch on the three rotational degrees of freedom for each molecule and equilibrate for an additional 100,000 Monte Carlo steps per particle. In one Monte Carlo step, each randomly selected molecule attempts three translations and three rotations. To characterize the phase behavior of the system, we particularly use three order parameters. First, the nematic order tensor $`Q`$ represents the strength and direction of orientational order of the molecules. It is defined as $$Q_{\alpha \beta }=\frac{1}{N}\underset{j=1}{\overset{N}{}}\left(\frac{3}{2}n_{j\alpha }n_{j\beta }\frac{1}{2}\delta _{\alpha \beta }\right),$$ (3) where $`\stackrel{}{n}_j`$ is the director along the core of molecule $`j`$ and $`N=500`$ is the number of molecules. The eigenvector corresponding to the maximum eigenvalue of $`Q`$ is the average director of the system. If the eigenvalue is 1, the molecular directors are completely aligned; if the value is lower it reflects less perfect alignment. Second, the polarization $`\stackrel{}{P}`$ represents the degree of orientation of the molecular dipole moments. It is defined as the vector average $$\stackrel{}{P}=\frac{1}{N}\underset{j=1}{\overset{N}{}}\stackrel{}{p}_j.$$ (4) Third, the smectic order parameter $`\sigma `$ represents the strength of the density modulation along the $`z`$ direction. It is defined as $$\sigma =\frac{1}{N}\underset{j=1}{\overset{N}{}}e^{2\pi iz_j/d},$$ (5) where $`z_j`$ is the $`z`$-coordinate of the center of mass of molecule $`j`$ and $`d`$ is the smectic layer wavelength, which is one-fifth of the $`z`$-dimension of the simulation cell. ## III Results ### A Forty-Five Degree Bent-Rod Molecules For the molecules with the large bend angle $`\theta =45^{}`$, the phase sequence is crystal-SmC-isotropic. The SmC phase is stable over a wide range of temperature, from approximately $`k_BT=0.5`$ to 1.5 (in Lennard-Jones units). A sample configuration is shown in Fig. 2. The polarization of any single layer, defined by Eq. (4), is $`P=0.85`$, indicating nearly perfect orientational order. In spite of the high orientational order within each layer, the local tilt direction of each layer is only loosely coupled to that of adjacent layers, and it tends to wander. This type of behavior was evident also in the simulation study of Affouard et al. on a related system. It is similar to the proposed random smectic-C<sub>R</sub> phase, which has been suggested as a model for the thresholdless switching observed experimentally in certain smectic liquid crystals . An alternative model has recently been proposed for these experiments , but the smectic-C<sub>R</sub> phase remains a theoretical possibility for future materials. Indeed, this proposed phase with random orientations of adjacent layers can be viewed as one version of the sliding phase that has been investigated in recent theoretical work . One possible explanation for the low interlayer correlations in our simulations is that there is very little interaction between the tilt directions in adjacent layers, because the intermolecular potential is purely repulsive and because there is hardly any interdigitation between the layers. As a result, the adjacent layers should have very little preference for synclinic (ferroelectric) or anticlinic (antiferroelectric) order, and they should be fairly free to wander between these extremes. An alternative explanation is that the layers might prefer anticlinic order, but they are frustrated because the system has an odd number of layers (five). It is interesting to note that Affouard’s simulation also included an odd number of layers (three). This latter explanation seems less likely, however, because the interaction between layers does not seem to favor anticlinic order. Note that we observe the SmC phase even though we have not included dipole-dipole interactions in our intermolecular potential, indicating that steric repulsion defined by molecular shape is sufficient to produce order in the molecular tilt direction. Electrostatic interactions are not required to produce a tilted smectic . Presumably the inclusion of dipole-dipole interactions in our simulation would increase the temperature range over which the SmC is stable, and it would likely increase the coupling between the tilt directions in adjacent layers. ### B Five Degree Bent-Rod Molecules The molecule with bend angle $`\theta =5^{}`$ looks very nearly like a rod, but it has properties quite different from a purely rod-shaped molecule. The $`\theta =5^{}`$ system has a stable SmA phase over a temperature range of $`k_BT=0.7`$ to 3.0. A sample configuration of the SmA phase is shown in Fig. 3a. The smectic order parameter defined by Eq. (5) is very high, about 0.9. The molecules in each layer of the SmA phase have approximately zero average tilt and no net polarization. However, a close look at the structure of an individual layer shows that the local molecular tilt is nonzero but that defects cause the net tilt to vanish, as shown in Fig 3b. In some configurations, these point defects in the local tilt appear to be vortices analogous to those seen in, for example, an $`xy`$ model . Comparison of defect structures in adjacent layers shows that there is no strong correlation in defect location between layers, indicating that these defects are truly point vortices and do not thread through all five layers of the system. In this respect, they are analogous to the “pancake” vortices seen in layered superconductors with weak interlayer coupling . When we apply an electric field in the SmA layer plane, the molecules tilt showing a clear electroclinic effect. The observed polarization responds rapidly to the applied field, coming close to its equilibrium value in only several thousand Monte Carlo steps, while the tilt angle takes up to 500,000 Monte Carlo steps to equilibrate. This equilibration would likely have been faster if we had implemented degrees of freedom that allowed shear deformation of the simulation cell, but clearly it is much longer than the equilibration time for the polarization. Figure 4a shows the SmA phase under a strong applied field $`E=10`$. When we examine a layer from this system, we observe that the vortex-like defects have vanished, and the molecules in the same layer are all closely aligned, as shown in Fig. 4b. We can compare the measured polarization response to the applied field with the prediction of a simple spin model. The molecules in a smectic layer are localized with directors pointing in almost the same direction. The most active movement is the rotation of the molecular dipole moment around the director. In view of this property, we consider the molecules as two-dimensional independent dipoles with only one effective rotational degree of freedom. The net polarization can then be written as $$P=\frac{_0^{2\pi }p\mathrm{cos}\theta e^{Ep\mathrm{cos}\theta /k_BT}d\theta }{_0^{2\pi }e^{Ep\mathrm{cos}\theta /k_BT}𝑑\theta }=\frac{I_1(Ep/k_BT)}{I_0(Ep/k_BT)}$$ (6) where $`I_0`$ and $`I_1`$ are modified Bessel functions and $`p=1`$ is the magnitude of the dipole moment of a single molecule. The simulation results for polarization vs. field are plotted together with the analytic prediction in Fig. 5a, and are in close agreement. Indeed, the agreement is much closer than one would expect from such a simple model—one would expect the interactions among the molecules to give collective order that would give a higher initial slope to the the polarization vs. field curve. This agreement shows that the alignment of molecular dipole moments with the electric field is a single-molecule effect rather than a collective effect for these nearly rod-like molecules. Collective effects should become more important if we increase the interaction between the directions of the molecular dipoles—either by including dipole-dipole interactions in our simulation model or by increasing the bend angle $`\theta `$ to make the molecules more oblique. In addition to these results for the polarization, we also measure the molecular tilt angle in the simulations. The average tilt angle for the system is extracted from the nematic order tensor $`Q`$ defined by Eq. (3). From the eigenvector corresponding to the maximum eigenvalue of $`Q`$, we can calculate the tilt angle away from the layer normal. Using this technique, we measure the tilt angle as a function of applied electric field in the simulation for two temperatures, $`k_BT=0.7`$ and $`k_BT=1.3`$. The results are shown in Fig. 5b. We observe that the tilt angle responds more sharply to the applied field at lower temperature than at higher temperature; that is, the electroclinic coefficient drops with increasing temperature. This temperature dependence is similar to the temperature dependence of the electric susceptibility shown in Fig. 5a. At high applied field, the tilt angle saturates at about $`19^{}`$ for both temperatures. We note that the tilt angle, in contrast with the polarization, is a collective effect rather than a single-molecule effect in this simulation, as shown by the much longer equilibration time for the tilt angle. Thus, the saturated tilt angle of $`19^{}`$ is not simply related to the molecular geometry, but depends on the collective order of many molecules whose transverse dipoles have been aligned by the applied electric field. When the molecules tilt under an electric field, the thickness of the smectic layers shrinks. Because the simulation cell is flexible, the $`z`$-dimension of the cell also shrinks. At the temperature $`k_BT=0.7`$, the $`z`$-dimension of the cell changes from 36.7 at $`E=0`$ to 35.0 at $`E=10`$, which is a contraction by a factor of 0.954. This contraction is analogous to the change in the smectic layer spacing under an electric field observed in x-ray diffraction experiments , and the contraction factor can be interpreted as the cosine of an x-ray tilt angle of $`17.5^{}`$. This x-ray tilt angle is somewhat smaller than the tilt angle of $`19^{}`$ associated with the eigenvectors of $`Q`$, which corresponds to the orientational ordering of the molecular cores observed in optical experiments. ### C Field-Induced Phase Transition We carried out further studies of the five-degree molecular system in a larger temperature range, and located the SmA-nematic transition at approximately $`k_BT=3.0`$. Above that temperature, the SmA phase melts and the system is stable as a nematic state, with low positional correlations (smectic order parameter below 0.3) but with very high orientational order. Figure 6a shows a slice of the nematic system at $`k_BT=3.3`$, viewed from the $`x`$ direction. When the temperature is lowered from $`k_BT=3.3`$ to $`k_BT=2.1`$, the system returns to the SmA phase with clearly defined layers (smectic order parameter of about 0.8). This is evidence that the system has a stable and reversible SmA-nematic phase transition. Under a strong electric field, the nematic phase has a surprising behavior. We apply $`E=10`$ to the nematic system at $`k_BT=3.3`$, not far above the nematic-SmA transition temperature. The system regains a large smectic order parameter and again forms clearly defined layers, as shown in Fig. 6b. This figure shows a slice of the system, with five layers in cross section. Thus we observe in this simulation a field-induced nematic-SmA phase transition. In experiments, electric-field-induced isotropic-nematic-smectic phase transitions have been observed in thermotropic liquid crystals , and the critical behavior of the field-induced molecular tilt near the nematic-SmA transition has been investigated . A good understanding of these effects in simulation will contribute to a better understanding of field-induced phase transitions in experiment. ## IV Discussion This simulation study shows that the molecular shape is very important for the phase behavior of liquid crystals. In the system with the $`45^{}`$ molecular bend angle, the steric repulsion based on molecular shape provides the driving force for molecular tilt order in a SmC phase, even without intermolecular dipole-dipole interactions. In the system with the $`5^{}`$ bend angle, the molecules are closer to rigid rods, so they do not exhibit a SmC phase with spontaneous tilt order. Still, even a $`5^{}`$ molecular bend leads to a substantial electroclinic effect, which would be totally absent for rigid rods. (Rigid rods with transverse electric dipoles would align their dipoles with an applied electric field, but this alignment would not lead to any molecular tilt.) Preliminary simulation results for molecules with a $`9^{}`$ bend angle (not presented here) suggest that the phase transitions shift dramatically from the $`5^{}`$ molecules, confirming the influence of small changes in molecular shape. Hence, one conclusion of this study is that collective intermolecular properties like molecular tilt and transition temperatures are quite sensitive to slight details of molecular shape. This conclusion is somewhat disappointing from the perspective of modeling unique properties of particular liquid-crystal compounds, as opposed to generic properties based on molecular symmetry, because it implies that one must describe the molecular structure very precisely in order to predict properties like tilt and transition temperatures. Another conclusion of this study is that the distribution of molecular tilts in the SmA phase is more complex than is often supposed. In the absence of an applied electric field, the molecules do not stand up as rigid rods along the layer normal. Rather, there is disorder in the molecular tilt, with all of the molecules tilting away from the layer normal in random azimuthal directions. Some of this disorder takes the form of vortices in the tilt projected into the smectic layer plane. When an electric field is applied, it has two effects: it increases the magnitude of the tilt angles and it increases the order in the azimuthal direction of the tilt. These two effects combine to give the electroclinic tilt angle associated with the eigenvectors of the nematic order tensor $`Q`$. For that reason, this tilt angle is somewhat greater than the x-ray tilt angle associated with the contraction of the smectic layers, which arises only from the increase in the magnitude of the molecular tilt angles. This result suggests that experimental measurements of the electroclinic effect cannot be interpreted purely as tilting of rigid rods or as ordering of $`xy`$ spins, but rather as a combination of both. The vortices observed in the SmA phase of the simulation are particularly intriguing defects. These vortices appear to be equivalent to the topological defects that mediate the Kosterlitz-Thouless ordering transition in the two-dimensional $`xy`$ model . Thus, they suggest that the SmA phase is analogous to the disordered phase of the $`xy`$ model and the SmC phase to the ordered phase. It is surprising that our three-dimensional simulation shows point vortices that are uncoupled from one smectic layer to the next, and do not thread through all five layers of the system. This uncoupling presumably occurs because, as noted earlier, there is very little interaction between the tilt directions in adjacent layers due to the short-range repulsive potential and the lack of interdigitation between layers. The observation of these defects leads to several questions for future research. For example, how do the defects evolve as a small electric field is applied? In a system with a SmA-SmC transition, what happens to the defects when the temperature drops toward the transition? Furthermore, if the interaction between molecules had a longer range, would the point-like “pancake” vortices turn into vortex lines as in conventional type II superconductors , or would they be driven out of the system completely? This final question is a key issue for experimental systems in which the tilt directions of adjacent layers are strongly coupled. In summary, we have simulated smectic ordering in liquid crystals composed of bent-rod molecules interacting through a soft repulsive potential. The system of highly bent molecules shows a SmC phase with spontaneous tilt ordering, while the system of only slightly bent molecules shows a SmA phase with a substantial induced tilt under an applied electric field. These results show the high sensitivity of molecular tilt ordering to the molecular shape, and show the distribution of molecular tilt that controls the electroclinic effect. ###### Acknowledgements. This work was supported by the U. S. Navy Grant No. N00014-97-1-G003, the National Science Foundation Grant No. DMR-9702234-1, and the Donors of the Petroleum Research Fund, administered by the American Chemical Society.
no-problem/9906/quant-ph9906046.html
ar5iv
text
# Untitled Document Derivation of the Pauli exchange principle A. A. Broyles Department of Physics, Institute for Fundamental Theory University of Florida, Gainesville, Fl 32611 When the time-independent wave function of a multicomponent system is computed, it is necessary to require that it be consistent with the Pauli exchange principle which states that, for systems involving identical elements, wave functions must be antisymmetric to pair exchange if twice the spin $`s`$ for each one is an odd integer and symmetric to pair exchange if the spin is an integer. For identical spin sets, each set containing spin $`s`$, this can be expressed mathematically, for a single exchange, in the form, $$\mathrm{\Psi }(𝐱_1\underset{¯}{s}\mathrm{}𝐱_a\underset{¯}{s}\mathrm{}𝐱_b\underset{¯}{s}\mathrm{}𝐱_n\underset{¯}{s})=(1)^{2s}\mathrm{\Psi }(𝐱_1\underset{¯}{s}\mathrm{}𝐱_b\underset{¯}{s}\mathrm{}𝐱_a\underset{¯}{s}\mathrm{}𝐱_n\underset{¯}{s})$$ $`(1)`$ where $`𝐱`$ is a set of space coordinates, and $`\underset{¯}{s}`$ represents a set of spin quantum numbers. Pauli showed that this relation connecting fermion and boson statistics with spin can be derived with the aid of relativistic quantum field theory. As will be seen below, it can also be derived directly from the properties of wave functions. Papers have been published by Feynman, Schwinger, and others in an effort to find a simpler and more understandable proof. Their work is discussed in a recent review article in which we find this statement, “Finally we are forced to conclude that although the Spin-Statistics Theorem is simply stated, it is by no means simply understood or simply proved.” Feynman’s position on the lack of a simple derivation in 1963 was expressed in his physics course where he said, “This probably means we do not have a complete understanding of the fundamental principle involved.” Broyles published an elementary proof of this theorem in 1976 in which he used an exchange operator defined in terms of rotation operators. Apparently unaware of this, Feynman presented a lecture giving essentially the same proof in 1986. However, Duck and Sudarshan objected to these proofs by expressing the need for evidence that the wave functions before and after the exchange are identical. This paper is written to answer this objection and to summarize the proof. Feynman introduced the term, “probability amplitude”, into quantum mechanics. The square of the magnitude of this complex number gives the probability of obtaining a specified collection of quantum numbers as the result of measurements. In order to identify this amplitude uniquely (aside from an overall constant phase), a complete set of quantum numbers must be specified. As an example, Kemble relates how he and Feenberg proved that the probability amplitude over all space for position coordinates can be determined by repeated measurements. These measurements can be used to construct the probability distribution and its time derivative. They show that a phase factor, constant in space and time, remains undetermined. Phases of probability amplitudes can be measured relative to a standard by means of diffraction experiments , but no method is available for measuring the phase of the standard. We shall assume here that a complete set of quantum numbers can be found to determine any probability amplitude to within a constant phase. These quantum numbers are associated in sets as though each set belonged to a particle. Each set can include the configuration space coordinates $`𝐱`$ as well as the spin coordinates in the subset $`\underset{¯}{s}`$. We shall include the spin $`s`$ and a spin component $`s_z`$ in $`\underset{¯}{s}`$. In addition, any other quantum numbers such as mass, charge, flavor, etc. that may be needed to identify the set, can be included in $`\underset{¯}{s}`$. In association with each set $`𝐱\underset{¯}{s}`$ , we can imagine a point in a representative spin configuration space whose coordinates are the components of $`𝐱\underset{¯}{s}`$. In Fig. 1a, points representing three sets, that we shall take to have identical spin quantum numbers $`\underset{¯}{s}`$ , are pictured. We shall be interested in systems where all the sets $`\underset{¯}{s}`$ are identical except possibly for the $`s_z`$ quantum numbers. Let us first consider cases where all the $`s_z`$’s are the same and where each one has its maximum value $`s`$. The wave function $`\mathrm{\Psi }`$ in Eq. (1) is a convenient way to write down this probability amplitude and its associated sets of quantum numbers. However it is clear that additional information has been incorporated in this wave function, additional information beyond that contained in the quantum numbers that determine the probably amplitude. This information is provided by the arrangement of the sets in the argument of the wave function. The number represented by the function $`\mathrm{\Psi }`$ will, in general, depend upon the order in which the sets of quantum numbers are written. The probability amplitude, on the other hand, depends only on what quantum number sets are present, not on the order in which they appear. Thus if the wave function is to represent the probability amplitude, we must impose a restriction on it to remove the dependence on the order. If the spins are zero, all we need to do is make the wave function symmetric to the pair exchange of identical sets. For nonvanishing spins, however, the restriction is more complicated. It is instructive to recognize a difference between the spatial exchange of two apparently identical classical objects and the exchange of two electron coordinate sets determining a probability amplitude. For example if we exchange two billiard balls, we produce a new state, a new configuration. Of course, if the two balls had exactly the same number of atoms of each kind, it might be possible to make them truly indistinguishable so that the state after the exchange is the same as the one before. Since, however, it is essentially impossible to equalize the number of atoms in the two balls, the probabilities are overwhelming that the initial and final states will be different. If however, two electron coordinate sets, $`𝐱_a\underset{¯}{s}`$ and $`𝐱_b\underset{¯}{s}`$, with identical $`\underset{¯}{s}`$ subsets, are exchanged, the quantum numbers present are the same after the exchange as before. Thus the probability amplitude, since it depends only on the sets that are present and not on their order, will be unaltered by the exchange. This is illustrated in Fig. 1 where the exchange in Figs 1b and 1c leaves Fig 1d identical to 1a. This is equivalent to saying that the $`𝐱\underset{¯}{s}`$’s that specify a probability amplitude are generic sets rather than specific . The argument of a wave function, is normally written down with the quantum number sets in specific positions. Thus there is one set in the first position to which we can assign a number one and label the symbols with subscripts so that we have $`𝐱_1\underset{¯}{s}_1`$. Similarly, the particular set in the second position can be written as $`𝐱_2\underset{¯}{s}_2`$, etc. Then an application of the exchange operator to sets $`a`$ and $`b`$ changes their positions in the argument of the wave function as we see in Eq. (1). In order to make the wave function equal the probability amplitude, it must not be allowed to change value as a result of this exchange. To generate the N! specific positions that can appear in the argument of the wave function from the values of the wave function at the N generic positions that are available from the probability amplitudes, we must first identify each generic set. This can be done by assigning an order number to it. One way of accomplishing this is to select a point in the representation space to act as the center. The generic positions can then be numbered according to their distances from this center. The point nearest to the center will be point number one, the next nearest will be point number two, etc. The distance from the center to the nearest point can then be represented by $`r_1`$, to the next nearest one by $`r_2`$, etc. Then we have $`r_1<r_2<r_3<\mathrm{}r_n`$. The probability amplitudes for these N points then compose the wave function for the region where the $`r`$’s satisfy these inequalities. We shall call this, Region A. Since the above procedure gives $`\mathrm{\Psi }`$ in a sizeable fraction of space, Region A, the assumption is made here that the $`\mathrm{\Psi }`$ in the remainder of space (where the $`r`$’s satisfy different inequalities) can be computed by analytic continuation. Since any order of $`r`$’s can be altered to a new order by exchanging two points in the representation space, it is convenient to obtain the collection of Taylor’s series for this continuation by expanding the rotation operators that move two points through angles. Such an exchange is illustrated in Figs. 1b and 1c. If it exchanges the points labeled 1 and 2 in the last paragraph, the distances from the center in the representation space are in the order $`r_2<r_1<r_3\mathrm{}r_n`$. These points are in a region other than Region A. If the sets to be exchanged are $`𝐱_a\underset{¯}{s}_a`$ and $`𝐱_b\underset{¯}{s}_b`$, the rotations required can be described by first placing the representation space coordinate system so that its $`x`$ axis coincides with the line connecting the two points at $`𝐱_a`$ and $`𝐱_b`$ and with the origin (marked O in Figs. 1b and 1c) half way between these points. Then the operator (with the subscript $`a,b`$ indicating “$`a`$ or $`b`$”), $$R_{a,b}^\varphi =e^{i\varphi _{a,b}J_{za,b}}$$ $`(2)`$ where $$J_{za,b}=L_{za,b}+S_{za,b}$$ $`(3)`$ and $$L_{za,b}=i\mathrm{}\frac{}{\varphi _{a,b}},$$ $`(4)`$ ($`S_{za,b}`$ is the generator of the spin component in the $`z`$ direction.) can be applied to $`\mathrm{\Psi }`$ to produce the $`\mathrm{\Psi }`$ after the required rotations. This rotation operator or its series can be used to compute the wave function with $`𝐱_a`$ or $`𝐱_b`$ moved by a rotation through an angle $`\varphi _{a,b}`$. If the spins in the argument of $`\mathrm{\Psi }`$ are zero, the spin generator in Eq. (3) can be omitted. However, if the spins do not vanish, the spin generator will be required since the direction of spin in the argument of $`\mathrm{\Psi }`$ after the rotation may be different from the initial one. When $`\varphi _a`$ equals $`\pi `$, $`\mathrm{\Psi }`$ will be calculated at the points resulting from the rotation of $`𝐱_a\underset{¯}{s}_a`$ as shown in Fig. 1b. Next, setting $`\varphi _b`$ to $`\pi `$ provides the $`\mathrm{\Psi }`$ after the rotation of $`𝐱_b,\underset{¯}{s}_b`$ through $`\pi `$ as shown in Fig 1c. The two rotations together $`R_aR_b`$ then produce the $`\mathrm{\Psi }`$ with the two points exchanged. See References and for additional discussions. By applying the proper choice of these rotation operators to $`\mathrm{\Psi }`$, the wave function with any exchange of sets can be computed. By computing the proper choice of exchanges, the wave function for any specific configuration of sets can be determined from a generic one. In the case under consideration where the spin components are identical, that is, $`s_{za}=s_{zb}=s`$, the $`L_z`$ operators will exchange the $`𝐱`$’s while the $`S_z`$’s can be replaced by their eigenvalues. As a result, the two spin rotation operators applied to $`\mathrm{\Psi }`$ to produce an exchange will give a factor, $`e^{i2s_{za}\pi }=(1)^{2s}`$. Thus $$R_aR_b\mathrm{\Psi }(\mathrm{}𝐱_a\underset{¯}{s}\mathrm{}𝐱_b\underset{¯}{s}\mathrm{})=(1)^{2s}\mathrm{\Psi }(\mathrm{}𝐱_b\underset{¯}{s}\mathrm{}𝐱_a\underset{¯}{s}\mathrm{}).$$ $`(5)`$ where the $`\varphi _{a,b}`$ in Eq. (2) equals $`\pi `$ in $`R_{a,b}`$. But, as we have seen, exchanging these coordinate sets does not introduce new sets or remove old ones and, therefore, does not change the probability amplitude. Thus $`R_aR_b`$ must have no effect on $`\mathrm{\Psi }`$ if $`\mathrm{\Psi }`$ is to equal the probability amplitude. As a result, the left hand side is equal to $`\mathrm{\Psi }(\mathrm{}𝐱_a\underset{¯}{s}\mathrm{}𝐱_b\underset{¯}{s}\mathrm{})`$. This makes the last equation equivalent to Eq. (1). The effect of exchanging two points lying on a node can be computed by a limiting procedure. If the quantum numbers in the various spin sets in the argument of $`\mathrm{\Psi }`$ are identical except for the values of $`s_z`$, the Pauli principle again applies. To prove this, it is convenient to make use of the complete set of spin wave functions (defined with the aid of an integer $`l`$) with the property, $$\chi (s,\theta _l)=e^{i\theta _lS_x}\chi (s,0),\theta _l=l\pi /2s,0l2s,$$ $`(6)`$ introduced in the appendix of Reference . The $`\theta _l`$’s then replace the $`s_z`$’s in the argument of the wave function. A wave function like that in Eq.(1) can then be written in the form, $$\mathrm{\Psi }(\mathrm{}𝐱_as\theta _{la}\mathrm{}𝐱_bs\theta _{lb}\mathrm{})=\mathrm{}e^{i\theta _{la}S_{xa}}\mathrm{}e^{i\theta _{lb}S_{xb}}\mathrm{}\mathrm{\Psi }(\mathrm{}𝐱_as0\mathrm{}𝐱_bs0\mathrm{})$$ $`(7)`$ where the zero’s indicate values of the $`\theta _l`$’s. It is understood that the subscript $`j`$ on $`S_{xj}`$ is the same as the subscript on the $`𝐱_j`$ in the set on which it operates. Since, as we have seen in the last section, $`\mathrm{\Psi }(\mathrm{}𝐱_as0\mathrm{}𝐱_bs0\mathrm{})`$ satisfies Eq.(1), and the operators, $`e^{i\theta _{la}S_{xa}},e^{i\theta _{lb}S_{xb}},\mathrm{}`$ commute, it follows that $`\mathrm{\Psi }(\mathrm{}𝐱_as\theta _a\mathrm{}𝐱_bs\theta _b\mathrm{})`$ will also satisfy Eq.(1). The author is indebted to J. R. Klauder, R. L. Coldwell, C. B. Thorn, and H. P. Hanson for discussions about this manuscript. REFERENCES W. Pauli, Phys. Rev. 58, 716 (1940). I. Duck and E. C. G. Sudarshan, Am. J. Phys. 66, 284 (1998). R. P. Feynman, R. B. Leighton, and M. Sands, The Feynman Lectures on Physics (Addison-Wesley, Reading, Mass., 1965) Vol. 3, p. 162. A. A. Broyles, Am. J. Phys. 44, 340 (1976). , “The Reason for Antiparticles” in R. P. Feynman and S. Weinberg, Elementary Particles and the Laws of Physics (Cambridge University Press, New York, 1987). R. P. Feynman, The Theory of Fundamental Processes (W. A. Benjamin, New York, 1962), Chapter 1. E. C. Kemble, The Fundamental Principles of Quantum Mechanics (McGraw-Hill, New York, 1937), First Edition, p. 71. J. J. Sakurai, Modern Quantum Mechanics (Benjamin/Cummings, Menlo Park, CA) p. 162. D. ter Haar, Elements of Statistical Mechanics (Rinehart, New York, 1954), p. 136, last paragraph. Ramamurti Shankar, Principles of Quantum Mechanics (Plenum Press, New York, 1980) Chapter Sec. 12.5. Figure 1. The exchange, by means of rotations, of two points representing two sets of quantum numbers. Labels are unnecessary because we take the spin quantum numbers to be identical. The coordinate axes are located so that the origin is half way between the two points to be exchanged. 1a. The projection into the $`xy`$ plane of three points representing three sets of quantum numbers. 1b. The rotation of one of the points through an angle $`\pi `$ around the $`z`$ axis with the origin at $`O`$. 1c. The rotation of the other point through an angle $`\pi `$. 1d. The representative points after the two rotations that result in the exchange. Fig. 1d is identical to 1a.
no-problem/9906/astro-ph9906363.html
ar5iv
text
# Iron line signatures in X-ray afterglows of GRB by BeppoSAX ## 1 Introduction Distance - scale determination of Gamma-ray bursts (GRB) has been one of the most important achievements of astrophysics in recent years. Accurate and fast localization of the prompt and afterglow emission (Piro et al.1998a ; Costa et al. (1997)) by BeppoSAX (Piro, Scarsi & Butler (1995); Boella et al. (1997)) led to the identification of optical counterparts (van Paradijs et al. (1997)) and ultimately to spectral measurements of a redshift (Metzger et al. (1997)). While the extragalactic origin of GRB has gathered solid evidence in its support, the source of the large energy implied by their distance is still speculative. The measurement of X-ray Fe lines emitted directly by the GRB or its afterglow could provide a direct measurement of the distance and probe into the nature of the central environment (Perna & Loeb (1998); Mészáros & Rees (1998); Boettcher et al. (1999); Ghisellini et al. (1998)). Neutron star – neutron star merging should happen in a fairly clean environment, with line intensities much below the sensitivity of current experiments. In contrast, Mészáros and Rees (1998) have shown that the circumburst environment created by the stellar wind before the explosion of the hypernova could yield a line of substantial intensity. A similarly favourable situation should be expected in a variation of the hypernova scenario, – the SupraNova (Vietri & Stella (1998)), where the GRB is shortly preceded by a supernova explosion with the ejection of an iron–rich massive shell. It is also conceivable that the impact of the relativistic shell that produced the original GRB on these ejecta could provide an additional energy input in the afterglow. Motivated by these expectations, we have started a detailed analysis of afterglow spectra to look for the presence of features. The first and most promising candidate is GB970508. It is characterized by a large outbursting event during its afterglow phase (Piro et al.1998b ) and has the highest signal to noise ratio of the BeppoSAX GRB afterglows. Here we summarize the results obtained in this burst (reported more extensively in Piro et al.1999) and present the first results of the analysis in other afterglows. ## 2 Possible evidence of Fe line feature in GB970508 We have searched the X–ray spectrum of GB970508’s afterglow for an iron line, located at the system’s redhift (z = 0.835, Bloom et al. 1998); we found such a line with limited statistical significance ($`99.3\%`$) in the early part (first $`16h`$) of the afterglow; the line decreases in the later part of the observations ($`1`$ day after the burst) by at least a factor 2, enough to make it undetectable with current apparatus. Simultaneously with the line disappearance, the X–ray flux both rises and hardens ($`\alpha =0.4\pm 0.6`$, while $`\alpha =1.5\pm 0.6`$ before the reburst), consistent with the appearance of a new shock. Then, at the end of the outburst, the spectrum steepens. ## 3 Search of Fe line in other GRB’s afterglows An extensive analysis is under way to search for features in all the X-ray afterglows observed by BeppoSAX. We report here the results derived in relatively bright afterglow sources (flux in the first part of the observation $`>3\times 10^{13}\mathrm{erg}\mathrm{cm}^2\mathrm{s}^1`$) which have a spectroscopic redshift determination, thus allowing to fix the energy of the line. In the two GRB’s that meet these conditions (GRB971214 and GRB980613) we find upper limits that are roughly consistent with the intensity measured in GB970508 (Tab.1) ## 4 Origin of the line and constraints on the emitting region It is most likely that the line of GB970508 is produced by fluorescence and recombination of Fe atoms ionized by the intense flux of the GRB and its afterglow (Piro et al. (1999), Lazzati et al.1999) In the early phases of the GRB the radiation field is so high that iron atoms are completely stripped of their electrons: the Compton temperature is very high and then recombination is not very efficient in producing line photons. When the flux decreases, about $`10^4s`$ after the burst, fluorescence becomes an effective process. We note, in passing, that the intensity of the line is therefore not correlated with the luminosity of the burst: for example, with a luminosity a factor of 10 larger, the medium would have remained completely ionized upto about 1 day after the burst, producing therefore a line with a lower intensity. The minimum mass needed to produce the line is (Piro et al.1999, Lazzati et al.1999) $`M_{min}=0.5M_{}A_{Fe}^1`$, where $`A_{Fe}`$ is the iron abundance normalized to the solar value. From the line variability, intensity and width we deduce that this medium should be located at a distance of $`3\times 10^{15}`$cm from the central source, it is moving with subrelativistic speed, it should have a large density ($`n>5\times 10^9cm^3`$), and it should lie sideways respect to the observer, otherways it would smear out the short timescale structure of the burst with Thomson scattering (Boettcher et al. (1999)). In order to reach such distance, this material must have been pre–ejected by the source originating the burst, shortly (perhaps a year, for typical SN expansion speeds) before the burst. We stress that these observations contain two coincidences: on the one hand, this is the only burst in which a reburst and a line have been observed by BeppoSAX; on the other, the iron line disappears exactly at the moment of the reburst. We also point out that a line feature, with a similar significance, has been found by ASCA in another burst, GRB970828, which also shows an event of rebursting during the X-ray afterglow (Yoshida et al. (1999)). On the contrary, neither GB971214 nor GB98613 show rebursting (Costa 1999). ###### Acknowledgements. We thank the BeppoSAX team for the support with observations. BeppoSAX is a program of the Italian space agency (ASI) with the participation of the Dutch space agency (NIVR).
no-problem/9906/cond-mat9906001.html
ar5iv
text
# Impurity-Induced Bound Excitations on the Surface of Bi2Sr2CaCu2O8 \[ ## Abstract We have probed the effects of atomic-scale impurities on superconductivity in Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> by performing low-temperature tunneling spectroscopy measurements with a scanning tunneling microscope. Our results show that non-magnetic defect structures at the surface create localized low-energy excitations in their immediate vicinity. The impurity-induced excitations occur over a range of energies including the middle of the superconducting gap, at the Fermi level. Such a zero bias state is a predicted feature for strong non-magnetic scattering in a d-wave superconductor. PACES numbers: 74.50.+r,73.20.Hb,74.72Hs \] There is now a great deal of evidence that the superconducting state in a number of high-T<sub>c</sub> superconductors has a dominant d$`_{x^2y^2}`$ symmetry . One key evidence is that non-magnetic impurities, which only affect conventional s-wave superconductors weakly, can act as strong pair-breakers in the high-T<sub>c</sub> superconductors. However, to date, there are no atomic scale studies of superconductivity in the immediate vicinity of individual impurities in a high-T<sub>c</sub> superconductor. Such experiments are motivated by recent theoretical studies of the local response of a d-wave superconductor to individual impurities. These theories predict local signatures of d-wave pairing and emphasize the importance of local variations in the electronic properties of a d-wave superconductor. There is also now growing theoretical and experimental evidence that quasi-particle scattering from surfaces and twin boundaries of d-wave superconductors gives rise to effects that have no analog in conventional s-wave superconductors. The scanning tunneling microscope (STM) offers a direct method to examine the spatial variation of the electronic properties in the superconducting state near defect structures. In this paper, we report on STM measurements of the local density of states near regions of a d-wave superconductor that have been perturbed by the presence of atomic-scale impurities. Our results show that non-magnetic impurities induce low-energy excitations in a d-wave superconductor, which are localized on length scales comparable to the superconducting coherence length. Such impurity-bound excitations for non-magnetic impurities is a predicted signature of a d-wave superconductor and is in stark contrast to behavior of conventional s-wave superconductors, in which to create similar effects, the impurities need to be magnetic. More specifically, we observe that the impurity-induced excitations for a d-wave superconductor can occur as a pronounced and narrow resonance at energies close to the Fermi level, E<sub>F</sub>. Such a zero bias feature has been predicted to occur for impurity scattering in a d-wave superconductors in the unitarity limit or for strong impurities when considering the order parameter suppression. From a different perspective, a zero bias resonance has also been predicted and observed for Andreev scattering from interfaces at which quasi-particles experience a sign change of the d-wave order parameter and form a surface bound state at E<sub>F</sub>. The similarity suggests that multiple Andreev scattering similar to that proposed at interfaces is at work in the bulk at strongly scattering impurities. We performed our experiments using an ultrahigh vacuum (UHV) STM which operates at low temperatures. The Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> single-crystal samples were grown using directional solidification technique and re-oxygenated prior to the STM measurements. The experiments reported here were carried out on overdoped single-crystal samples with a superconducting transition temperature at 74K and a transition width of 3K, as characterized by magnetometry measurements. Samples were introduced into the UHV chamber at room temperature and mechanically cleaved prior to STM measurements performed at low temperatures (T=5K.) We used a polycrystalline Au wire as our tip; however, the chemical identity of the last atom on the tip is unknown. The local quasi-particle density of states (LDOS) of the Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> surface was obtained from measurements of the differential conductance $`dI/dV`$ (where I is the current) of the STM junction versus sample bias voltage V (with respect to the tip) performed under open feedback loop conditions. Figure 1 shows a constant current STM topograph of the cleaved Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> surface. The weakest bond in the crystal is between the two adjacent BiO layers; therefore it is most likely that the top-most atomic layer of a cleaved sample is BiO. The image in Figure 1 shows a long length scale modulation (period of 27Å) which has been previously observed with the STM and has been associated with the relaxation of Bi atoms in BiO layer. Imaging the Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> surface in several regions showed that there are intrinsic defects on the surface. The intrinsic impurities are difficult to identify, except in some cases, such as that shown in the inset of Figure 1, where defects appeared as protrusions in STM topographs. The intrinsic defects are of fundamental interest, since they are considered crucial to understanding the deviations of low-temperature properties of high-T<sub>c</sub> superconductors from those expected for a clean d-wave superconductor. Our STM measurements near intrinsic surface defects show that such defects indeed modify the LDOS of Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> surface. For example, Figure 2 shows the measurements of LDOS with the tip “on” and “off” the intrinsic surface defect shown in the inset of Figure 1. There is a clear enhancement of the LDOS at low-energies at this impurity site, as compared to that measured at lateral distance of about a coherence length away. The change in $`dI/dV`$ near $`E_F`$ corresponds to about 20% of that measured at voltages above the superconducting gap. However, we found other intrinsic surface defects which did not modify the low-energy LDOS as much, but only affected the quasi-particle peaks or the LDOS background measured at higher voltages. Overall, the STM spectra over many regions of the “bare” surface, away from any surface defects, were comparable to those previously reported in the overdoped regime. In the inset of figure 2 we show a collection of spectra from different regions of two samples with similar doping level. The values of the maximum gap ($`2\mathrm{\Delta }_{max}/k_bT_c`$ 9), the asymmetric background, and other features of our spectra such as the width of the quasi-particle peaks are similar to those previously reported by Renner et al. for a sample with a similar doping level and T<sub>c</sub>. The intrinsic surface defects that do modify the LDOS at low-energies can be considered as evidence for impurity-induced excitations due to non-magnetic scattering in a d-wave superconductor. However, the magnetic state of such defects is unknown, hence to illustrate the main point of this work, we focus the rest of this paper on ostensibly non-magnetic defects created with the Au STM tip. Such defects were deposited onto the surface by bringing the Au STM tip close enough to the surface to cause transfer of atoms from the tip to the sample. We monitored the current through the STM junction during this procedure and retracted the tip once the current began to saturate signaling the formation of a small contact. In the regime of close contact, it has been shown that an STM junction resistance, $`R_J25k\mathrm{\Omega }`$, is indicative of formation of a single metal atom contact between the tip and the surface. The STM topographs of the Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> surface measured after making such contacts revealed the deposition of atoms from the tip onto the sample, similar to that observed when Au tips contact other conducting surfaces. An example of tip-deposited defects is shown in the topograph in Figure 3A, which shows a 36Å square area of the surface with two defects appearing as protrusions with an apparent height of 0.8Å. The LDOS of the Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> surface is dramatically modified by defects deposited from the Au STM tip. For the defect in the center of figure 3A these modifications are demonstrated by the data in figure 4A. This figure shows measurements of the LDOS with the tip over the defect structure (centered at maximum height) and that at lateral distances away (in the direction indicated in Figure 3A). Comparing these spectra shows that the defect induces low-energy excitations, within the superconducting gap, in its immediate vicinity. At the impurity site, the value of $`dI/dV`$ measured close to $`E_F`$ is about five times larger than that measured at voltages above the superconducting gap. A closer examination of data in Figure 2A shows that the impurity-induced resonance is made of two asymmetric peaks–one above and one below $`E_F`$ (within 1meV). It is also interesting to note that this asymmetric behavior is similar to that of the quasi-particle peaks measured with tip over a region far away from the impurity. The enhancement of the low-energy excitations is a common characteristic among the different defects deposited from the Au STM-tip. This behavior is illustrated in Figure 4B for another defect. The inset of this figure also shows difference spectra for several different defects with similar behavior (see caption for details). Another common characteristic is that the impurity-induced excitations are localized to within lateral distances of about 20Å around the impurities. This spatial characteristic was also imaged by measuring the AC $`dI/dV`$ at a fixed voltage while scanning the tip in constant DC current mode. Such an image is shown in Figure 3B, along with the constant current topographs (Figure 3A) acquired simultaneously from an area of the surface which includes two defect structures. The bound excitations for each of the defects are localized at the dark regions in this gray scale image. The data above demonstrate the central result of this paper, that non-magnetic defects alter the local LDOS of Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> surface by inducing low-energy excitations. These excitations are bound to the impurities on length scales comparable to the superconducting coherence length. We emphasize that similar experiments on a Nb surface have previously shown the LDOS of a conventional superconductor to be insensitive to the presence of atomic-scale non-magnetic defects deposited from a Au tip . Theoretical works on impurity scattering in a d-wave superconductor have predicted that an isolated non-magnetic defect gives rise to a semi-bound excitation with energy $`E_B<\mathrm{\Delta }_{max}`$. In this theoretical picture, $`E_B`$ is determined by the strength of the impurity scattering potential, with $`E_B=E_F`$ for scattering in the unitarity limit. More recent efforts show that including the local suppression of the superconducting order parameter in the calculations drives the effective scattering strength of even a strong impurity to the unitarity limit. The impurities discussed above modify the LDOS over a range of energies, many of which induce sharp resonances very close to $`E_F`$, and are in the unitarity limit. The asymmetric or splitting of measured resonances near $`E_F`$ support the possibility that impurities locally break the electron-hole symmetry . However, the underlying electronic states of Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub> has also been suggested as a cause for the asymmetry.. The spatial structure of the resonance for an isolated impurity has been predicted to be cross-shaped, reflecting the anisotropy in the superconducting order parameter. Theoretical efforts have emphasized that the spatial and angular character of single-impurity resonances greatly affect their overlap and the nature of the many-impurity state. The overlap between these resonances for bulk impurities determines the formation of low-energy impurity bands and the behavior of the thermodynamic properties at low temperatures.. The single impurity-induced excitations measured here appear to be localized to a region of about a coherence length, and do not show the predicted cross-shaped pattern, which has a delocalized nature along the nodes of the d-wave order parameter. More detailed high-resolution spatial measurements may be required to resolve this fine structure. The observation of the impurity-induced excitation is also intriguing within the context of local electronic properties of a d-wave superconductor near other spatial perturbations such as surfaces, crystal twins, and grain boundaries. The scattering from these boundaries is expected to give rise to Andreev bound states whenever the incident and reflected quasi-particles experience the sign change of the d-wave order parameter, i.e. when the boundary is normal to the direction in momentum space along d-wave nodes. Directional tunneling spectroscopy in planar tunnel junction, point contact spectroscopy, and grain boundary tunnel junctions have shown evidence for such Andreev states. The experimental data reported here show that the scattering at isolated atomic-scale impurities can also induce similar resonances at $`E_F`$. In fact, a theoretical connection between scattering processes at impurities and those at interfaces can be made within a semi-classical approximation. Perturbations such as defects or surfaces cause scattering and interference between electronic states from different regions of k-space. These processes together with the sign change of the order parameter inherent to a d-wave superconductor give rise to the zero energy resonances observed in the experiments. In conclusion, we have shown that ordinary non-magnetic impurities can induce localized low-energy excitations in a d-wave superconductor. For some of the impurities, we observed the impurity-induced resonance to be essentially in the middle of the gap at $`E_F`$, as predicted for impurity scattering in the unitarity limit. Such a zero bias state is also the hallmark of Andreev scattering in a d-wave superconductor whenever scattering of the quasi-particles can explore the sign change of the d-wave order parameter. The experimental data described here were obtained at IBM Almaden Research Center. It is our pleasure to thank L. H. Greene, D. J. Van Harlingen, J. A. Sauls, G. Blumberg, D. K. Campbell, and especially A. V. Balatsky, M. E. Flatté, A. Shnirman, I. Adagideli, and P. M. Goldbart for fruitful discussions. Note Added: After submission of this manuscript, we have become aware of another STM study of defects in Bi<sub>2</sub>Sr<sub>2</sub>CaCu<sub>2</sub>O<sub>8</sub>..
no-problem/9906/astro-ph9906458.html
ar5iv
text
# ENERGY CRISIS IN ASTROPHYSICS (Black Holes vs. N-Body Metrics) ## 1 Introduction The difficulty experienced by general relativity in accounting for the prodigious energy released in the gamma ray bursters, particularly GRB 990123 (Schilling (1999)) focusses attention on the problems in the treatment of gravitational energy which the theory has had from its inception. (The quantity proposed by Einstein for the gravitational field stress-energy turned out not to be a tensor (Weyl (1922)).) The Schwarzschild metric solution, with the interpretation of ”event horizon” and ”black hole”, limits the mass of a neutron star to about $`2.8M_{\mathrm{}}`$. But of even greater importance, the existence of interactive N-body solutions is incompatible with the presence of an event horizon. N-body interactive solutions, however, are necessary to describe the properties of neutron stars, both in the sense of their being a collection of neutrons (Kapusta (1989)), and in mergers involving two or more neutron stars as macroscopic N-body systems. We also need such solutions to study the formation of neutron stars themselves during gravitational collapse and the conversion of gravitational energy into radiation in the case of merging neutron stars or in accretion onto neutron stars in X-ray binaries. These are just some important examples. All of astrophysics requires interactive N-body solutions which are not present in general relativity. A resolution of this problem exists in the relativistic curved spacetime gravitation theory of Yilmaz which completes the approach initiated by Einstein. The fundamental differences between the two theories are presented in this paper in the form of a theorem about the N-body solutions. The intent is to provide a brief theoretical underpinning for Robertson’s proposed explanation of GRB 990123 (Robertson 1999b ) and to note other astrophysical consequences of this new view. ## 2 The Problem of Interactive N-Body (Multiparticle) Solutions We begin with the obvious remark that, in order to do physics with any set of objects, we must have more than one object so that we may study their relationships, their interactions, their scatterings, their coalescence, and so on. In other words, an acceptable physical theory must have “interactive N-body solutions.” By interactive, we mean the bodies exert forces on each other, or accelerate in each other’s fields when free of constraints. We therefore propose to investigate whether a theory of gravity has N-body interactive solutions. In 1974 the British Canadian mathematician Brian O.J. Tupper has shown that in any four-dimensional spacetime theory of gravity viable with respect to the four classical weak field tests the slow motion (sometimes called static) limit $`u^i0`$, $`u_0u^01`$ the field equations are of the form (Tupper 1974a ; 1974b ) $$\frac{1}{2}G_\mu ^\nu =\tau _\mu ^\nu +\lambda t_\mu ^\nu $$ (1) $$\tau _\mu ^\nu =\sigma u_\mu u^\nu ,$$ (2) $$\sqrt{g}\sigma =\mathrm{\Sigma }_Am_A\delta ^3(𝒙𝒙_A)$$ (3) where $`\tau _\mu ^\nu `$ is the Einstein “matter-stress-energy” tensor, and $$t_\mu ^\nu =_\mu \varphi ^\nu \varphi +\frac{1}{2}\delta _\mu ^\nu ^\rho \varphi _\rho \varphi $$ (4) is the Yılmaz gravitational “field stress-energy” tensor and $`\lambda `$ is an arbitrary numerical parameter passing through $`\lambda =0`$ (Einstein’s theory) and $`\lambda =1`$ (Yılmaz’ theory). The $`\varphi `$ is the low velocity limit of $`\varphi _0^0`$ when $`\varphi =\mathrm{trace}\varphi _\mu ^\nu =\varphi _0^0`$. Thus in this limit $`\varphi `$ is a scalar. Remarkably, Tupper was able to solve the above equations exactly for arbitrary $`\lambda `$ as $$ds^2=Adt^2A^1(1ϵ^2\varphi ^2/4)^2(dx^2+dy^2+dz^2)$$ (5) $$\varphi =m/r$$ (6) where $`A=[(1ϵ\varphi /2)/(1+ϵ\varphi /2)]^{2/ϵ}`$, $`\lambda =1ϵ^2`$. (We use $`2\lambda `$ where Tupper used $`\lambda `$. Note also that Tupper used spherical coordinates whereas we use its transform into cartesian coordinates in order to analyze more simply the N-body solutions.) Tupper noted that for $`\lambda =0`$ (that is, $`ϵ=\pm 1`$) and for $`\lambda =1`$ (that is, $`ϵ=0`$) this metric indeed reduces to the Schwarzschild and Yılmaz metrics respectively. What Tupper did not emphasize is that, while the Schwarzschild metric $$ds^2=[(1\varphi /2)/(1+\varphi /2)]^2dt^2(1+\varphi /2)^4(dx^2+dy^2+dz^2)$$ (7) $$\varphi =m/r$$ (8) is only a 1-body solution $`\varphi =m/r`$, the Yılmaz metric is an N-body solution (Yılmaz (1958)) $$ds^2=e^{2\varphi }dt^2e^{2\varphi }(dx^2+dy^2+dz^2)$$ (9) $$\varphi =\mathrm{\Sigma }_Am_A/r_A+C$$ (10) where $`r_A=|𝒙𝒙_A|`$. It reduces to $`\varphi =m/r`$ only as a special case when only one body is present. Note also that in the $`ϵ0`$ case one has an event horizon at $`r_{eh}=ϵm/2`$. In the Yılmaz case $`(ϵ=0)`$ there are no event horizons (no black holes). Thus we will show that there is a strict mathematical anticorrelation between having an event horizon and having N-body solutions. The importance of this result, namely the existence of interactive N-body solutions in Yılmaz’ theory was not appreciated until recently because it was always assumed (or hoped) that such N-body solutions would someday be found in general relativity. However, despite the many able mathematicians and well motivated groups and alliances working on it during the last eighty years, no N-body interactive solution has been found in general relativity. It is only relatively recently that, using the general $`\lambda `$-parametric solution of Tupper, Yılmaz proved that (Yılmaz 1987, 1992) they do not exist except when $`\lambda =1`$. Below we present this important theorem and discuss its consequences, including the energy requirement of GRB 990123. ## 3 Proof of the N-Body Theorem Since no other generalization than Tupper’s $`\lambda `$-parametric form is viable, and since the exact solution for arbitrary $`\lambda `$ is already given, a most interesting thing to do would be to evaluate (taking Eqs. (1) to (5) into account) both sides of the field equations with an unspecified $`\varphi `$ and see what happens (what conditions there are in order for the $`\varphi `$ to be a solution, where the Laplacians occur, etc.). In fact, something remarkable happens – one gets the following exact result: | | | Left Hand Side | | Right Hand Side | | --- | --- | --- | --- | --- | | $`\frac{1}{2}\sqrt{g}G_0^0`$ | $`:`$ | $`\mathrm{\Delta }\varphi +\frac{1}{4}(\lambda 1)\mathrm{\Omega }_{00}+\lambda t_{00}`$ | $`=`$ | $`\mathrm{\Sigma }_Am_A\delta ^3(𝒙𝒙_A)+\lambda t_{00}`$ | | $`\frac{1}{2}\sqrt{g}G_i^k`$ | $`:`$ | $`\frac{1}{4}(\lambda 1)\mathrm{\Omega }_{ik}+\lambda t_{ik}`$ | $`=`$ | $`\lambda t_{ik}`$ | where $`\mathrm{\Delta }`$ is the ordinary Laplacian. $`\mathrm{\Omega }_{00}`$ and $`\mathrm{\Omega }_{ik}`$ are given by $$\mathrm{\Omega }_{00}=2\varphi \mathrm{\Delta }\varphi $$ (11) $$\mathrm{\Omega }_{ik}=\varphi _i_k\varphi 3_i\varphi _k\varphi +\delta _{ik}^j\varphi _j\varphi \delta _{ik}\varphi \mathrm{\Delta }\varphi .$$ (12) We can see immediately that if $`\lambda =1`$ we would have the (needed) N-body solutions of the form (computations are done by a Mathematica symbolic manipulation program) $$\varphi =\mathrm{\Sigma }_Am_A/r_A+C$$ (13) but if $`\lambda 1`$, $`\mathrm{\Omega }_{00}`$ and $`\mathrm{\Omega }_{ik}`$ will have to vanish. The question is, can we get these two terms to vanish. Let us first note that, if we ignore the additive constant $`(C=0)`$, this can be done in the case of $`\mathrm{\Omega }_{00}`$. For, we can take the body as a small sphere of constant matter density in which case the potential may be assumed to start from the center as $`\sigma r^2/6`$, hence $`\mathrm{\Omega }_{00}=\sigma ^2r^2/3`$. Since this expression can be made as small as one likes, the $`G_0^0`$ part of the field equations allows an N-body solution of the type $`\varphi =\mathrm{\Sigma }_Am_A/r_A`$. But no matter how hard we try (including the above trick on its last term) we cannot get $`\mathrm{\Omega }_{ik}`$ to vanish with an N-body solution where $`N`$ is greater than one (for $`N>1`$, $`\mathrm{\Omega }_{ik}`$ has no roots). But for $`N=1`$, that is for $`\varphi =m/r`$, it can easily be shown that this special case is allowed (because $`\mathrm{\Omega }_{ik}=0`$ for $`\varphi =m/r`$) which is the original $`\varphi =m/r`$ in the Schwarzschild metric. Thus for the $`\lambda =0`$ case we have a strange situation where the $`G_0^0`$ component of the field equations allows an N-body solution $`\varphi =\mathrm{\Sigma }_Am_A/r_A`$ but the $`G_i^k`$ components of the same equations do not allow that solution. We cannot even argue that, due to nonlinearity, $`\mathrm{\Omega }_{ik}=0`$ may require a form different than the $`\varphi =\mathrm{\Sigma }_Am_A/r_A`$ because the $`\mathrm{\Omega }_{00}=0`$ part already accepts that form. This failure is normally overlooked. The N-body solution is usually assumed for general relativity in passing to a Newtonian limit. But in general relativity the Newtonian limit is satisfied only in first order. Here we are concerned with second order quantities, $`t_{ik}`$, $`\mathrm{\Omega }_{00}`$ and $`\mathrm{\Omega }_{ik}`$. When we include the requirement of the additive constant the situation gets worse. For, in this case we cannot get even the original Schwarzschild solution. The reason is that with the additive constant $`C`$, neither $`\mathrm{\Omega }_{00}=2\sigma C`$ nor $`\mathrm{\Omega }_{ik}=C\varphi _{ik}\delta _{ik}\sigma C`$ can be made zero independently of $`C`$ so as to satisfy the field equations. On the other hand, the existence of a $`\varphi +C`$ is of utmost importance. The essence of the additive constant is that if $`\varphi `$ is a solution to the field equations, then $`\varphi +C`$ must also be a solution to the same field equations. This means that the equations (and therefore the metric) must not depend on the absolute value of the potential $`\varphi `$. They must depend only on the “potential differences”. With the $`C`$ invariance also imposed in general relativity we cannot get any solution at all. The only way to get the desired N-body solutions is to set $`\lambda =1`$. In this unique case the event horizon disappears and we have no black holes. This case allows one to set the zero of the potential (more generally the potentials) at the observation point as $`\varphi (x)\varphi (x)\varphi (x_0)`$ where $`x_0`$ is the position of the observer. Kinematics then becomes locally Minkowskian $`g_{\mu \nu }\eta _{\mu \nu }`$ which allows one to take over the measurement procedure of special relativity locally. This means that the local vacuum velocity of light is c irrespective of accelerations of the observer. This interpretation is a fundamental prediction of the $`\lambda =1`$ theory and can be tested experimentally (Yılmaz (1987)). To this end, measurements of one-way light propagation times, using 100 picosecond pulses of laser light and transported hydrogen maser clocks, have been made over a 20 km East-West component path. So far the results are inconclusive as to whether the East-West and West-East times are equal or not on the rotating Earth. With improved equipment, now available, conclusive measurements can be obtained (Alley (1992)). ## 4 Interactive Nature of the N-Body Solutions As we have emphasized in Section 2, to have N-body solutions is not enough. One must also show that the bodies interact, and in a way consistent with observations. At first sight the linearity of the Poisson equation leading to the N-body solutions may give the wrong impression that, in the $`\lambda =1`$ case, there may be no interaction between the bodies hence no accelerations. Such a conclusion is not correct, because, in the Newtonian theory too, we have such a linear potential and in the Newtonian theory we have interactions and accelerations. Here, as the bodies move, other components $`\varphi _0^k`$, $`\varphi _i^k`$ of the field also develop. To exhibit the interactive nature of the N-bodies most clearly we now introduce the more general form of the Yılmaz theory. Written in a Minkowski background (for purposes of correspondence to special relativity, see Note 1) the Yılmaz theory can be given by two equations plus a coordinate (gauge) condition for definiteness, as (Yılmaz (1992)) $$\frac{1}{2}G_\mu ^\nu =\tau _\mu ^\nu +t_\mu ^\nu $$ (14) $$\sigma du_\mu /ds=\frac{1}{2}_\mu g_{\alpha \beta }(\tau ^{\alpha \beta }+t^{\alpha \beta })$$ (15) $$_\nu (\sqrt{g}g^{\mu \nu })=0$$ (16) First of all these equations show a relationship between general relativity and Yılmaz’ theory. The former is a truncated case where the $`t_\mu ^\nu `$ is removed as in $`t_\mu ^\nu \lambda t_\mu ^\nu `$, $`\lambda =0`$. But the presence of $`t_\mu ^\nu `$ is of crucial importance because there exists an identity which states that (Note (2)) $$\frac{1}{2}_\mu g_{\alpha \beta }(\tau ^{\alpha \beta }+t^{\alpha \beta })(\sqrt{g})^1_\nu (\sqrt{g}t_\mu ^\nu )$$ (17) hence the equations of motion can also be written as $$\sigma du_\mu /ds=(\sqrt{g})^1_\nu (\sqrt{g}t_\mu ^\nu ).$$ (18) This equation shows clearly the essential point that the $`t_\mu ^\nu `$ is the carrier (mediator) of interactions and in its absence there will be no accelerations. This is what we mean with the requirement of “interactive N-body” solutions and here we see that $`\lambda =1`$ theory has them. Multiplying by $`\sqrt{g}`$ and integrating over the volume containing one of the particles, for example $`m_1=m`$, $$mdu_\mu /ds=\mathrm{\Delta }\varphi _\mu \varphi =m_\mu \varphi $$ (19) which is the equation of motion in the slow motion limit. Since, upon calculation the term $`1/2_\mu g_{\alpha \beta }(\tau ^{\alpha \beta }+t^{\alpha \beta })`$ gives the same result (note that in this limit $`_\mu g_{\alpha \beta }t^{\alpha \beta }=0`$), we have the geodesic limit $`(du_\mu /ds=_\mu \varphi )`$ and the strong principle of equivalence satisfied (Yılmaz (1992)) $$m_i=m_a=m_p$$ (20) since $`\mathrm{\Delta }\varphi =\sqrt{g}\sigma `$ is the density of “active mass”. This calculation also shows that it will be difficult, if not impossible, to satisfy the “strong principle of equivalence” without the $`t_\mu ^\nu `$ because the active mass comes in by the density divergence of $`t_\mu ^\nu `$. The theory describes interactive multiparticle dynamics in the sense of Hamiltonian particle mechanics; the continuum limit is allowed by statistical averaging, in which case one needs two or more functions to describe the details of the equation of state. Can there be noninteractive N-body solutions? It is found that in some simple symmetries, extended bodies such as parallel slabs and spherical shells, there may be N of them even when $`t_\mu ^\nu `$ is zero. However, by the above equations of motion Eq. (18), the forces between them, hence also their accelerations are zero (Alley (1994)). If the $`t_\mu ^\nu `$ is present, they do interact (consistent with the Newtonian correspondence). These results can be verified by hand or by computer calculations. If the solution contains only one object, then, of course, there cannot be any interaction as there would be nothing else to interact with. As to the test-body theories having a single central body plus test particles put by hand, they contain an implicit assumption, namely, the central body must have infinite inertial mass and finite active mass which we know is false and is against the strong principle of equivalence. Of course, particles put by hand cannot have active mass and cannot generate gravitational fields. Such particles are called test-particles. A test-particle theory violates the universal interparticle symmetry of gravitation because the central body is in the solution but the test-particles are not (Yılmaz (1988)). The difference between an N-body theory and a test-body theory shows up most dramatically in the calculation of the motions of the planetary perihelia. Thus for example perihelion of Mercury advances 575” per century of which 532” is due to Mercury’s interactions with other planets and 43” per century to relativistic correction. The 532” interactive part is predicted by the N-body theory but not by the test-body theory since test bodies do not interact. The situation is the same for the other eight planets all of which have even larger interactive perihelion shifts. The $`\lambda =1`$ theory predicts the total perihelion motions in a seamless way. It is usually believed that in papers published in 1938 and 1940 (Einstein, Infeld & Hoffman (1938); Einstein & Infeld (1940)) Einstein, Infeld and Hoffman (EIH) obtained N-body equations of motion in the slow motion limit. This belief is unfounded. As described by P. G. Bergmann in his well-known book (Bergmann (1942)), the situation is as follows: With Eqs. (15.12) on page 230, Einstein’s equations are satisfied in first order (right hand sides are put to zero in vacuum), but with Eqs. (15.25) on page 234 they are not satisfied in second order (they are not put to zero in vacuum). They are left unspecified. Yet, as stated on page 232, to obtain the equations of motion one must carry the field equations to second order. Thus the question arises: What should these unspecified second order terms be in order to get the N-body interactive solutions to be used later to obtain the N-body equations of motion (15.49) on page 240? It turns out that they cannot be zero, as Einstein’s theory requires. They rather demand $`\frac{1}{2}G_\mu ^\nu =t_\mu ^\nu `$ in vacuum where the $`t_\mu ^\nu `$ is the Yilmaz stress-energy tensor for the N-body field $`\varphi =\mathrm{\Sigma }_Am_A/r_A+C`$. (The (-) sign is due to the definition of $`G_\mu ^\nu `$ in Bergmann’s book as the negative of Yilmaz’ definition). In other words, Eqs.(15.49) are true in Yilmaz’ theory and not in Einstein’s theory. In fact, the Yilmaz exponential metric, our eq.(9), can be derived from the condition that, in the Newtonian limit, the equations of motion will be of the form (15.49) of Bergmann. ## 5 Discussion The recent discovery of the gamma ray burster GRB 990123, requiring energies exceeding the limit allowable by general relativity for neutron star mergers, created an energy crisis in astrophysics (Schilling (1999)). The limiting factor seems to be that, according to general relativity, a neutron star (or a merger of stars) exceeding a total of 2.8$`M_{\mathrm{}}`$ would become a black hole and thereafter little radiation could escape whereas the energy required for GRB 990123 seems to be at least 2$`M_{\mathrm{}}c^2`$ to properly account for the gamma and other emissions. In fact, according to the N-body theorem there cannot be such energy producing mergers in general relativity. If the obstacle event horizon did not exist, the interaction energy released from the deeper regions, surfaces, magnetic fields, etc., can provide the required energy. (Note that the massive neutron stars can possess magnetic moments – the ”black holes have no hair” theorem does not apply in the new theory. Note also that radially directed light can always escape, although substantially redshifted.) In two recent articles by S. L. Robertson such an explanation is already proposed (Robertson 1999a ; Robertson 1999b ). Summarizing: a) The long sought N-body interactive solutions in curved spacetime theory of gravity are found which merits immediate attention in its own right. b) The test-body (1-body) nondynamical metrics are replaced by N-body dynamical metrics free of event horizons. A natural explanation of the GRB 990123 energy requirement becomes possible via a merger of two massive neutron stars (called Yılmaz stars by Robertson) which are not black holes. In the past, in times of great theoretical and observational crises, like the ones we are now having, patching up old theories did not help. Instead, a new paradigm emerged which organized known facts in a more systematic manner as well as overcoming the prevalent difficulties and predicting new effects. We may be witnessing here a similar situation in the equations of general relativity. In both the field equations and the equations of motion, the “matter alone” paradigm is allowed to go over into a new paradigm “matter plus field”. (More precisely, $`\tau _\mu ^\nu \tau _\mu ^\nu +t_\mu ^\nu `$.) This change in paradigm makes it possible to treat the GRB 990123 as a merger or collision of two massive neutron stars, with some beaming if needed, whereas general relativity seems to be in a bind, since it has only a 1-body solution (a solitary black hole) with which none of these models is feasible. Quite independently of the energy crisis at hand this shift in paradigm has many important consequences in other respects. a) The theory becomes a standard local gauge-field theory in curved spacetime. b) It is a dynamical theory (not a test-body theory), hence the planetary perturbations are treatable in a seamless way along with the relativistic effects. c) It does not lead to event horizons, hence physical properties such as magnetic moments are allowed. d) It has a higher critical mass, hence more energy is available in mergers and collisions. e) As far as we know, it is quantizable (Yılmaz 1997, Alley 1995). These and other important features will be described in a larger paper in preparation. “The hallmark of a successful theory is that it predicts correctly facts which were not known when the theory was presented or, better still, which were then known incorrectly.” Francis Crick (Life Itself, Simon and Schuster, 1981) We thank Dr. Ching Yun Ren, Mr. Kirk Burrows and Mr. Per Kennet Aschan for their expert help in computer calculations. Two of us (HY and YM) would like to thank Mr. T. Hiruma, President of Hamamatsu Photonics, K.K. for encouragement and dedicate our efforts to the memory of the late Executive Vice President of Hamamatsu Photonics, Dr. Sakio Suzuki.
no-problem/9906/astro-ph9906143.html
ar5iv
text
# 1 INTRODUCTION ## 1 INTRODUCTION Several remarkable discoveries have renewed interest in solar system formation. Recent surveys have detected many small icy bodies beyond the orbit of Neptune (e.g., Jewitt & Luu (1993); Williams et al. (1995); Jewitt et al. (1996); Luu et al. (1997); Gladman & Kavelaars (1997); Chiang & Brown (1999)). Assuming a geometric albedo of 4%, these Kuiper Belt objects (KBOs) have radii of 50–400 km; the derived size distribution implies a significant population of smaller objects. With semi-major axes of 40–50 AU and orbital inclinations of 0–30, the orbits of known KBOs suggest an annulus of planetesimals formed in situ and left over from the planetary formation epoch (Holman & Wisdom (1993)). The presumed structure of this annulus resembles the dusty disks recently discovered around several nearby stars (Smith & Terrile (1984); Aumann et al. (1984); Jayawardhana et al. (1998); Koerner et al. (1998); Greaves et al. (1998)). Planets similar to those in our solar system have not been detected in any of these disks, but direct images and radial velocity measurements of other nearby stars already imply the existence of more than one dozen extra-solar planets of several Jupiter masses (Latham et al. (1989); Marcy & Butler (1996); Cochran et al. (1997); Noyes et al. (1997); DelFosse et al. (1998); for a review see Marcy (1999)). These discoveries challenge planet formation theories. Most theories presume that planets grow by accretion of small planetesimals in a gaseous circumstellar disk (Safronov (1969); Goldreich & Ward (1973); see also Lissauer & Stewart (1993), Boss (1997), and references therein). Hitherto, numerical studies have focused on the formation of the prototypical terrestrial and gas giant planets, Earth and Jupiter (Greenberg et al. (1978), 1984; Nakagawa et al. (1983); Ohtsuki et al. (1988); Wetherill & Stewart (1989), 1993; Barge & Pellat (1990); Ruden & Pollack (1991); Pollack et al. (1996); Weidenschilling et al. (1997)). If the initial disk mass is comparable to the Minimum Mass Solar Nebula<sup>1</sup><sup>1</sup>1 The Minimum Mass Solar Nebula has a surface density $`\mathrm{\Sigma }=\mathrm{\Sigma }_0(R/R_0)^{3/2}`$, where $`\mathrm{\Sigma }_0`$ is the surface density of solid material at $`R_0`$ = 1 AU. We adopt $`\mathrm{\Sigma }_0`$ = 45 g cm<sup>-2</sup> (Hayashi (1981); see also Weidenschilling (1977); Bailey (1994)). This definition yields a total mass of solids $`M_0`$ 100 $`M_E`$ inside the orbit of Neptune ($`R<`$ 30 AU) and $`M_0`$ 10 $`M_E`$ in the inner part of the Kuiper Belt ($`R`$ = 32–38 AU), where 1 $`M_E=6\times 10^{27}`$ g is the mass of the Earth., these calculations often have difficulty producing objects similar to the known terrestrial or gas giant planets during the estimated disk lifetime of $``$ 10–30 Myr (Pollack et al. (1996); Weidenschilling et al. (1997)). This problem is exacerbated in the outer solar system, where numerical calculations yield formation times exceeding 100 Myr for 500–1000 km radius KBOs (Stern (1995), 1996; Stern & Colwell 1997a ,b). KBOs must form on shorter timescales in parallel with Neptune. Otherwise, Neptune’s gravity increases the velocities of nearby planetesimals, including those in the inner Kuiper Belt, on timescales of $``$ 10 Myr (Malhotra (1996)). This process prevents the growth of KBOs with radii exceeding 100–200 km, because large velocities hinder agglomeration. We recently began to consider KBO formation in the outer solar system using an evolution code that follows planetesimal growth in the annulus of a circumstellar disk. Initial results indicate that KBOs can form at 30–50 AU on timescales of 10–100 Myr in disks with 1–3 times the Minimum Mass Solar Nebula when collisional disruption of planetesimals is unimportant (Kenyon & Luu 1998; hereafter KL (98)). Further calculations with an algorithm that includes disruptive processes lead to similarly short timescales for a wide range of initial conditions (Kenyon & Luu 1999; hereafter KL (99)). Here, we briefly summarize these new results, compare the theoretical model with current observations, and make predictions for comparison with future observations of KBOs. ## 2 MODEL Our accretion code is based on the particle-in-a-box method, where planetesimals are a statistical ensemble of bodies with a distribution of horizontal and vertical velocities about Keplerian orbits (Safronov (1969)). We perform calculations for a single annulus of width $`\mathrm{\Delta }a`$ centered at a heliocentric distance $`a`$. We approximate the continuous distribution of particle masses with $`i`$ discrete batches having particle populations $`n_i(t)`$ and total masses $`M_i(t)`$. The horizontal and vertical velocity dispersions are $`h_i(t)`$ and $`v_i(t)`$ (Wetherill & Stewart (1993)). The average mass of a batch, $`m_i(t)`$ = $`M_i(t)/n_i(t)`$, changes with time as collisions add and remove bodies from the batch. This procedure conserves mass and provides a statistical method to follow the growth of $``$ $`10^{20}`$ small planetesimals into a few planets. Detailed $`n`$-body calculations confirm the basic features of particle-in-a-box calculations for the early stages of planet growth described here (Ida & Makino (1992); Kokubo & Ida (1996)). To evolve the initial size distribution in time, we calculate collision rates for the coagulation equation, determine the outcome of each collision, and compute velocity changes due to collisions and long-range gravitational interactions (see KL99). Each two-body collision can produce (1) merger into a single body with no escaping debris (very low impact velocity), (2) merger into a single body with escaping debris (‘cratering’; low impact velocity), (3) rebound with or without cratering (modest impact velocity), or (4) catastrophic disruption into numerous smaller bodies (high impact velocity). The collision outcomes depend on the ratio of the impact energy $`Q_f`$ to the disruption energy $`Q_d`$ of two colliding planetesimals (Greenberg et al. (1978); Wetherill & Stewart (1993); Davis et al. (1994)). Collisions with $`Q_f>Q_d`$ disrupt planetesimals into many small fragments. Collisions with $`Q_f<Q_d`$ yield a merged planetesimal and some small fragments if the collision velocity $`V_c`$ exceeds the minimum velocity for cratering $`V_f`$. Collisions with $`Q_f<Q_d`$ and $`V_c<V_f`$ yield a merged planetesimal with no cratering debris. We use an energy-scaling formalism to compute $`Q_d`$ as the sum of the intrinsic tensile strength $`S_0`$ and the gravitational binding energy (Davis et al. (1985), 1994). The intrinsic strength is the dominant component of $`Q_d`$ for bodies with $`r_i`$ 1 km; gravitational binding dominates $`S_0`$ for larger bodies. For each collision, a velocity evolution algorithm distributes the kinetic energy among the resulting bodies and then accounts for collisional damping, kinetic energy transfer during elastic collisions (“dynamical friction”), angular momentum transfer during elastic collisions (“viscous stirring”), and gas drag (Hornung et al. (1985); see also Wetherill & Stewart (1993); KL (98)). Dynamical friction tries to enforce equipartition of kinetic energy between mass batches; viscous stirring increases the velocities of all bodies. Gas drag removes objects from the annulus and reduces the velocities of small objects which are well coupled to gas in the disk. We tested the code against analytical solutions and published numerical results (KL (98), KL (99)). We reproduced previous calculations for accretion at 1 AU (Wetherill & Stewart (1993)) and collisional disruption of pre-existing large KBOs at 40 AU (Davis & Farinella (1997)). Our calculations match analytical solutions well when the mass spacing between successive batches, $`\delta `$ = $`m_{i+1}/m_i`$ = 1.1–1.4. Numerical solutions lag the analytic results by $``$ 10% when $`\delta `$ = 1.4–2. The timescale to produce objects of a given size increases with $`\delta `$, because poorer resolution prevents growth of large objects (see KL (98)). Table 1 lists basic input parameters. The input cumulative size distribution $`N_C`$ has the form $`N_Cr_i^{q_0}`$, with initial radii $`r_i`$ = 1–80 m. The total mass in the annulus is $`M_0`$; $`M_0`$ 10 $`M_E`$ for a Minimum Mass Solar Nebula. All batches start with the same initial velocity. We tested a range of initial velocities corresponding to initial eccentricities of $`e_0=10^4`$ to $`10^2`$, as is expected for planetesimals in the early solar nebula (Malhotra (1995)). The adopted mass density, $`\rho _0`$ = 1.5 g cm<sup>-3</sup>, is appropriate for icy bodies with a small rocky component. The fragmentation parameters – $`V_f`$, $`S_0`$, $`Q_c`$, $`f_{KE}`$, $`c_1`$, and $`c_2`$ – are adopted from earlier work. KL99 describe these parameters in more detail. To provide observational constraints on the models, we note that the known Kuiper Belt population contains at least one body with a radius of $``$ 1000 km (Pluto), and $`10^5`$ KBOs with radii $`r_i`$ 50 km between 30–50 AU. The cumulative size distribution of known KBOs can be fitted with $`N_Cr_i^{q_{obs}}`$, with $`q_{obs}=3\pm 0.5`$ (Jewitt et al. (1998); see also Chiang & Brown (1999)). Successful models should reproduce these observations on timescales comparable to (a) the estimated lifetimes of the solar nebula and gaseous disks surrounding nearby young stars, $`10^7`$ yr (Russell et al. (1996); Hartmann et al. (1998)) and (b) the formation timescale for Neptune, $`10^8`$ yr (Lissauer et al. (1996)). ## 3 NUMERICAL RESULTS We separate the growth of KBOs into three regimes. Early in the evolution, frequent collisions damp the velocity dispersions of small bodies. These bodies slowly grow into 1 km objects on a timescale that is approximated by $`\tau _{1\mathrm{km}}`$ 8 Myr $`(M_0/10M_E)(e_0/10^3)^{0.65}`$. This linear growth phase ends when the gravitational range of the largest objects exceeds their geometric cross-section. This “gravitational focusing” enhances the collision rate by factors of 10–1000. The largest objects then begin a period of “runaway growth”, when their radii grow from $``$ 1 km to $``$ 100 km in several Myr. During this phase, dynamical friction and viscous stirring increase the velocity dispersions of the smallest bodies from $``$ 1 m s<sup>-1</sup> up to $``$ 40 m s<sup>-1</sup>. This velocity evolution reduces gravitational focusing factors and ends runaway growth. The largest objects then grow slowly to 1000+ km sizes on timescales that again depend on $`M_0`$ and $`e_0`$. Column (5) in Table 2 lists timescales to form Pluto-size objects $`\tau _P`$ as a function of the input parameters $`M_0`$, $`\delta `$, $`e_0`$, and $`q_0`$. Fig. 1 shows cumulative size distributions for a model with $`M_0=10M_E`$, $`q_0`$ = 3, and $`S_0=2\times 10^6`$ erg g<sup>-1</sup>. The shapes of these curves depend on two competing physical processes: (1) growth by mergers and (2) erosion by high velocity collisions. In this example, collisions result in growth because the velocity dispersion is less than the catastrophic disruption threshold. However, the collision velocity exceeds the cratering threshold $`V_f`$. Cratering adds debris to all low mass batches. Gas drag removes material from low mass batches ($`r_i`$ 10 m), but is ineffective at removing larger objects. The size distribution thus becomes shallower at small masses. At large masses, mergers produce a group of growing planetesimals with a steep size distribution. Once gravitational focusing becomes effective, the largest of these objects ‘run away’ from the rest of the ensemble to produce a smooth power law with a maximum radius $`r_{max}`$. As the evolution proceeds, $`r_{max}`$ increases but the slope of the smooth power law remains nearly constant. The main features of these results depend little on the input parameters. All calculations produce two cumulative power law size distributions connected by a transition region having an ‘excess’ of planetesimals (the “bump” in the curves in Fig. 1). The characteristic radius of this transition region increases from 0.3 km at $`e_0=10^4`$ to 3 km at $`e_0=10^2`$. If fitted with a power law of the form $`N_Cr_i^{q_f}`$ at small masses, the cumulative size distribution follows the predicted limit for collisional evolution, $`q_f=2.5`$ (Dohnanyi (1969)). We perform least-square fits to obtain $`q_f`$ at larger masses; column (6) of Table 2 lists derived values for $`q_f`$ along with the 1$`\sigma `$ error. Column (7) lists the radius range for each fit. The results are surprisingly independent of the input parameters. We find the small range $`q_f`$ = 2.75–3.25 for calculations with $`M_0`$ = 1–100 $`M_E`$, $`e_0=10^4`$ to $`10^2`$, $`q_0`$ = 1.5–4.5, and $`S_0`$ = 10 erg g<sup>-1</sup> to $`3\times 10^6`$ erg g<sup>-1</sup>. This model result is consistent with the observed slope, $`q_{obs}=3\pm 0.5`$ (e.g., Jewitt et al. 1998). ## 4 COMPARISONS WITH OBSERVATIONS As shown in Table 2, several calculations meet the success criteria defined in $`\mathrm{\S }2`$. Annuli with $`M_010M_E`$ produce Pluto-sized objects on short timescales, $`\tau _P`$ 50 Myr (for $`e_010^3`$). Models with smaller initial masses or larger initial eccentricities form Plutos on longer timescales, $`\tau _P`$ 50 Myr. Plausible ranges of other input parameters – such as $`q_0`$, $`S_0`$, and $`f_{KE}`$ – yield $`\pm `$20% variations about these timescales. The results are insensitive to $`V_f`$ and other collision parameters (KL (99)). The crosses in Figure 1 compare our calculations directly with several observational constraints. The cross at $`r_i`$ = 50 km indicates the number of KBOs with $`r_i`$ 50 km estimated from recent ground-based surveys (Jewitt et al. (1998); see also Chiang & Brown (1999)); the one at $`r_i`$ = 10 km shows limits derived from a single, controversial measurement with Hubble Space Telescope (HST; Cochran et al. (1995), 1998; Brown et al. (1997)). The third cross plots limits at $`r_i`$ = 1 km derived from theoretical attempts to explain the frequency of short-period comets from the Kuiper Belt (Davis & Farinella (1997); Duncan & Levison (1997); Levison & Duncan (1997)). Our predictions agree with ground-based surveys at 50 km and theoretical limits at 1 km, but fall a factor of $``$ 10 short of the HST measurement at 10 km. To compare with observations in more detail, we predict the luminosity function (LF) of KBOs directly from the computed number distribution. We use a Monte Carlo calculation of objects selected randomly from the cumulative size distribution $`N_C`$. We assign each object a distance from the Sun $`d_{\mathrm{}}`$ and a random phase angle $`\beta `$ between the line-of-sight from the Earth to the object and the line-of-sight from the Sun to the object. This phase angle lies between 0 and a maximum phase angle that is distance-dependent. The distance of the object from the Earth is then $`d_E=d_{\mathrm{}}\mathrm{cos}\beta (1+d_{\mathrm{}}^2(\mathrm{cos}^2\beta 1))^{1/2}`$. We derive the red magnitude of this object from a two parameter magnitude relation for asteroids, $`m_{R,KBO}=R_0+2.5\mathrm{log}(t_1/t_2)5\mathrm{log}r_{KBO}`$, where $`R_0`$ is the zero point of the magnitude scale, $`r_{KBO}`$ is the radius of the KBO, $`t_1=2d_{\mathrm{}}d_E`$, and $`t_2=\omega ((1g)\varphi _1+g\varphi _2)`$ (Bowell et al. (1989)). In this last expression, $`\omega `$ is the albedo, and $`g`$ is the slope parameter; $`\varphi _1`$ and $`\varphi _2`$ are phase functions that describe the visibility of the illuminated hemisphere of the object as a function of $`\beta `$. We adopt standard values, $`\omega =0.04`$ and $`g=0.15`$, appropriate for comet nuclei (Jewitt et al. (1998)). The zero point $`R_0`$ is the apparent red magnitude of the Sun, $`m_{R,\mathrm{}}`$ = $``$27.11, with a correction for the V–R color of a KBO, $`R_0`$ = $`m_{R,\mathrm{}}`$ \+ $`\delta `$(V–R)<sub>KBO</sub>. Observations suggest that KBOs have colors that range from roughly $`0.1`$ to 0.3 mag redder than the Sun. We treat this uncertainty by allowing the color to vary randomly in this range. The important parameters in the model LF are the distributions of input sizes (derived from the accretion calculations), distances, and orbital parameters. We assume KBOs are evenly distributed between “Plutinos,” objects in 3:2 orbital resonance with Neptune having semimajor axes of 39.4$`\pm `$0.2 AU, and “classical” KBOs with semimajor axes between 42–50 AU. The distance parameters are set by observations (Jewitt et al. (1998)). This distance distribution is different from the 32–38 AU adopted for the coagulation calculations. Several tests show that accretion results at 42–50 AU are identical to those at 32–38 AU, except that the timescale to produce Pluto-sized objects is 50%–100% longer. To compute the model LF from the Monte Carlo magnitude distribution of KBOs, we scale the mass in the 32–38 AU annulus to match the mass in a 42–50 AU annulus, add in an equal number of Plutinos, and divide by the sky area. The distribution of KBO orbital parameters is poorly known. We adopt circular orbits to derive magnitudes; the LF is insensitive to other choices. We assume orbital inclinations of $`i=`$ 0 to 5 to compute the sky area, which is a compromise between the $`i`$ 0–5 of classical KBOs and the $`i`$ 10–30 of Plutinos. The model LFs scale inversely with sin $`i`$. Figure 2 compares several models with the observed LF. The left panel shows models with $`e_0`$ = $`10^3`$ and different masses; the right panel shows models with the mass of a Minimum Mass Solar Nebula and different $`e_0`$. Model LFs with the Minimum Mass and any $`e_0`$ agree with current observations. The good agreement of all models at $`m_R`$ 20, where the uncertainties are largest, depends on the assumed maximum radius in the model distribution. We picked 1000 km for convenience. Model LFs for $`m_R`$ 20 are independent of this choice. To quantify the comparison between models and observations, we fit model LFs to log $`\mathrm{\Sigma }(m_R)`$ = $`\alpha (m_Rm_0)`$ over $`20.5m_R26.5`$. Table 3 lists the fitted $`\alpha `$ and $`m_0`$ as a function of the mass in classical KBOs (in units of the Minimum Mass Solar Nebula), $`e_0`$, the inner annulus boundary $`R_{in}`$, and the outer annulus boundary $`R_{out}`$. The small range in $`\alpha `$ for model LFs agrees with published values derived from observations<sup>2</sup><sup>2</sup>2Gladman et al. (1998) report $`\alpha =0.76_{0.11}^{+0.10}`$ and $`m_0=23.4_{0.18}^{+0.20}`$ from a maximum likelihood analysis of previous surveys with magnitude limits, 20 $`m_R`$ 28. Surface densities in their Table 3 yield $`\alpha 0.6`$ and $`m_0`$ 22.4. Jewitt et al. (1998; see also Luu & Jewitt 1998) quote $`\alpha =0.54\pm 0.04`$ and $`m_0=23.2\pm 0.10`$ for 20 $`m_R`$ 26. Chiang & Brown (1999) prefer $`\alpha =0.52\pm 0.02`$ and $`m_0`$ = 23.5 for 20 $`m_R`$ 27; they note that the slope depends on which survey data are used in the fit.. The model $`\alpha `$ is independent of the relative numbers of Plutinos and classical KBOs, and the distance distribution of classical KBOs. The observed zero-point of the LF, $`m_0`$ 23.2–23.5, favors models with masses comparable to the Minimum Mass Solar Nebula and any initial eccentricity. These data rule out models with $`30\%`$ of the Minimum Mass at the 3$`\sigma `$ level. Smaller Plutino fractions require larger masses: if Plutinos are 10%–25% of the total KBO population, as indicated by recent observations (Jewitt et al. (1998)), the needed mass is 2–4 times the Minimum Mass. There are two main uncertainties in comparing our model LFs with the data, the evolution of the KBO LF with time and the current orbital parameters of KBOs. The initial mass in KBOs was larger than implied by a direct comparison between the data and model LFs, because large velocity collisions and dynamical encounters with Neptune have eroded the Kuiper Belt over time (Holman & Wisdom (1993); Davis & Farinella (1997); see also Levison & Duncan (1993); Duncan et al. (1995)). Erosion from high velocity collisions probably does not change the slope of the LF significantly. Massive KBOs with $`r_i`$ 50 km ($`m_R`$ 26–27) are probably safe from collisional disruption (Davis & Farinella (1997)). Disruption of smaller bodies depends on the unknown bulk properties and the poorly known orbital parameters of KBOs. These uncertainties are not important for comparisons of models and observations for $`m_R`$ 26–27, but can bias future comparisons at fainter magnitude limits. Gravitational perturbations from Neptune should affect all KBO masses equally and simply reduce the total mass in KBOs with time (Holman & Wisdom (1993)). Despite the uncertainty in the total amount of mass lost from the Kuiper Belt, we are encouraged that the mass needed to explain current observations of KBOs is at least the Minimum Mass Solar Nebula. Future calculations will allow us to place better constraints on the initial mass in the Kuiper Belt. The uncertain distribution of KBO orbital parameters also affects the initial mass estimates. Our assumption of KBOs uniformly distributed in distance $`d_{\mathrm{}}`$, orbital eccentricity $`e`$, and inclination $`i`$ is probably incorrect for Plutinos in specific orbital resonances with Neptune. Larger adopted volumes for current Plutinos require larger initial disk masses in the Kuiper Belt. A uniform distribution is probably reasonable for classical KBOs, but the observed range in $`d_{\mathrm{}}`$ and $`i`$ is not well-known. Allowing classical KBOs to occupy a larger range in semi-major axis reduces our mass estimates; a larger range in sin $`i`$ increases our mass estimates. We suspect that the uncertainties currently are a factor of $``$ 2–3. Future large area surveys will provide better knowledge of KBO orbital parameters and allow more accurate models for the observed LF. In addition to the reasonably good fit for $`20.5m_R26.5`$, our calculations predict 1–5 ‘Plutos’ with $`m_R`$ 20 over the entire sky. This number is uncertain, because we do not understand completely the mechanism that ends accretion and sets the maximum size of KBOs. Our calculations indicate that planetary accretion at 35–50 AU is self-limiting: once objects reach radii of $``$ 1000 km, they stir up smaller bodies sufficiently to limit additional growth. The formation of nearby Neptune should have also limited the growth of the largest bodies (Morbidelli & Valsecchi (1997)). Better constraints on the radial distribution of 500+ km KBOs would test the relative importance of these two mechanisms. Observations at fainter magnitude limits will provide additional constraints on KBO formation. Imaging data acquired at the Keck and Palomar telescopes detect KBOs with $`m_R`$ 25–26.5, where models with $`e_010^2`$ predict the LF to rise sharply. The apparent lack of a significant upturn in the LF at $`m_R`$ 25 implies $`e_0`$ a few $`\times 10^2`$. In contrast, the current limit on the KBO population at $`m_R`$ 28 implies a substantial population of 10 km radius KBOs which is inconsistent with our calculations. Deeper ground-based surveys or new HST data could resolve the controversy surrounding this observation and place better constraints on $`e_0`$. Finally, the proposed Next Generation Space Telescope (NGST) will probe the size distribution of 1 km radius KBOs where models with $`e_010^3`$ predict a sharp upturn in the observed LF. If such small bodies can survive for the age of the solar system, NGST observations would provide important constraints on the initial mass and dynamics of the outer solar system. We thank B. Bromley for making it possible to run our code on the JPL Cray T3D ‘Cosmos’ and the HP Exemplar ‘Neptune’ and for a generous allotment of computer time through funding from the NASA Offices of Mission to Planet Earth, Aeronautics, and Space Science. Comments from F. Franklin, M. Geller, and M. Holman greatly improved our presentation. | Table 1. Basic Model Parameters | | | | | --- | --- | --- | --- | | Parameter | Symbol | | Value | | Width of annulus | $`\delta a`$ | | 6 AU | | Initial velocity | $`V_0`$ | | 0.45–45 m s<sup>-1</sup> | | Particle mass density | $`\rho _0`$ | | 1.5 g cm<sup>-3</sup> | | Relative gas velocity | $`\eta `$ | | 30 m s<sup>-1</sup> | | Time step | $`\delta t`$ | | 1–250 yr | | Number of mass bins | $`N`$ | | 64–256 | | Mass spacing of bins | $`\delta `$ | | 1.25-2.0 | | Minimum velocity for cratering | $`V_f`$ | | 1 cm s<sup>-1</sup> | | Impact strength | $`S_0`$ | | $`2\times 10^6`$ erg g<sup>-1</sup> | | Crushing energy | $`Q_c`$ | | $`5\times 10^7`$ erg g<sup>-1</sup> | | Fraction of KE in ejecta | $`f_{KE}`$ | | 0.05 | | Coefficient of restitution | $`c_1`$ | | $`10^2`$ | | Coefficient of restitution | $`c_2`$ | | $`10^3`$ | | Table 2. Model Results at 32–38 AU | | | | | | | | --- | --- | --- | --- | --- | --- | --- | | $`M_0(M_E)`$ | $`\delta `$ | $`e_0`$ | $`q_0`$ | $`\tau _P`$ (Myr) | $`q_f`$ | Range (km) | | 1 | 1.4 | $`10^4`$ | $`3.0`$ | 448 | $`2.86\pm 0.04`$ | 2–300 | | 10 | 1.4 | $`10^4`$ | $`3.0`$ | 20 | $`2.90\pm 0.02`$ | 1–930 | | 1 | 1.4 | $`10^3`$ | $`3.0`$ | 893 | $`2.76\pm 0.05`$ | 6–400 | | 3 | 1.4 | $`10^3`$ | $`3.0`$ | 184 | $`2.78\pm 0.03`$ | 5–600 | | 10 | 1.4 | $`10^3`$ | $`3.0`$ | 37 | $`2.91\pm 0.03`$ | 6–800 | | 30 | 1.4 | $`10^3`$ | $`3.0`$ | 10 | $`3.02\pm 0.03`$ | 6–930 | | 100 | 1.4 | $`10^3`$ | $`3.0`$ | 3 | $`2.97\pm 0.03`$ | 7–600 | | 10 | 1.4 | $`10^2`$ | $`3.0`$ | 428 | $`3.15\pm 0.10`$ | 50–700 | | 100 | 1.4 | $`10^2`$ | $`3.0`$ | 25 | $`3.23\pm 0.06`$ | 20–600 | | 10 | 1.25 | $`10^3`$ | $`1.5`$ | 40 | $`2.97\pm 0.02`$ | 7–700 | | 10 | 1.25 | $`10^3`$ | $`3.0`$ | 35 | $`3.03\pm 0.03`$ | 9–650 | | 10 | 1.25 | $`10^3`$ | $`4.5`$ | 30 | $`2.90\pm 0.02`$ | 4–750 | | Table 3. Luminosity Function Parameters | | | | | | | | --- | --- | --- | --- | --- | --- | --- | | $`M_0/M_{MMSN}`$ | $`e_0`$ | $`R_{in}`$ (AU) | $`R_{out}`$ (AU) | | $`\alpha `$ | $`m_0`$ | | 0.3 | $`10^3`$ | 42 | 50 | | 0.56 $`\pm `$ 0.01 | 24.03 $`\pm `$ 0.16 | | 1.0 | $`10^4`$ | 42 | 50 | | 0.58 $`\pm `$ 0.01 | 23.36 $`\pm `$ 0.12 | | 1.0 | $`10^3`$ | 42 | 50 | | 0.57 $`\pm `$ 0.02 | 23.16 $`\pm `$ 0.15 | | 1.0 | $`10^2`$ | 42 | 50 | | 0.60 $`\pm `$ 0.01 | 23.53 $`\pm `$ 0.18 | | 3.0 | $`10^3`$ | 42 | 50 | | 0.58 $`\pm `$ 0.01 | 22.42 $`\pm `$ 0.17 | | 0.3 | $`10^3`$ | 42 | 60 | | 0.56 $`\pm `$ 0.01 | 23.50 $`\pm `$ 0.13 | | 1.0 | $`10^4`$ | 42 | 60 | | 0.57 $`\pm `$ 0.01 | 22.85 $`\pm `$ 0.11 | | 1.0 | $`10^3`$ | 42 | 60 | | 0.56 $`\pm `$ 0.02 | 22.63 $`\pm `$ 0.13 | | 1.0 | $`10^2`$ | 42 | 60 | | 0.63 $`\pm `$ 0.02 | 23.31 $`\pm `$ 0.19 | | 3.0 | $`10^3`$ | 42 | 60 | | 0.58 $`\pm `$ 0.01 | 21.92 $`\pm `$ 0.17 |
no-problem/9906/cond-mat9906050.html
ar5iv
text
# 1 Number of excitations, 𝑁, versus frequency, Λ, found numerically from Eq. () for 𝑠=∞ (full), 4 (dotted), 3 (short-dashed), 2.5 (long-dashed), 2 (short-long-dashed), 1.9 (dashed-dotted). Functional Materials 5, No. 3 (1998) 315–318. EFFECTS OF NOISE AND NONLOCAL INTERACTIONS IN NONLINEAR DYNAMICS OF MOLECULAR SYSTEMS P.L. Christiansen<sup>1</sup>, G.I. Gaididei<sup>2</sup>, Yu.B. Gaididei<sup>3</sup>, M. Johansson<sup>4</sup>, S.F. Mingaleev<sup>3</sup>, and K.Ø. Rasmussen<sup>5</sup> <sup>1</sup>Department of Mathematical Modelling, Technical University of Denmark, DK-2800 Lyngby, Denmark <sup>2</sup>Physics Department of Kyiv University, 252022, Kyiv, Ukraine <sup>3</sup>Bogolyubov Institute for Theoretical Physics, 252 143 Kyiv, Ukraine <sup>4</sup>Department of Physics and Measurement Technology, Linköping University, S-581 83 Linköping, Sweden <sup>5</sup>Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA We show that the NLS systems with multiplicative noise, nonlinear damping and nonlocal dispersion exhibit a variety of interesting effects which may be useful for modelling the dynamical behavior of one- and two-dimensional systems. 1 INTRODUCTION We consider the nonlinear Schrödinger (NLS) system with cubic nonlinearity. This system models optical media, molecular thin films in the continuum limit, deep water waves and many other physical systems which exhibit weak nonlinearity and strong dispersion. The effects we shall discuss are observed in our earlier works and cited therein. The motivation for studying the 2-D NLS equation with thermal fluctuations is our intension to model efficient energy transfer in Scheibe-aggregates. Our starting point is a two-dimensional Davydov model with nonlinear coupling between the exciton and phonon system, and white noise in the phonon system. In Section 2 we derive a single equation for the exciton system with multiplicative colored noise and a nonlinear damping term. In the continuum limit the collective coordinate approach indicates that an energy balance between energy input (from the noise term) and dissipation can be established. Thus, this model may describe the state of thermal equilibrium in the molecular aggregate. The coherent exciton moving on the aggregate is modelled by the ground state solution to the 2-D NLS equation, and the lifetime has been related to the collapse time of the ground state . For sufficiently strong nonlinearity the thermal fluctuations will slow down the collapse. As a result this lifetime increases with the variance of the fluctuations, i.e. the temperature. In Section 3 the role of long-range dispersive interaction in the one-dimensional molecular system is investigated. Dispersive interactions of two types (the power and the exponential dependences of the interaction intensity on the distance) are studied. If the intaraction decreases with the distance slowly, there is an interval of bistability where two stable stationary states: narrow, pinned states and broad, mobile states exist at each value of the excitation energy. For cubic nonlinearity the bistability of the solitons occurs already for dipole-dipole dispersive interaction. We demonstrate a possibility of the controlled switching between pinned and mobile states applying a spatially symmetric perturbation in the form of a parametric kick. The mechanism could be important for controlling energy storage and transport in molecular systems. 2 NOISE AND DAMPING Following the derivation given in Ref. , we start by assuming that the coupled exciton-phonon system can be described by the following pair of equations $$i\mathrm{}\dot{\psi }_n+\underset{n^{}}{}J_{nn^{}}\psi _n^{}+\chi u_n\psi _n=0,$$ (1) $$M\ddot{u}_n+M\lambda \dot{u}_n+M\omega _o^2u_n\chi |\psi _n|^2=\eta _n(t).$$ (2) Here $`\psi _n`$ is the amplitude of the exciton wave function corresponding to site $`n`$ and $`u_n`$ represents the elastic degree of freedom at site $`n`$. Furthermore, $`J_{nn^{}}`$ is the dipole-dipole interaction energy, $`\chi `$ is the exciton-phonon coupling constant, $`M`$ is the molecular mass, $`\lambda `$ is the damping coefficient, $`\omega _0`$ is the Einstein frequency of each oscillator, and $`\eta _n(t)`$ is an external force acting on the phonon system. To describe the interaction of the phonon system with a thermal reservoir at temperature $`T`$, $`\eta _n(t)`$ is assumed to be Gaussian white noise with zero mean and the autocorrelation function $$\eta _n(t)\eta _n^{}(t^{})=2M\lambda k_BT\delta (tt^{})\delta _{nn^{}},$$ (3) in accordance with the fluctuation-dissipation theorem ensuring thermal equilibrium. In order to derive a single equation for the dynamics of the exciton system, we start by writing the solution to Eq. (2) in the integral form. Neglecting all exponentially decaying transient terms and making the additional assumption that $`\psi _n`$ varies slowly in space and that only nearest-neighbor coupling $`J`$ is of importance, we obtain in the continuum approximation for the continuous exciton field $`\psi (x,y,t)\sqrt{V/J}e^{4iJt/\mathrm{}}\psi _n(t)`$ the equation of motion $`i\psi _t+^2\psi +|\psi |^2\psi \mathrm{\Lambda }\psi (|\psi |^2)_t+\sigma \psi =0,`$ (4) where $`x`$ and $`y`$ are scaled on the distance $`\mathrm{}`$ between nearest neighbors, a noise density $`\sigma (x,y,t)`$ is not white, but strongly colored , and time was transformed into the dimensionless variable: $`Jt/\mathrm{}t`$. The nonlinear damping parameter $`\mathrm{\Lambda }=\frac{\lambda J}{\mathrm{}\omega _0^2}`$. It can easily be shown that in spite of the presence of the nonlinear damping and multiplicative noise terms in Eq. (4), the norm, defined as $`N={\displaystyle |\psi (x,y,t)|^2𝑑x𝑑y}`$ (5) will still be a conserved quantity. To investigate the influence on the collapse process of the damping and noise terms in Eq. (4), we will use the method of collective coordinates. To this end, we will make some simplifying assumptions. We will assume isotropy, which effectively reduces the problem to one space dimension with the radial coordinate $`r=\sqrt{x^2+y^2}`$. We also assume that the noise $`\sigma `$ can be approximated by radially isotropic Gaussian white noise. The validity of the approximation was discussed in Ref. . Finally, we assume that the collapse process can be described in terms of collective coordinates using the following self-similar trial function for the exciton wave function $`\psi (r,t)`$ $`\psi (r,t)=A(t)\text{sech}\left({\displaystyle \frac{r}{B(t)}}\right)e^{i\alpha (t)r^2}.`$ (6) This trial function, with three real time-dependent parameters $`A`$, $`B`$, and $`\alpha `$ determining the amplitude, width, resp. phase of the wave function, was used in Refs. to investigate the case when $`\mathrm{\Lambda }=0`$ in Eq. (4). The choice of this particular type of trial function can be motivated by regarding it as a generalization of the approximate ground state solution to the ordinary 2-D NLS found in Ref. . From the definition (5) of the norm, we immediately obtain the relation between amplitude and width $`A(t)\sqrt{N}/B(t)`$. In analogy with the treatment for the undamped case in Ref. , we find that it is possible to arrive to the following differential equation for the width $`B`$ of the exciton wave function $`\ddot{B}={\displaystyle \frac{\mathrm{\Delta }}{B^3}}{\displaystyle \frac{\mathrm{\Gamma }\dot{B}}{B^4}}+{\displaystyle \frac{h(t)}{B^2}},`$ (7) where the constants $`\mathrm{\Delta }`$ and $`\mathrm{\Gamma }`$ are some functions of $`N`$ and $`\mathrm{\Lambda }`$. Note that $`\mathrm{\Delta }`$ and $`\mathrm{\Gamma }`$ depend on the initial conditions via $`N`$, and that while $`\mathrm{\Delta }`$ can be either positive or negative, $`\mathrm{\Gamma }`$ is always positive. In the absence of noise and damping the collapse will occur if and only if $`\mathrm{\Delta }<0`$. The white noise $`h(t)`$ has the autocorrelation $`h(t)h(t^{})=2D\delta (tt^{}),`$ (8) where $`D`$ is the dimensionless noise variance. Our numerical calculations show that for $`D<D_{\text{crit}}0.15`$, the effect of the noise is to delay the pseudo-collapse in terms of the ensemble average of the width, in analogy with the similar result obtained in Ref. for the undamped case. For $`D>D_{\text{crit}}`$, we observe a non-monotonic behavior of $`B(t)`$. Initially, the average width will decrease in a similar way as when $`D<D_{\text{crit}}`$, but after some time $`B(t)`$ will reach a minimum value and diverge as $`t\mathrm{}`$. This is due to the fact that for $`D>D_{\text{crit}}`$, the noise is strong enough to destroy the pseudo-collapse and cause dispersion for some of the systems in the ensemble. As $`t\mathrm{}`$ the dominating contribution to $`B(t)`$ will come from the dispersing systems for which $`B\mathrm{}`$, and consequently $`B(t)`$ will diverge for $`D>D_{\text{crit}}`$. The minimum value of $`B(t)`$ will increase towards $`B(0)`$ as $`D`$ increases. 3 NONLOCAL INTERACTIONS In the main part of the previous studies of the discrete NLS models the dispersive interaction was assumed to be short-ranged and a nearest-neighbor approximation was used. However, there exist physical situations that definitely can not be described in the framework of this approximation. The DNA molecule contains charged groups, with long-range Coulomb interaction ($`1/r`$) between them. The excitation transfer in molecular crystals and the vibron energy transport in biopolymers are due to transition dipole-dipole interaction with $`1/r^3`$ dependence on the distance, $`r`$. The nonlocal (long-range) dispersive interaction in these systems provides the existence of additional length-scale: the radius of the dispersive interaction. We will show that it leads to the bifurcative properties of the system due to both the competition between nonlinearity and dispersion, and the interplay of long-range interactions and lattice discreteness. In some approximation the equation of motion is the nonlocal discrete NLS equation of the form $`i{\displaystyle \frac{d}{dt}}\psi _n+{\displaystyle \underset{mn}{}}J_{nm}(\psi _m\psi _n)+|\psi _n|^2\psi _n=0,`$ (9) where the long-range dispersive coupling is taken to be either exponentially, $`J_n=Je^{\beta |n|}`$, or algebraically, $`J_n=J|n|^s`$, decreasing with the distance $`n`$ between lattice sites. The parameters $`\beta `$ and $`s`$ are introduced to cover different physical situations from the nearest-neighbor approximation ($`\beta \mathrm{},s\mathrm{}`$) to the quadrupole-quadrupole ($`s=5`$) and dipole-dipole ($`s=3`$) interactions. The equation (9) conserves the number of excitations $`N=\underset{n}{}|\psi _n|^2`$. We are interested in stationary solutions of Eq. (9) of the form $`\psi _n(t)=\varphi _n\mathrm{exp}(i\mathrm{\Lambda }t)`$ with a real shape function $`\varphi _n`$ and a frequency $`\mathrm{\Lambda }`$. This gives the governing equation for $`\varphi _n`$ $$\mathrm{\Lambda }\varphi _n=\underset{mn}{}J_{nm}(\varphi _m\varphi _n)+\varphi _n^3.$$ (10) Figure 1 shows the dependence $`N(\mathrm{\Lambda })`$ obtained from direct numerical solution of Eq. (10) for algebraically decaying $`J_{nm}`$. A monotonic function is obtained only for $`s>s_{cr}`$. For $`2<s<s_{cr}`$ the dependence becomes nonmonotonic (of $`𝒩`$-type) with a local maximum and a local minimum. These extrema coalesce at $`s=s_{cr}3.03`$. For $`s<2`$ the local maximum disappears. The dependence $`N(\mathrm{\Lambda })`$ obtained analytically using the variational approach is in a good qualitative agreement with the dependence obtained numerically (see ). Thus the main features of all discrete NLS models with dispersive interaction $`J_{nm}`$ decreasing faster than $`|nm|^{s_{cr}}`$ coincide qualitatively with the features obtained in the nearest-neighbor approximation where only one stationary state exists for any number of excitations, $`N`$. However in the case of long-range nonlocal NLS equation (9), i.e. for $`2<s<s_{cr}`$, there exist for each $`N`$ in the interval $`[N_l(s),N_u(s)]`$ three stationary states with frequencies $`\mathrm{\Lambda }_1(N)<\mathrm{\Lambda }_2(N)<\mathrm{\Lambda }_3(N)`$. In particular, this means that in the case of dipole-dipole interaction ($`s=3`$) multiple solutions exist. It is noteworthy that similar results are also obtained for the dispersive interaction of the exponentially decaying form. In this case the bistability takes place for $`\beta \mathrm{\hspace{0.17em}1.67}`$. According to the theorem which was proven in , the necessary and sufficient stability criterion for the stationary states is $`dN/d\mathrm{\Lambda }>0`$. Therefore, we can conclude that in the interval $`[N_l(s),N_u(s)]`$ there are only two linearly stable stationary states: $`\mathrm{\Lambda }_1(N)`$ and $`\mathrm{\Lambda }_3(N)`$. The intermediate state is unstable since $`dN/d\mathrm{\Lambda }<0`$ at $`\mathrm{\Lambda }=\mathrm{\Lambda }_2`$. The low frequency states are wide and continuum-like while the high frequency solutions represents intrinsically localized states with a width of a few lattice spacings. It can be shown that the existence of two so different soliton states for one value of the excitation number, $`N`$, is due to the presence of two different length scales in the system: the usual scale of the NLS model which is related to the competition between nonlinearity and dispersion (expressed in terms of the ratio $`N/J`$ ) and the range of the dispersive interaction $`\xi `$. Having established the existence of bistable stationary states in the nonlocal discrete NLS system, a natural question that arises concerns the role of these states in the full dynamics of the model. In particular, it is of interest to investigate the possibility of switching between the stable states under the influence of external perturbations, and to clear up what type of perturbations can be used to control the switching. Switching of this type is important for example in the description of nonlinear transport and storage of energy in biomolecules like the DNA, since a mobile continuum-like excitation can provide action at distance while the switching to a discrete, pinned state can facilitate the structural changes of the DNA . As it was shown recently in , switching will occur if the system is perturbed in a way so that an internal, spatially localized and symmetrical mode (’breathing mode’) of the stationary state is excited above a threshold value.
no-problem/9906/cond-mat9906200.html
ar5iv
text
# Quantum cavitation in liquid 3He: dissipation effects ## Abstract We have investigated the effect that dissipation may have on the cavitation process in normal liquid <sup>3</sup>He. Our results indicate that a rather small dissipation decreases sizeably the quantum-to-thermal crossover temperature $`T^{}`$ for cavitation in normal liquid <sup>3</sup>He. This is a possible explanation why recent experiments have not yet found clear evidence of quantum cavitation at temperatures below the $`T^{}`$ predicted by calculations which neglect dissipation. Quantum cavitation in superfluid liquid <sup>4</sup>He has been unambiguously observed using ultrasound experimental techniques . These experiments have shown that quantum cavitation takes over thermal cavitation at a temperature ($`T`$) around 200 mK, in good agreement with theoretical calculations , so that the problem of cavitation in liquid <sup>4</sup>He can be considered as satisfactorily settled. The crossover temperature corresponding to <sup>3</sup>He has also been calculated, predicting that $`T^{}`$ 120 mK. It turns out that preliminary results obtained in a recent experiment have not shown clear evidence of quantum cavitation for temperatures even below that value. However, the phenomenon has been firmly established as a stochastic process. A possible explanation is that thermal cavitation is still the dominant process down to temperatures lower than predicted. The method of Ref. (see also Ref. ) is based, on the one hand, in using a density functional that reproduces the thermodynamical properties of liquid <sup>3</sup>He at zero temperature (equation of state, effective mass, etc), as well as the properties of the <sup>3</sup>He free surface. A major advantage of using a density functional is that one can handle bubbles in the vicinity of the spinodal region, where they are not empty objects and any attempt to describe the critical bubble in terms of a sharp surface radius fails . On the other hand, we have used a functional-integral approach especially well suited to find $`T^{}`$. This gives us some confidence on the values obtained for the crossover temperature, and inclines us to think that any appreciable discrepancy between theory and experiment has to be attributed not to the method itself, but to some physical ingredient which has been overlooked in the formalism. One such ingredient in the case of liquid <sup>3</sup>He is dissipation, which is known to decrease $`T^{}`$. Since <sup>4</sup>He is superfluid below the lambda temperature, we are actually treating both quantum fluids within the same framework, the behavior of <sup>4</sup>He being accounted for by the dissipationless version of the general formalism. Our starting point is the real time Lagrangian density $`(\rho ,s)`$ $$(\rho ,s)=m\dot{\rho }s(\rho ,s),$$ (1) where $`\rho (\stackrel{}{r},t)`$ denotes the particle density, $`m`$ the <sup>3</sup>He atomic mass, and $`s(r,t)`$ is the velocity potential, i.e, the collective velocity is $`\stackrel{}{u}(\stackrel{}{r},t)=s(\stackrel{}{r},t)`$. The Hamiltonian density $`(\rho ,s)`$ reads $$(\rho ,s)=\frac{1}{2}m\rho \stackrel{}{u}^2+\left[\omega (\rho )\omega (\rho _m)\right],$$ (2) where $`\omega (\rho )`$ is the grand potential density of the system and $`\rho _m`$ is the density of the metastable homogeneous liquid. We refer the reader to Ref. and references therein for details. To describe the dynamics in the dissipative regime while still being able to deal with inhomogeneous <sup>3</sup>He, which is crucial for a proper description of cavitation in liquid helium, we have introduced a phenomenologycal Rayleigh’s dissipation function $``$ $$=\frac{1}{2}\xi \frac{\dot{\rho }^2}{\rho ^2}.$$ (3) From Lagrange’s equations $$\frac{}{t}\left(\frac{\delta }{\delta \dot{x}}\right)\frac{\delta }{\delta x}=\frac{}{\dot{x}},$$ (4) with $`x`$ being either $`s`$ or $`\rho `$, one gets the continuity and motion equation, respectively: $$\dot{\rho }+(\rho \stackrel{}{u})=0$$ (5) $$m\left\{\frac{u_k}{t}+u_i_ku_i\right\}=_k\left(\frac{\delta \omega }{\delta \rho }\right)+_k\left[\xi \frac{1}{\rho ^2}(\rho \stackrel{}{u})\right].$$ (6) For an homegeneous fluid, the equation of motion ressembles the Navier-Stokes equation $$m\rho \left\{\frac{u_k}{t}+u_i_ku_i\right\}=_kP+\eta \mathrm{\Delta }u_k+\left(\zeta +\frac{1}{3}\eta \right)_k(\stackrel{}{u}),$$ (7) where $`P`$ is the pressure. For liquid <sup>3</sup>He at low $`T`$, dissipation depends on the mean free path of quasiparticles, and a precise estimation of the magnitude of this effect in the tunneling process is difficult. Since our interest here is to explore the effect of a small viscosity on $`T^{}`$, we have adopted the pragmatic point of view of identifying $`\xi `$ with $`\zeta +\eta /3`$ and presenting results for different $`\xi `$’s close to the experimental $`\eta `$ value (it is known that at low temperatures, the shear viscosity coefficient $`\eta `$ is much larger than the bulk viscosity coefficient $`\zeta `$, see for example Ref. ). Using the macroscopic viscosity coefficient, one should have in mind that we are likely overestimating the dissipation effects. To obtain $`T^{}`$ we have proceeded as indicated in Ref. , writing the above equations in imaginary time $`\tau =it`$ and linearizing them around the critical bubble density $`\rho _0`$, seeking solutions of the kind: $$\rho (\stackrel{}{r},\tau )\rho _0(r)+\rho ^1(r)e^{i\omega _s\tau }.$$ (8) Upon linearization, we end up with the following equation for $`\omega _s`$ and $`\rho ^1(r)`$: $$\rho ^1(r)\left[m\omega _s^2_1\xi \omega _s_2\right]\rho ^1(r)=0.$$ (9) The differential operators $`_1`$ and $`_2`$ in Eq. (9) are, respectively, the linearization of $$\left\{\rho \left(\frac{\delta \omega }{\delta \rho }\right)\right\}\mathrm{and}\left\{\rho \left(\frac{1}{\rho ^2}\right)\right\},$$ (10) in which only first order terms in $`\rho ^1(r)`$ and its derivatives have been kept . Since $`\xi `$ depends on the density as $`\rho ^{5/3}`$, in actual calculations we have made a local density approximation, using as form factor in Eq. (3) the expression $`1/(\rho _{sat}^{5/3}\rho ^{1/3}(r))`$, where $`\rho _{sat}`$ is the density of the liquid at $`T`$ = 0 and $`P`$ = 0, and $`\xi `$ is then density-independent. Eq. (9) is a fourth-order linear differential, generalized eigenvalue equation, whose physical solutions have to fulfill $`\rho ^1(0)=\rho ^{1\prime \prime \prime }(0)=0`$, and fall exponentially to zero at large distances. We have solved it as indicated in Ref. . Once the largest dissipation-renormalized frequency $`\omega _s`$ has been determined, the crossover temperature is obtained as $`T^{}=\mathrm{}\omega _s/(2\pi )`$. Table I collects the equation of state near the spinodal point ($`\rho _{sp}`$ = 0.01191 $`\mathrm{\AA }^3`$, $`P_{sp}=3.102`$ bar), and other quantities which are of interest to analyze the experimental results . Our spinodal point compares very well with recent Monte Carlo calculations ($`\rho _{sp}`$ = 0.0121 $`\mathrm{\AA }^3`$, $`P_{sp}=3.12\pm 0.10`$ bar), and also with other phenomenological approaches . We show $`T^{}`$ in Fig. 1 as a function of pressure for different $`\xi `$ values. In particular, $`\xi `$ = 100 $`\mu `$P roughly corresponds to the experimental value of $`\eta `$ at $`P`$ = 0 and $`T`$ = 100 mK. The associated effective quantum action $`𝒮`$ obtained as $`𝒮=\mathrm{\Delta }\mathrm{\Omega }/T^{}`$, where $`\mathrm{\Delta }\mathrm{\Omega }`$ is the maximum of the energy barrier, is displayed in Fig. 2. Fig. 3 shows $`\rho ^1(r)`$ at $`P=3`$ bar for three $`\xi `$ values, as well as the critical bubble density $`\rho _0(r)`$. The linearized continuity equation $`\rho ^1(r)(\rho _0\stackrel{}{u})`$ implies that $`\rho ^1(r)`$ must have nodes, as it imposes that the integral of $`\rho ^1(r)`$ is zero when taken over the whole space. When $`\xi `$ is small enough and the $`_2`$ term in Eq. (9) can be treated perturbatively, a straightforward calculation yields $$\omega _s=\sqrt{\omega _{0,0}^2+\left(\frac{\xi \mu _{\mathrm{\hspace{0.17em}2}}}{2m}\right)^2}\frac{\xi \mu _{\mathrm{\hspace{0.17em}2}}}{2m},$$ (11) where we have used a standard matrix notation to denote as $`\omega _{0,0}`$ and $`|\rho _0^{1(0)}`$ the higher frequency solution of the non-viscous problem $`(m\omega _{0,n}^2_1)|\rho _n^{1(0)}`$ = 0, and have defined $`\mu _{\mathrm{\hspace{0.17em}2}}\rho _0^{1(0)}|_2|\rho _0^{1(0)}>0`$. Equation (11) is similar to that given in Ref. for the dissipation-renormalized frequency $`\omega _s`$ in the case of frequency-independent damping. Figures (1-2) indicate that for viscosity values of the order of the experimental one, a sizeable decrease of the crossover temperature occurs. However, the present model still predicts that a transition from thermal to quantum cavitation takes place in liquid <sup>3</sup>He. We finally obtain the homogeneous cavitation pressure $`P_h`$ from the equation : $$1=(Vt)_{exp}J_0e^𝒮,$$ (12) taking for the experimental volume$`\times `$time $`(Vt)_{exp}`$ a typical value of $`10^8\mathrm{\AA }^3`$ s, which correspods to <sup>4</sup>He experiments . We have adopted for $`J_0`$ the same prescription as in Ref. . Figure 4 shows the homogeneous cavitation pressure as a function of $`T`$ for the $`\xi `$ values we have been using. In conclusion, we have developed a phenomenological model to size the effect of dissipation in the cavitation process in liquid <sup>3</sup>He that allows one to handle realistic critical cavitation configurations near the spinodal line, and to treat both helium isotopes within the same frame, using the dissipationless limit of the method in the case of <sup>4</sup>He. The results we have obtained indicate that for liquid <sup>3</sup>He even a moderate dissipation may reduce the crossover temperature in a non-negligible amount, displacing the homogeneous cavitation pressure towards the spinodal value. Viscosity may then be the reason of the inconclusive results for quantum cavitation reported in Ref. which, if confirmed, would indicate that dissipation plays a crucial role in quantum cavitation in liquid helium. The experimental study of cavitation in undersaturated <sup>3</sup>He-<sup>4</sup>He mixtures might then uncover a structure much richer than that theoretically described in Ref. , since <sup>4</sup>He is still superfluid and <sup>3</sup>He is in the normal phase. This would open the possibility of studying the influence of dissipation in the cavitation process varying the <sup>3</sup>He concentration. We would like to thank Sebastien Balibar, Eugene Chudnovsky and Patrick Roche for useful discussions. This work has been supported in part by the DGICYT (Spain), grant PB95-1249, and by the Generalitat de Catalunya ‘Accions Integrades’ and 1998SGR-00011 programs.
no-problem/9906/astro-ph9906209.html
ar5iv
text
# 1 Introduction: ## 1 Introduction: Blazars are a class of active galactic nuclei whose emission is believed to arise predominantly from a relativistic jet whose axis makes a small angle with our line of sight. More than 60 blazars have been detected with the Energetic Gamma-Ray Experiment (EGRET) (Hartman et al. 1999). A few BL Lacertae objects (BL Lacs), a sub-class of blazars, have also been detected as TeV $`\gamma `$-ray emitters (Ong 1998). No model for the origin of the $`\gamma `$-ray emission is generally accepted at this time. Two popular classes are those in which high energy electrons produce the $`\gamma `$-rays by inverse Compton scattering of low energy photons (e.g., Maraschi, Ghisellini & Celotti 1992; Dermer, Schlickeiser & Mastichiadis 1992; Sikora, Begelman & Rees 1994) and those in which high energy protons produce $`\gamma `$-rays by initiating cascades in the jets (e.g., Mannheim 1993). Contemporaneous observations at several wavelengths can be used to derive physical conditions in and around the blazar jet and may resolve which emission mechanism operates in the objects. Markarian 421 (Mrk 421) is the closest known BL Lac ($`z=0.031`$) and is an established very high energy (VHE, E$`>`$250 GeV) $`\gamma `$-ray source (e.g., Punch et al. 1992; Petry et al. 1996). Mrk 421 is also an EGRET source (Hartman et al. 1999). The VHE emission from Mrk 421 is extremely variable, showing flaring activity on time scales as short as 15 minutes (Gaidos et al. 1996) with little or no baseline level emission (Buckley et al. 1996). The spectrum of Mrk 421 is consistent with a power law that extends to at least 10 TeV with no evidence of a sharp cut-off, and no evidence of variability (e.g., Krennrich et al. 1999a,b). Multiwavelength campaigns on Mrk 421 (e.g., Buckley et al. 1996) have revealed correlations between X-rays and VHE $`\gamma `$-rays and evidence for correlated optical/UV variability. The flux amplitude of the X-ray and VHE $`\gamma `$-ray variations was similar and the variability time profiles were the same, on day-scales. These rapid, correlated variations have permitted stringent limits to be placed on the Doppler factor and magnetic fields of the Mrk 421 jet (e.g., Buckley et al. 1996) and these data have become important tests for emission models (e.g., Mannheim 1998; Buckley 1998; Tavecchio, Maraschi & Ghisellini 1998). Despite the successes of these campaigns, the light curves were not densely sampled, so the multiwavelength variability could not be measured on time-scales less than one day. In order to better measure the flaring behavior of Mrk 421, several more intense multiwavelength campaigns were conducted in 1998 using longer exposures in X-rays and VHE $`\gamma `$-rays, and combining the data from several VHE $`\gamma `$-ray telescopes. Here, we present the results of a campaign in 1998 April with BeppoSAX and the Whipple $`\gamma `$-ray telescope. ## 2 Observations: ### 2.1 BeppoSAX: The scientific payload carried by BeppoSAX is fully described in Boella et al. (1997a). The data of interest here derive from three coaligned instruments, the Low Energy Concentrator Spectrometer (LECS, 0.1-10 keV, Parmar et al. 1997), the Medium Energy Concentrator Spectrometer (MECS, 2-10 keV, Boella et al. 1997b) and the Phoswich Detector System (PDS, 12-300 keV, Frontera et al. 1997). The observations with BeppoSAX reported here consist of two exposures lasting approximately 100 kiloseconds each. The data reduction for the PDS was done using the XAS software (Chiappetti & Dal Fiume 1997), while for the LECS and MECS linearized cleaned event files generated at the BeppoSAX Science Data Centre (SDC) were used. No appreciable difference was found extracting the MECS data with the XAS software. Light curves were accumulated from each instrument with the usual choices for extraction radius and background subtraction as described in Chiappetti et al. (1999). ### 2.2 Whipple: The VHE $`\gamma `$-ray observations were made with the Whipple Observatory 10 m telescope (Cawley et al. 1990). At the time of these observations, the telescope camera consisted of 331 photomultiplier tubes with a combined field of view of 4.8. Also, light-cones were not in place and this, as well as reduced reflectivity of the mirrors, resulted in a somewhat higher energy threshold than usual for the telescope, $``$500 GeV. Events were parameterized with a standard moment analysis and candidate $`\gamma `$-rays were selected using a variation of the Supercuts analysis (Reynolds et al. 1993) appropriate for the large camera field of view and for maintaining a constant energy threshold as a function of observation elevation (see Table 1 and discussion below). Observations were taken on the nights of 1998 April 21, 22, 23 and 24. To permit longer observations within each night, data were taken over a large range of zenith angles ($``$7 to 60). The collection area and energy threshold increase with zenith angle and the $`\gamma `$-ray selection is a function of zenith angle, so the observed $`\gamma `$-ray rates can change as a function of elevation even from a source of constant $`\gamma `$-ray emission. To obtain a light-curve which shows only the intrinsic source variations, it was necessary to determine software cuts which result in a constant energy threshold as a function of elevation and to calculate collection areas as a function of elevation for those energy thresholds in order to normalize the $`\gamma `$-ray rates. Because contemporaneous observations of the Crab Nebula were not available with sufficient statistics over all the zenith angles, the analysis presented here relies entirely on Monte Carlo shower simulations. Simulated $`\gamma `$-ray induced showers at zenith angles of 20, 40, 45, and 55 were generated with ISUSIM (Mohanty et al. 1998) to determine the size cut required at each zenith angle to obtain a common energy threshold of 2 TeV and to estimate the collection areas at these zenith angles for the 2 TeV energy threshold. To normalize the $`>`$2 TeV rate measurements at the different elevations we multiply them by the ratio of effective collection area at a given zenith angle to the effective collection area at a zenith angle of 20. The software trigger threshold applied at each zenith angle to set the energy threshold at 2 TeV and the ratio of effective areas for the four zenith angle ranges are shown in Table 1. The results reported here are based on limited statistics and are therefore preliminary. The aim of this analysis is to derive normalized fluxes as a function of time rather than absolute fluxes and energy spectra. ## 3 Results: The light curves for the $`\gamma `$-ray and X-ray observations are shown in Figure 1. Three X-ray energy bands are shown and the $`\gamma `$-ray light curve shows the normalized E$`>`$2 TeV data for the measurements. Each $`\gamma `$-ray point represents a 28 minute observation. The count rates for the measurements are normalized to their respective averages for these observations. The rise and fall of a large amplitude flare is clearly evident in all data sets on the first day of observations. Observations after the first day did have detectable fluxes, but showed no significant variability on day or shorter time-scales. For the observations on April 21, the amplitude of the X-ray flaring increases with increasing energy, but it is close to a factor of 2 in all three bands. The $`>`$2 TeV light curve shows a 4-fold variation in flux. The flux in the 0.1-2 keV, 4-10 keV, and $`>`$2 TeV energy bands peaks at approximately the same time (within one-half of one hour), but the decay time for the TeV light curve is significantly shorter than that of the LECS and MECS light curves. The 12-26 keV light curve measured by the PDS instrument appears to peak slightly later than the others, but the statistical uncertainty in the data precludes a definitive measurement. A detailed investigation of possible leads or lags in the data is underway. The TeV spectrum does not change significantly during the rapid flare, nor is it significantly different than previous measurements (Krennrich et al. 1999b). ## 4 Discussion and Conclusions: These observations show, for the first time, that the TeV and keV fluxes from Mrk 421 are correlated on hour time scales while at the same time indicate that the $`\gamma `$-rays and X-rays are not completely correlated. Neither the larger variability amplitude at TeV energies than at X-rays nor the difference in the variability time scales have been seen previously. The reason for the difference in the decay time-scale of the flare at TeV and keV energies is not clear. The differences could reflect the nature of the flaring mechanism. For example, a variation in the electron spectrum and the energy density of the low energy photons up-scattered to TeV energies (Maraschi et al. 1999) might produce such a flare. The differences may also indicate that the particles which produce the X-rays are not the same as the particles which produce the TeV $`\gamma `$-rays. This is possible in models where the progenitor particles are electrons or protons. In addition, the region of the broadband spectrum of Mrk 421 observed by BeppoSAX spans the end of the synchrotron emission and the onset of the high energy emission (c.f., Buckley et al. 1996). As such, the X-ray emission may reflect contributions from more than one population of source particles, regardless of the emission mechanism. Detailed model fitting of this data, which is beyond the scope of this paper, is necessary to investigate these possibilities. References Boella, G., et al. 1997a, A&AS, 122, 299 Boella, G., et al. 1997b, A&AS, 122, 327 Buckley, J.H., et al. 1996, ApJ, 472, L9 Buckley, J.H. 1998, Science, 279, 676 Cawley, M.F., et al. 1990, Exp. Astron., 1, 173 Chiappetti, L., & Dal Fiume, D. 1996, in Proc. of the 5th Workshop “Data Analysis in Astronomy” (Erice), ed. V. Di Gesú, et al., 101 Chiappetti, L., et al. 1999, ApJ, in press Dermer, C.D., Schlickeiser, R., & Mastichiadis, A. 1992, A&A, 256, L27 Frontera, F., et al. 1997, A&AS, 122, 357 Gaidos, J.A., et al. 1996, Nature, 383, 319 Hartman, R.C., et al. 1999, ApJS, in press Krennrich, F., et al. 1999b, Proc. 26th ICRC (Salt Lake City), OG 2.1.02 Mannheim, K. 1993, A&A, 269, 67 Mannheim, K. 1998, Science, 279, 684 Maraschi, L., Ghisellini, G., & Celotti, A. 1992, ApJ, 397, L5 Maraschi, L., et al. 1999, in TeV Astrophysics of Extragalactic Sources, ed. M. Catanese & T.C. Weekes, Astrop. Phys., in press Mohanty, G., et al. 1998, Astrop. Phys., 9, 15 Ong, R.A. 1998, Phys. Reports, 305, 93 Parmar, A.N., et al. 1997, A&AS, 122, 309 Petry, D., et al. 1996, A&A, 311, L13 Punch, M., et al. 1992, Nature, 358, 477 Reynolds, P.T., et al. 1993, ApJ, 404, 206 Sikora, M., Begelman, M.C., & Rees, M.J. 1994, ApJ, 421, 153 Tavecchio, F., Maraschi, L., & Ghisellini, G. 1998, ApJ, 509, 608
no-problem/9906/cond-mat9906230.html
ar5iv
text
# Examination of self interaction correction methods for Na clusters ## I Introduction The local density approximation (LDA) provides a powerful practical technique to apply the Kohn-Sham framework to interacting many body problems. A problem of this method is unphysical self interaction, i.e. the interaction of a particle with itself. Perdew and Zunger proposed a prescription to remedy this shortcoming, which has been used for atoms, molecules, bulk systems and also metal clusters. Though there still remain some effects of self interaction, the major part of the problem is removed in this method. Compared with the Hartree-Fock theory, which is free from the self interaction problem, the local density approximation with the self interaction correction (SIC) has advantages such as, i) exchange and correlation energies can be relatively easily handled in the same manner, ii) the resultant single particle energies well approximate the physical removal energy from each orbit, iii) the numerical calculation is much lighter, especially for three dimensional calculations. A characteristic feature of the SIC method of Perdew and Zunger is that the energy functional depends not only on the total density, but explicitly also on the density of each single particle orbital. In almost all calculations for closed shell atoms and metal clusters, the single particle densities in the energy functional are substituted by the spherically averaged densities. The central single particle potential is then deduced by taking functional derivative of the resultant energy functional with respect to the spherically averaged single particle density. Following more closely the original idea of Perdew and Zunger, on the other hand, Harrison proposed a method of using the original single particle densities without introducing spherical averaging. Since the energy functional is not invariant under unitary transformation of single particle orbitals, Harrison represented the single particle orbitals by either spherical harmonics or Cartesian basis as two choices. Though the method by Harrison works well for atoms, it has not been tested for metal clusters. We address this question in this paper by taking Na clusters as an example. We show that it does not work well, especially for large clusters which have single particle orbitals with large angular momentum. We confine our study to the exchange energy without referring to the correlation energy in order to make the argument clear and compare the results with those of Hartree-Fock calculations. We show that a better agreement with the Hartree-Fock calculations is obtained if one applies Harrison’s method only to the Hartree term. In addition to the validity of Harrison’s method, we discuss in this paper the problem of non-diagonal Lagrange multipliers originating from the orthonormality of single particle orbitals in the self interaction correction method of Perdew and Zunger. We show that the non-diagonal property of the Lagrange multipliers introduces a negligible effect to single particle energies as well as the total energy. The paper is organized as follows. In Sec.II the SIC method of Perdew and Zunger and Harrison’s approach are briefly explained. In Sec.III A the exchange energies calculated by several SIC methods are compared, and the effect of non-diagonal Lagrange multipliers is discussed in Sec.III B. Summary and conclusion are given in Sec.IV. ## II SIC formalism and Harrison’s method The total energy in the self interaction corrected local density approximation (SIC-LDA) by Perdew and Zunger is expressed as $`E_{\mathrm{TOT}}=T+E_{\mathrm{ext}}+E_H+E_X^{\mathrm{SIC}},`$ (1) where $`T={\displaystyle \frac{1}{2}}{\displaystyle \underset{i}{}}{\displaystyle d^3r\psi _i^{}(𝐫)^2\psi _i(𝐫)},`$ (2) $`E_{\mathrm{ext}}[\rho ]={\displaystyle d^3rv_{\mathrm{ext}}(𝐫)\rho (𝐫)},`$ (3) $`E_H[\rho ]={\displaystyle \frac{1}{2}}{\displaystyle d^3rd^3r^{}\frac{\rho (𝐫)\rho (𝐫^{})}{\left|𝐫𝐫^{}\right|}},`$ (4) $`E_X^{\mathrm{SIC}}=E_X^{\mathrm{LDA}}[\rho _{},\rho _{}]{\displaystyle \underset{i=1}{\overset{N}{}}}\left\{E_H[\rho _i]+E_X^{\mathrm{LDA}}[\rho _i,0]\right\},`$ (5) $`E_X^{\mathrm{LDA}}[\rho _{},\rho _{}]={\displaystyle \frac{3}{2}}\left({\displaystyle \frac{3}{4\pi }}\right)^{1/3}{\displaystyle \underset{\sigma =,}{}}{\displaystyle d^3r\rho _\sigma (𝐫)^{4/3}}.`$ (6) Here all quantities are in Hartree atomic units, i.e. $`m=e^2=\mathrm{}=1`$. As mentioned in the introduction, the correlation energy has been neglected. We take the effects of ions into account in the spherical jellium model. The external potential is then given by $`v_{\mathrm{ext}}(𝐫)`$ $`=`$ $`\{\begin{array}{cc}Z/(2R_{\mathrm{jell}})\left\{3\left(r/R_{\mathrm{jell}}\right)^2\right\}\hfill & rR_{\mathrm{jell}}\hfill \\ Z/r\hfill & r>R_{\mathrm{jell}},\hfill \end{array}`$ (9) where the jellium radius $`R_{\mathrm{jell}}`$ is related to the number of atoms in the cluster $`Z`$ by $`R_{\mathrm{jell}}=r_sZ^{1/3}`$, $`r_s`$ being the bulk Wigner-Seitz radius which is 4 a.u. for Na. The Euler equation under the orthonormality condition $`{\displaystyle \frac{\delta }{\delta \psi _i^{}(𝐫)}}\left\{E_{\mathrm{tot}}+{\displaystyle \underset{ij}{}}ϵ_{ij}\left(\delta _{ij}{\displaystyle d^3r\psi _j^{}(𝐫)\psi _i(𝐫)}\right)\right\}=0`$ (10) results in the following coupled equations for the single particle wave functions $`\left\{{\displaystyle \frac{1}{2}}^2+v_{\mathrm{ext}}(𝐫)+{\displaystyle d^3r^{}\frac{\rho (𝐫^{})}{\left|𝐫𝐫^{}\right|}}+v_X^{\mathrm{S}IC(i)}(𝐫)\right\}\psi _i(𝐫)={\displaystyle \underset{j}{}}ϵ_{ij}\psi _j(𝐫),`$ (11) $`v_X^{\mathrm{S}IC(i)}(𝐫)=\left({\displaystyle \frac{3}{\pi }}\right)^{1/3}\rho ^{1/3}(𝐫)\left\{{\displaystyle d^3r^{}\frac{\rho _i(𝐫^{})}{\left|𝐫𝐫^{}\right|}}2\left({\displaystyle \frac{3}{4\pi }}\right)^{1/3}\rho _i^{1/3}(𝐫)\right\}.`$ (12) The Lagrange multiplier $`ϵ_{ij}`$ becomes non-diagonal because of the orbital dependence of the self interaction corrected exchange potential. In the following, we approximate it by the diagonal components and solve the following non-coupled equations $`\left\{{\displaystyle \frac{1}{2}}^2+v_{\mathrm{ext}}(𝐫)+{\displaystyle d^3r^{}\frac{\rho (𝐫^{})}{\left|𝐫𝐫^{}\right|}}+v_X^{\mathrm{S}IC(i)}(𝐫)\right\}\psi _i(𝐫)=ϵ_i\psi _i(𝐫)`$ (13) We discuss the validity of this approximation in Sec. III B for Na clusters. As we see in Eq.(5) the SIC method of Perdew and Zunger is characteristic in that the total energy functional depends explicitly on the individual orbital density. A consequence is that it loses invariance under the unitary transformation of single particle orbitals. Another problem is that the numerical load is heavy, because one has to solve three-dimensional equations instead of the one-dimensional equations for the radial motion of electrons even for closed shell atoms and metal clusters. In applying this formalism to those systems, one usually replaces $`\rho _i(𝐫)`$ in the curly brackets in Eqs.(5) and (12) by the spherically averaged orbital density given by $`\stackrel{~}{\rho }_i(r)`$ $`\stackrel{~}{\rho }_i(r)`$ $`=`$ $`{\displaystyle \frac{1}{4\pi }}{\displaystyle 𝑑\widehat{𝐫}\rho _i(𝐫)}.`$ (14) This prescription certainly reduces the numerical load, because the resultant single particle potentials become central potentials. However, it does not optimize the SIC following the original scheme of Perdew and Zunger. Harrison proposed an alternative procedure, which avoids replacing the single particle densities in the energy functional by spherically averaged ones. As mentioned in the introduction, he expressed them in the spherical harmonic basis $`\rho _{nlm}^{\mathrm{SH}}(𝐫)=\left|{\displaystyle \frac{u_{nl}(r)}{r}}Y_l^m(\widehat{𝐫})\right|^2,`$ (15) or in Cartesian basis $`\rho _{nlm}^\mathrm{C}(𝐫)=\{\begin{array}{cc}\left|{\displaystyle \frac{u_{nl}(r)}{r}}Y_l^0(\widehat{𝐫})\right|^2\hfill & m=0\hfill \\ \left|{\displaystyle \frac{u_{nl}(r)}{r}}{\displaystyle \frac{Y_l^{|m|}(\widehat{𝐫})\pm Y_l^{|m|}(\widehat{𝐫})}{\sqrt{2}}}\right|^2\hfill & m0.\hfill \end{array}`$ (18) He noticed that the resultant energy functional can be expressed in terms of the spherically averaged orbital density after the integration over angle, $`E_X^{\mathrm{SIC}}`$ $`=`$ $`E_X^{\mathrm{LDA}}[\rho _{},\rho _{}]{\displaystyle \underset{n,l}{}}2(2l+1)\left\{{\displaystyle \underset{k=0}{\overset{2l}{}}}E_H^{k,l}[\stackrel{~}{\rho }_{nl}]+c_X^lE_X^{\mathrm{LDA}}[\stackrel{~}{\rho }_{nl},0]\right\},`$ (19) $`E_H^{k,l}[\stackrel{~}{\rho }_{nl}]`$ $`=`$ $`{\displaystyle \frac{1}{2}}c_H^{k,l}{\displaystyle 𝑑r𝑑r^{}\left|u_{nl}(r)\right|^2\left|u_{nl}(r^{})\right|^2\frac{r_<^k}{r_>^{k+1}}}.`$ (20) The coefficients $`c_H^{k,l}`$ and $`c_X^l`$ for each of the spherical harmonic and Cartesian representations are listed in Tables I and II. Harrison then calculated the corresponding central potential for each set of quantum nembers $`n,l`$ by taking the functional derivative with respect to the spherically averaged orbital density $`\stackrel{~}{\rho }_{nl}(r)`$ $`v_X^{\mathrm{S}IC(nl)}(r)`$ $`=`$ $`{\displaystyle \frac{\delta E_X^{\mathrm{SIC}}[\stackrel{~}{\rho }_{nl}]}{\delta \stackrel{~}{\rho }_{nl}(r)}}.`$ (21) Though this method restricts the variational space smaller than that in the original scheme of Perdew and Zunger since it presumes a spherically symmetric potential from the beginning, Harrison showed that his method still improves both the exchange energy and the total energy for atoms compared with the simple procedure where $`\rho _i(𝐫)`$ is replaced by the spherically averaged orbital density. It is an interesting question to see whether Harrison’s method can be applied to metal clusters. A simple minded consideration would suggest that Harrison’s method becomes more powerful in metal clusters. This is because Harrison’s treatment should have a large effect on high angular momentum orbitals which play more important roles in metal clusters than in atoms where the main contribution to the SIC originates from the $`1s`$ state . In the next section, we apply Harrison’s method to Na clusters, and show that it does not work well contrary to the simple expectation. ## III Results and discussion ### A Exchange and total energy We compare in Table III the exchange energy per electron for Na clusters calculated by several methods. In this table and in what follows, the abbreviations SA-SICX, SH-SICX and C-SICX stand for the SICX calculation using the spherically averaged, spherical harmonic and Cartesian orbital densities, respectively. The difference between the HF and the other calculations represents the error of each method since the HF calculation provides the exact exchange energy. Strictly speaking, the exact exchange energy in the Kohn-Sham formalism is the one given by the optimized effective potential method . It is known, however, that it is nearly the same as that given by the HF calculation for atoms. We first compare the results of HF, LDAX and SA-SICX for four different Na clusters. The order of the estimated exchange energies, which are negative, is LDAX $`>`$ SA-SICX $`>`$ HF irrespectively of the size of the Na cluster. This is different from the order for atoms, where LDAX $`>`$ HF $`>`$ SA-SICX . The deviation of the result of LDAX from that of HF gets smaller with the size of the Na cluster, while that of SA-SICX is almost constant. We next compare the results of HF and three SIC methods. The absolute value of the exchange energy calculated by using the spherical harmonic and Cartesian orbital densities becomes smaller than that estimated by the SA-SICX. Including the negative sign, the order is C-SICX $`>`$ SH-SICX $`>`$ SA-SICX. Consequently, the deviation of the results of SH-SICX and C-SICX from that of the HF gets larger than that for SA-SICX. These two calculations are even worse than LDAX for large systems. Their deviation from the HF calculation gets larger with increasing size of the cluster. For atoms, the order of the exchange energies C-SICX $`>`$ SH-SICX $`>`$ SA-SICX is the same as that for Na clusters. However, the result of C-SICX is still below that of HF. This means that both of SH-SICX and C-SICX have smaller deviation from the HF result than SA-SICX and provide better prescriptions for atoms. We now investigate the reason why the SH-SICX and C-SICX methods are inferior to SA-SICX for Na clusters. Since the self-consistent density in the three SICX methods is almost the same, the first term of Eq.(19) is nearly the same for the three methods. The major difference among them is therefore associated with the second term of Eq.(19), which consists of contributions from orbitals with various $`n,l`$ quantum numbers. In order to see the physics clearly, we consider the following self interaction averaged over the azimuthal quantum number for each set of quantum numbers $`n,l`$, $`{\displaystyle \frac{1}{2l+1}}{\displaystyle \underset{m}{}}\left\{E_H[\rho _{lnm}]+E_X^{\mathrm{LDA}}[\rho _{nlm},0]\right\}.`$ (22) Its value evaluated by inserting the orbital densities given by Eqs.(14), (15) and (18) is compared in Fig.1 for the occupied orbitals in Na<sub>40</sub>. The three methods agree well for the $`s`$ orbitals. The small differences originate from a subtle difference of the total densities in the three methods. On the other hand, a large difference appears among them for the finite angular momentum states, i.e. for the $`p`$, $`d`$ and $`f`$ orbitals. Two specific features can be remarked, i) always SA-SICX $`>`$ SH-SICX $`>`$ C-SICX, and ii) the higher the orbital angular momentum is, the larger the differences among them are. The key point to understand these features is the degree of localization of single particle density adopted in the three SICX methods. Clearly they differ in their angular treatment of the orbital density. C-SICX has the most distinct localization in angle, while SA-SICX has, of course, no angular localization. Symborically, we express this situation of the degree of localization as C-SICX $`>`$ SH-SICX $`>`$ SA-SICX. This difference gets more prominent for higher angular momentum. The problem is that the angular localization property affects the self Hartree term, i.e. the first term in the curly brackets in Eq.(5), and the self exchange terms, i.e. the second term, in different way. Since the self Hartree term uses the exact long range Coulomb interaction between electrons, it does not depend so much on the density profile of electrons including the angular localization. On the other hand, the self exchange term is evaluated based on the LDA. This is equivalent to assuming a short range $`\delta `$ interaction, which leads to a strong sensitivity of the exchange energy on the details of the density profile of electrons. The more the orbital localizes, the larger the absolute value of the self exchange energy is. In C-SICX, which has the most prominent angular localization, the negative self exchange energy increases with angular momentum and eventually overwhelms the self Hartree term resulting in the sign change of the total self interaction energy. This can be clearly seen in Fig.1. A less prominent, but a similar, trend can be seen also for SH-SICX. However, this strong sensitivity of the self correction energy to the angular localization of electronic density may be unphysical because the LDA for the exchange term cannot be justified for the localized density in SH-SICX and C-SICX. The use of spherically averaged density in the SA-SICX moderates to some extent the overamplification of the angular localization dependendence leading to a better approximation. The localization of radial wave functions in atoms is very different from that in Na clusters when one compares the states with the same nodal quantum number. For example, $`1s`$ and $`1p`$ orbitals have very similar radial distibution in Na clusters, while the latter ($`2p`$ in the atomic notation) is much more extended than the former in atoms. Consequently, the self interaction correction mainly originates from the most localized $`1s`$ state in atoms. Moreover, only small angular momentum orbitals appear in atoms. Thus, the problem stated above, i.e. the unphysical strong angular localization dependence does not cause so serious trouble in atoms. To remedy the problem for Na clusters, we propose to utilize spherical harmonic and Cartesian orbital densities only for the self Hartree term. This is a natural consequence of the considerations mentioned above. The exchange energy calculated in this way is given in the lower side in the rows for SH-SICX and C-SICX in Table III. They are designated by the label “(H only)”. All of the H-only exchange energies better agree with that in HF than the corresponding results of the SH-SICX and C-SICX which have been calculated using the spherical harmonic or Cartesian orbital densities both for the self Hartree and exchange terms (the upper side in each row). Especially, H-only SH-SICX is superior to the SA-SICX for most systems. We remark that a similar improvement has been obtained by the H-only method concerning the total energy. ### B Effects of non-diagonal Lagrange multipliers Before we conclude the paper we comment on the effects of off-diagonal Lagrange multipliers in Eq.(11). We performed calculations by keeping off-diagonal Lagrange multipliers for Na<sub>20</sub> and Na<sub>40</sub> with SA-SICX, SH-SICX and C-SICX. We found that the differences of the total energy and single particle energy obtained by self-consistently solving Eq.(11) and Eq.(13) are less than 10<sup>-1</sup> eV and 10<sup>-2</sup> eV, respectively. We therefore conjecture that one can safely ignore the off-diagonal components of the Lagrange multipliers for Na clusters. This result is consistent with that in Ref. for atoms. ## IV Summary and conclusion We have calculated the self interaction corrected exchange energy for Na clusters by using spherically averaged, spherical harmonic and Cartesian orbital densities. We found that both calculations using the spherical harmonic and Cartesian orbital densities deviate from the HF results more than the calculations with spherically averaged orbital densities. The deviation is especially large for systems with large angular momentum orbitals. We attribute this problem to the LDA to evaluate the self exchange term. From this consideration, we propose to use the spherical harmonic and Cartesian orbital deinsities only for the self Hartree term and to use spherically averaged orbital densities for the self exchange term. We have shown that this treatment improves indeed the exchange energy to well reproduce the results of HF calculations. We expect that the remaining errors can be diminished by the SIC using GGA(generalized gradient approximation) . ## Acknowledgement We are thankful to David M. Brink for enlightening discussions.
no-problem/9906/gr-qc9906040.html
ar5iv
text
# 1 Two beer cans representing de Sitter space. Surfaces of constant 𝑡 are pairs of disks representing the two hemispheres of a stereographically projected 2-sphere. The pair of disks on top of the cans represents 𝑖⁺. The shaded surface is a Carter-Penrose diagram, each interior point of which represents a circle. The curves on this surface are light rays. Such diagrams are usually drawn so that the slope of a radial light ray is unity but this is not the case here. Our radial coordinate is chosen so that the slope of any light ray depends on 𝑟 only. 1. INTRODUCTION. In a paper entitled ”Inflationary data for generic spatial topology” Morrow-Jones and Witt were able to show that it is possible to set initial data on three-spaces of generic topology which evolve to locally de Sitter spacetimes under the equation $$R_{\alpha \beta }=\lambda g_{\alpha \beta },\lambda >0.$$ (1) During inflation—if it occurs—this is indeed how the Universe evolves. Given the continuing interest in both non-trivial spatial topologies and inflation it seems worthwhile to explore these results a little further. Let us briefly summarize the main points of Morrow-Jones and Witt: According to a widely believed conjecture , all three-manifolds can be ”glued together” from certain basic building blocks. These building blocks are Thurston’s model geometries. There are altogether eight model geometries of which four are locally spherically symmetric, namely hyperbolic space $`𝐇^3`$, flat space $`𝐄^3`$, the three-sphere $`𝐒^3`$, and the handle $`𝐒^2\times 𝐄`$. Naturally we must allow the topologies that can be obtained by taking the quotient of a building block with some discrete isometry group $`\mathrm{\Gamma }`$. In a well defined sense ”generic” spatial topologies are necessarily of the form $`𝐇^3/\mathrm{\Gamma }`$ , so including further model geometries makes the topology generic by default. The first observation in ref. is that all the four model geometries that we have mentioned can be embedded as spatial hypersurfaces in de Sitter space, so that locally de Sitter initial data can be set on all of them as well as on their quotients. Next the ”gluing together” is done by taking a connected sum, which means that one removes a ball from each piece and identifies the resulting boundaries. While it is clear that this can be done in a way that preserves the local spherical symmetry of the three-metrics, it is not obvious that the full set of initial data for Einstein’s equations can be chosen so that spacetime remains locally de Sitter. Indeed since there are elliptic constraint equations one expects the presence of a handle on a sphere (say) to affect the geometry all over the sphere, so that attaching a second handle might prove difficult. Nevertheless Morrow-Jones and Witt were able to show that such ”Machian” behaviour is in fact absent provided that the minimum radius $`R`$ of the handles is large enough. To show that the connected sums can indeed evolve to locally de Sitter spacetimes Morrow-Jones and Witt present a slightly complicated argument that relies on a cosmic Birkhoff theorem and analyticity in the various charts. In this paper we will rederive and extend the above results in two different ways: First we rewrite the de Sitter metric by reparametrizing two of the coordinates. This brings two arbitrary functions of each coordinate into the metric, and we think of these two functions as providing the general solution of the two dimensional wave equation. Effectively, this means that we can solve the main problem of ref by choosing initial data for the wave equation rather than for Einstein’s equation. (This method can be applied in more general situations but turns out to be useful here only because the de Sitter metric is so simple.) Then we present a pictorial description of de Sitter space that allows us to drop the restriction of local spherical symmetry everywhere, and to see the global time evolution of the data at a glance. We use the pictorial method to address an interesting point raised by Morrow-Jones and Witt (and known to mathematicians ), namely the existence of smooth locally de Sitter domains of development whose universal covers cannot be embedded in de Sitter space. It turns out that $`𝐇^2\times 𝐄`$, the fifth of Thurston’s model geometries, serves as an example. (On the other hand we will find that the example proposed by Morrow-Jones and Witt does not work.) Throughout, a ”spacetime” means the domain of development of some complete smooth initial data surface. It will become clear that most of the spacetimes constructed are complete only in one time direction, chosen to be the future. If we evolve back into the past we will typically encounter either Cauchy horizons, Misner type singularities, or both. In some cases the solution can be analytically extended across the horizon, and then the spatial topology may be revealed as being in some sense not what we thought it was. An example would be an $`𝐇^3`$ initial data surface, for which the solution can be extended to ordinary de Sitter space whose topology is in fact $`𝐒^3\times 𝐑`$. In other cases there are singularities in the past, perhaps making it impossible to continue the solution across the Cauchy horizon. An example would be $`𝐇^3/\mathrm{\Gamma }`$ for a suitably chosen discrete group . We choose to ignore this point, which means that we have nothing to say about the issue of how inflation started. The organization of our paper is as follows: In section 2 we rederive the central result of ref. by writing the de Sitter metric in a form where picking a locally spherically symmetric geometry on an initial data slice is achieved by choosing initial data for the two dimensional wave equation. In section 3 we show—using the hyperboloid model and following Schrödinger —how some of Thurston’s model geometries can be embedded in a de Sitter space. In section 4 we explain how de Sitter space can be visualized in intrinsic terms; from here on the argument relies to a large extent on pictures. Section 5 is devoted to some of the wormhole topologies discussed by Morrow-Jones and Witt . Finally, in section 6 we construct examples of locally de Sitter spacetimes whose universal covers cannot be found as a subset of de Sitter space and in section 7 we summarize our conclusions. 2. A USEFUL FORM OF THE DE SITTER METRIC. We define a locally spherically symmetric three-space as a space that can be covered with charts in which the metric takes the form $$dl^2=a^2(dr^2+f^2(r)d\mathrm{\Omega }^2),$$ (2) where $`d\mathrm{\Omega }^2`$ is the metric on the two-sphere $`𝐒^2`$ and $`a`$ is some constant. Examples include four of the five model geometries mentioned above as well as their connected sums . The point is that the connected sum can be taken by removing a sphere symmetrically placed within a chart of this type. We then choose an $`f`$ that interpolates between what we need for the model geometries that are to be connected together, and we are done. The question to be investigated is therefore: What restrictions on $`f`$ are imposed by the requirement that such a three-space can be embedded in a locally de Sitter space? Our strategy in this section is to reparametrize the de Sitter metric in such a way that it contains a function $`F(t,r)`$; the metric will remain de Sitter if and only if $`F`$ is a solution of the two dimensional wave equation. By inspection we will then be able to see that choosing the function $`f(r)`$ and the constant $`a`$ in the induced 3-metric at constant $`t`$ corresponds to choosing initial data for the wave equation. A restriction on possible forms of $`f`$ will arise from the demand that the 3-metric shall have Euclidean signature. This then will be our answer to the question just asked. We begin with two standard expressions for the de Sitter metric, the static form $$ds^2=\frac{3}{\lambda }\left((1R^2)dT^2+\frac{dR^2}{1R^2}+R^2d\mathrm{\Omega }^2\right)R<1$$ (3) and the Kantowski-Sachs form $$ds^2=\frac{3}{\lambda }\left(\frac{dT^2}{T^21}+(T^21)dR^2+T^2d\mathrm{\Omega }^2\right)T>1$$ (4) where in both cases $`d\mathrm{\Omega }^2`$ is the metric on $`𝐒^2`$. Neither metric covers the entire spacetime, but they are locally de Sitter and this is all we need for our present purposes. The static form of the metric can be taken to the conformal gauge by means of a rescaling of $`R`$: $$\tau =T\mathrm{tanh}\rho =R$$ (5) $$ds^2=\frac{3}{\lambda }\left((1\mathrm{tanh}^2\rho )(d\tau ^2+d\rho ^2)+\mathrm{tanh}^2\rho d\mathrm{\Omega }^2\right).$$ (6) Define $$U=\tau \rho V=\tau +\rho $$ (7) and perform the reparametrization $$U=U(u)V=V(v).$$ (8) The metric then takes the form $$ds^2=\frac{3}{\lambda }\left((1\mathrm{tanh}^2F)\frac{U}{u}\frac{V}{v}dudv+\mathrm{tanh}^2Fd\mathrm{\Omega }^2\right)$$ (9) where $$F=\frac{1}{2}(V(v)U(u)).$$ (10) But this means that $`F`$ is an arbitrary solution of the two dimensional wave equation. Transforming to a new set of coordinates $$u=trv=t+r$$ (11) the metric takes the form $$ds^2=\frac{3}{\lambda }\left((1\mathrm{tanh}^2F)(F^2\dot{F}^2)(dt^2+dr^2)+\mathrm{tanh}^2Fd\mathrm{\Omega }^2\right),$$ (12) where $`F`$ is an arbitrary solution of the two dimensional wave equation in the coordinates $`r`$ and $`t`$, a prime denotes differentiation with respect to $`r`$ and a dot with respect to $`t`$. This is the form of the metric that we were after. Treating the Kantowski-Sachs metric in the same way we obtain $$ds^2=\frac{3}{\lambda }\left((\mathrm{coth}^2F1)(\dot{F}^2F^2)(dt^2+dr^2)+\mathrm{coth}^2Fd\mathrm{\Omega }^2\right).$$ (13) For our purposes this is all we need. To embed a locally spherically symmetric three-space with the metric (2) all that is needed is to choose initial data for the two dimensional wave equation. The initial values of $`F`$ and $`\dot{F}`$ at any fixed value of $`t=t_0`$ are at our disposal, so if $`|af|>\sqrt{3/\lambda }`$ we set $$g_{\theta \theta }=\frac{3}{\lambda }\mathrm{coth}^2F=a^2f^2$$ (14) $$g_{rr}=\frac{3}{\lambda }(\mathrm{coth}^2F1)(\dot{F}^2F^2)=a^2.$$ (15) and we are done: The spatial metric at $`t=t_0`$ is the locally spherically symmetric metric (2), and the initial values determine a solution $`F(r,t)`$ of the wave equation that yields the locally de Sitter metric (13). There are no additional restrictions on the function $`f`$ in this case. The story is different if $`|af|<\sqrt{3/\lambda }`$. This time we set $$g_{\theta \theta }=\frac{3}{\lambda }\mathrm{tanh}^2F=a^2f^2$$ (16) $$g_{rr}=\frac{3}{\lambda }(1\mathrm{tanh}^2F)(F^2\dot{F}^2)=a^2.$$ (17) Now there is an additional restriction since $`r`$ will be a spatial coordinate only if $$g_{rr}>0F^2\dot{F}^2>0.$$ (18) Hence $`\dot{F}`$ cannot be chosen quite arbitrarily. Moreover from the relation between $`f`$ and $`F`$ it follows that $$f^{}=0F^{}=0.$$ (19) Eqs. (18) and (19) imply that we cannot allow the function $`f`$ in the three-metric (2) to attain a minimum or a maximum at a radius smaller than $`\sqrt{3/\lambda }`$. In particular only handles whose minimum radius exceeds $`\sqrt{3/\lambda }`$ can occur. This completes our rederivation of the central result of ref. . The reader may worry that $`|af|`$ is a function, and it may be larger or smaller than $`\sqrt{3/\lambda }`$ depending on its argument. Actually there is no problem here, as an explicit example will make clear. Set $$F=\frac{1}{2}\mathrm{ln}\frac{rt}{r+t},$$ (20) which is a real solution of the wave equation if $`r>t`$. It is easy to show that $$\mathrm{coth}F=\frac{r}{t}$$ (21) and that the locally de Sitter metric (13) becomes $$ds^2=\frac{3}{\lambda }\frac{1}{t^2}(dt^2+dr^2+r^2d\mathrm{\Omega }^2),$$ (22) which represents a portion of de Sitter space foliated by flat spaces. When $`r<t`$ it happens that the argument of the logarithm becomes negative so that $`F`$ develops an imaginary part. This is just what is needed in order to turn the cotangens hyperbolicus into tangens hyperbolicus, so that the other form of the metric applies. Finally we observe that our argument can be varied a little. An alternative form of the intrinsic de Sitter metric is $$ds^2=\frac{3}{\lambda }\left(\frac{dT^2}{T^2+1}+(T^2+1)dR^2+T^2d\sigma ^2\right)$$ (23) where $`d\sigma ^2`$ is the metric on $`𝐇^2`$. Treating this line element in the same way as above we arrive at $$ds^2=\frac{3}{\lambda }\left((1+\mathrm{tan}^2F)(\dot{F}^2F^2)(dt^2+dr^2)+\mathrm{tan}^2Fd\sigma ^2\right).$$ (24) Possible initial data slices now include hyperbolic three-space $`𝐇^3`$ and the hyperbolic handle $`𝐇^2\times 𝐄`$. 3. MODEL GEOMETRIES. This section is intended to remind the reader of some well known facts . If (for convenience) we set $`\lambda =3`$ then de Sitter space can be defined as the hyperboloid $$X^2+Y^2+Z^2+U^2V^2=1$$ (25) sitting inside a five dimensional Minkowski space with the metric $$ds^2=dX^2+dY^2+dZ^2+dU^2dV^2.$$ (26) The isometry group of de Sitter space is therefore the Lorentz group $`SO(4,1)`$. The embedding space coordinates are very useful for many calculations. In particular, let us consider various spacelike hypersurfaces in de Sitter space. Intersecting the hyperboloid with the spacelike hyperplanes $$V=\mathrm{sinh}tX^2+Y^2+Z^2+U^2=\mathrm{cosh}^2t$$ (27) we obtain a foliation of de Sitter space with 3-spheres. These 3-spheres turn out to be totally umbilic (that is to say their first and second fundamental forms are proportional). They are totally geodesic and have unit radius if and only if the spacelike plane goes through the origin. Moving the family of spacelike planes around with isometries it is clear that there is a totally geodesic sphere through any point in de Sitter space. Intersecting the hyperboloid with the timelike hyperplanes $$U=\mathrm{cosh}t$$ (28) we get a foliation of a region of de Sitter space with totally umbilic hyperbolic 3-spaces. (Intersecting with a timelike hyperplane through the origin yields a totally geodesic three-dimensional de Sitter space.) Intersecting with null hyperplanes $$U+V=\frac{1}{t},t>0$$ (29) we obtain a foliation of ”one half” of de Sitter space with totally umbilic flat 3-spaces. Intersection with a null plane through the origin yields a totally geodesic null hypersurface. Two null planes in five dimensional Minkowski space intersect in a 3-space (unless the null planes are parallel), and the intersection of this 3-space with the de Sitter hyperboloid is a 2-sphere. We have seen how three of the model geometries ($`𝐒^3`$, $`𝐇^3`$ and $`𝐄^3`$) can be embedded in de Sitter space. To see that the handle can occur, intersect with the upper branch of the spacelike surface $$V^2U^2=\mathrm{sinh}^2tX^2+Y^2+Z^2=\mathrm{cosh}^2t.$$ (30) This is indeed the handle $`𝐒^2\times 𝐄`$, with the radius of the 2-sphere necessarily larger than one. Finally we can intersect the hyperboloid with the surface $$Z^2+U^2=\mathrm{cosh}^2tX^2+Y^2V^2=\mathrm{sinh}^2t.$$ (31) The intersection has the geometry of $`𝐇^2\times 𝐒^1`$. Now $`𝐇^2\times 𝐄`$ is also one of Thurston’s model geometries. If we can ”unwrap” the circle it follows that this model geometry also can occur in a locally de Sitter spacetime. Actually the function $`F`$ in the metric (24) can be arranged so that this is true but it appears that this spatial geometry cannot be embedded in de Sitter space. We will return to this point in section 6. Our last two examples are not umbilic surfaces but they do inherit enough Killing vectors from de Sitter space to be homogeneous spaces, and their second fundamental forms share the symmetries of the induced metric. 4. VISUALIZING DE SITTER SPACE. We would now like to show some pictures of the above. This can be done if we restrict ourselves to 2+1 dimensions and use the methods of ref. , which have proved quite useful to derive various properties of locally anti-de Sitter spaces. Naturally there are features of 3+1 dimensional spacetimes that cannot be seen in the lower dimension, but as far as the properties of locally spherically symmetric spacetimes are concerned the restriction is harmless. Thus we consider the hyperboloid $$X^2+Y^2+U^2V^2=1$$ (32) in ordinary Minkowski space. We use a function of $`V`$ as an intrinsic time coordinate; at constant $`V`$ space is a sphere. As map-makers know we cannot make a two dimensional map of a sphere unless we give up manifest spherical symmetry; our choice is to make a stereographic projection onto the unit disk of each hemisphere separately. Then we arrive at the coordinate system $`(x,y,t)`$, where $$X=\frac{2x}{1+r^2}\frac{1}{\mathrm{cos}t}Y=\frac{2y}{1+r^2}\frac{1}{\mathrm{cos}t}$$ (33) $$U=\pm \frac{1r^2}{1+r^2}\frac{1}{\mathrm{cos}t}V=\mathrm{tan}t,$$ (34) and we use the upper sign in the Northern hemisphere and the lower sign in the Southern. Note that $$r^2x^2+y^21.$$ (35) The metric in these coordinates is $$ds^2=\frac{1}{\mathrm{cos}^2t}\left(dt^2+\frac{4}{(1+r^2)^2}(dx^2+dy^2)\right).$$ (36) The time coordinate $`t`$ goes from $`\pi /2`$ to $`\pi /2`$, and we can attach a conformal boundary consisting of two spheres at $`t=\pm \pi /2`$. We can now draw pictures in coordinate space. Our first picture shows de Sitter space as two beer cans whose cylindrical boundaries are to be identified. The scaling of the radial coordinate was chosen so that the slope of a light ray is a function of $`r`$ only. We observe that all timelike and null geodesics acquire a future endpoint on future infinity $`i^+`$ and a past endpoint on past infinity $`i^{}`$, while all spatial geodesics are closed and have circumference $`2\pi `$. Totally geodesic surfaces are easy to find since they are the intersection of the hyperboloid with hyperplanes through the origin in embedding space. Let us begin with totally geodesic null surfaces which result if the hyperplane through the origin is chosen to be null. It is easy to check that there is a one-to-one correspondence between such surfaces and the set of lightcones with a vertex on $`i^+`$. (For the twistor theorist this means that $`i^+`$ can be thought of as mini-twistor space.) The next picture shows two such surfaces, chosen so that they are rotationally symmetric in our picture. The isometry group of 2+1 dimensional de Sitter space is the Lorentz group $`SO(3,1)`$, which acts like the group of Möbius transformations on $`i^+`$. In particular a Lorentz boost $`J_{UV}`$ has a bifurcate Killing horizon that coincides with the light cones with vertices at $`r=0`$; the flow is timelike inside the lightcones and spacelike outside. By symmetry all light cones with vertices on $`i^+`$ are Killing horizons. Two important properties follow immediately: First all spacelike loops going around such a cone have equal length (equal to $`2\pi `$ with our choice of $`\lambda `$). As a consequence the intersection of two light cones with vertices on $`i^+`$ is always a circle with circumference $`2\pi `$. Lorentz boosts correspond to hyperbolic Möbius transformations on $`i^+`$ while rotations correspond to elliptic Möbius transformations; rotations have two timelike lines of fixed points which can be made into world lines of point particles, if this is wanted . A general Lorentz transformation can be written as a combination of a boost and a rotation and is called a four screw. It has no fixed point inside de Sitter space but two fixed points on $`i^+`$, where it gives what one calls a loxodromic Möbius transformation. Finally a null boost has a single fixed point on $`i^+`$ and a single light cone for its (degenerate) Killing horizon; there are two lightlike lines of fixed points going through de Sitter space. We may note that if we identify points along the Killing flow lines of a boost (say $`J_{UV}`$ for definiteness) we obtain—in the regions where the flow is spacelike—the spatial topology of a torus. This is the de Sitter analogue of Misner space; the anti-de Sitter analogue is the BTZ black hole . Let us now consider foliations by spacelike surfaces. In particular we want to draw pictures of the two dimensional analogues $`𝐒^2`$, $`𝐄^2`$ and $`𝐒^1\times 𝐄`$ of the three dimensional model geometries discussed in section 3. Note that the distinct three dimensional possibilities $`𝐇^2\times 𝐒^1`$ and $`𝐒^2\times 𝐄`$ collapse to one example only in 2+1 dimensions. We begin with spheres. A foliation with surfaces of constant $`t`$ consists of contracting and expanding totally umbilic 2-spheres; only the pair of disks at $`t=0`$ represents a totally geodesic sphere. Moving this exceptional sphere around with suitable isometries will give a foliation with totally geodesic spheres of the interior of a pair of totally geodesic null surfaces. We will take special interest in surfaces that behave asymptotically like one of the model geometries $`𝐇^2`$, $`𝐄^2`$ and $`𝐒^1\times 𝐄`$. This means that they have to intersect $`i^+`$ in a circle, a point or two points, respectively (Figure 4). As we saw in section 3, intersecting the de Sitter hyperboloid with a family of null planes gives a foliation in terms of flat surfaces, while intersecting with timelike planes sufficiently far from the origin gives a foliation with expanding hyperbolic planes. Following section 3 we can also find a foliation with cylinders; we draw this in two different ways to remind the reader that if a given spatial geometry is displaced by an isometry then its appearance in our pictures will change. In all three cases the foliation covers only a region of de Sitter space. The boundary of this region is the Cauchy horizon, which is a null surface. In the case of the handle the Cauchy horizon grows from a spacelike geodesic which is a circle and has circumference $`2\pi `$. 5. SPACES WITH ASYMPTOTIC REGIONS. Given the pictures of the various model geometries it is straightforward to draw their connected sums. In fact we do not have to insist that the surfaces be everywhere locally spherically symmetric. We can wiggle them this way and that. As an example, let us (with Morrow-Jones and Witt ) consider the problem of attaching asymptotic regions to some spatial slice. Then the thing to remember is that a hyperbolic plane always intersects $`i^+`$ in a circle, while flat planes and handles intersect $`i^+`$ in respectively one and two points. In our picture what we have to do is to deform the sphere so that it touches $`i^+`$ in the ”right” way (Figure 5). It is interesting to ask for the minimum size of the disk that must be removed from the sphere when an asymptotic region is attached. Consider a flat asymptotic region, say. Then we must certainly remove that portion of the sphere that lies in the causal past of the point where the flat space touches $`i^+`$. This is a disk whose circumference is a loop surrounding a light cone with its vertex on $`i^+`$, and as we have seen it necessarily has circumference $`2\pi `$. But we can place the asymptotic region as close to this light cone as we want, and therefore the circumference of the disk that we must remove is bounded from below by $`2\pi `$. We get the same bound if the asymptotic region is a cylinder or a hyperbolic plane. Among the 3-spaces considered by Morrow-Jones and Witt is a set of three asymptotically flat spaces connected by two handles. They make the claim that although this space will (given a suitable form of its extrinsic curvature) evolve to a locally de Sitter spacetime, the universal cover of the domain of development cannot be embedded in de Sitter space. Actually this is incorrect. The argument goes as follows: In four spacetime dimensions the space is simply connected so that we do not have to discuss whether some quotient of de Sitter space will serve as embedding space—it must be de Sitter space itself. This is indeed so. Morrow-Jones and Witt next observe that two flat spaces in de Sitter space necessarily intersect in a sphere (unless they occur at different ”times”). This is also correct—in embedding space it is the statement that two non-parallel null planes will intersect in a 3-space, and this 3-space intersects the hyperboloid in a sphere. Finally it is concluded from the picture of the spatial slice that two of the flat regions do not so intersect—but this is not valid since in fact all the spatial planes do intersect, although the two at the ends do so (as it were) within the disks that were removed in attaching the handles that connect them to the plane in the middle. If this sounds obscure, we hope that it is made clear by a glance at our two beer cans, embellished with a spatial surface carrying precisely the geometry considered by Morrow-Jones and Witt (Figure 6). Although the picture is restricted to 2+1 dimensional de Sitter space the restriction is irrelevant here. Moreover, in drawing this picture we have in effect solved for the time evolution of these initial data. By the way, the restriction on the minimum radius of a connecting handle (given in ref. and in section 2) also follows from an argument that parallels the argument about the minimum size of the disk that must be removed when attaching an asymptotic region from the sphere. Nevertheless there are locally de Sitter spacetimes whose universal covers cannot be regarded as subsets of de Sitter space; we devote section 6 to this topic. 6. WHY DE SITTER SPACE IS NOT ENOUGH. A relativistic space form, by definition, is a complete Lorentzian manifold of constant curvature. They have been completely classified . The universal cover of a relativistic space form is de Sitter space, anti-de Sitter space or Minkowski space depending on the curvature. Our definition of a spacetime was made in the spirit of canonical gravity and does not require completeness. All that we require is the domain of development of some smooth initial data surface. The classification of such spacetimes is a much harder problem and—at least to our knowledge—it has been achieved only in 2+1 dimensions . In this connection it was pointed out by Mess that there are smooth locally de Sitter domains of development (complete to the future) whose universal covers cannot be embedded in de Sitter space. So how can we obtain a smooth surface carrying locally de Sitter data that cannot be embedded in de Sitter space? This is not so difficult. Suppose that we remove a pair of antipodal points from a sphere so that its topology becomes that of a cylinder, and then go to its universal covering space. A similar manœuvre carried through for 2+1 dimensional de Sitter space entails removing two antipodal timelike lines and going to the universal covering space of the remaining (incomplete) spacetime. To draw the picture it is convenient to place the timelike lines to be removed on the boundary of the cans. The universal covering space then consists of an infinite set of cans with lines cut out of the boundary; the sides of the cans are to be identified pairwise in an obvious manner (Figure 7). Now consider a particular can and insert a ”handle” (actually half a handle) that touches what used to be $`i^+`$ in precisely the two points that we have removed from the top. Draw a similar surface in all the other cans. This is a smooth spacelike surface of topology $`𝐑^2`$—in effect, we have ”unwrapped” the cylinder. Its intrinsic geometry is flat but unlike the family of embedded flat spaces considered earlier it is not an umbilic surface. Moreover its Cauchy development is smooth and locally de Sitter by construction, and it is complete to the future. If we try to add a conformal boundary $`i^+`$ we run into problems, but then a smooth conformal completion of the domain of development was never promised. In its past there is a Cauchy horizon that grows from a non-closed spacelike geodesic of infinite length—which is one way to see that this spacetime definitely cannot be embedded in de Sitter space since the latter does not have such geodesics. Hence this is the example that we were after. Let us mention in passing that if we start from the de Sitter analogue of Misner space (as briefly described in section 4) and then ”cut out a wedge” bounded by timelike surfaces—that is, if we identify points along the flow lines of the Killing vector $`J_{XY}`$—then the spatial topology is again a torus, but a torus whose covering space (in the generic case) is the spacetime that we have just discussed. The example just given works specifically in 2+1 dimensions, but it is not difficult to modify it so that it works in 3+1 dimensions. We start out by considering the model geometry $`𝐇^2\times 𝐒^1`$ embedded as a hypersurface in de Sitter space through $$Y^2+U^2=\mathrm{cosh}^2\tau X^2+Z^2V^2=\mathrm{sinh}^2\tau .$$ (37) Our aim is to ”unwrap” the circle. The Cauchy development of the embedded hypersurface is the future of the spatial geodesic $$X=Z=V=0.$$ (38) The circle becomes unwrapped if we are able to unwrap the flow lines of the Killing vector $`J_{YU}`$ (that generates rotations). To succeed in this we must remove the fixed points of this Killing vector from de Sitter space; in equations the fixed points are given by $$J_{YU}=Y_UU_Y=0Y=U=0.$$ (39) Metrically this is a 1+1 dimensional de Sitter space; for us it is important to observe that it has zero intersection with the Cauchy development of the embedded hypersurface. On the other hand the fixed point set and the embedded hypersurface touch $`i^+`$ in the same circle. If we think of de Sitter space as a set of foliating 3-spheres we see that the fixed point set gives a circle in each 3-sphere—and a circle is precisely what we must remove in order to make a 3-sphere multiply connected. Therefore once the fixed point set has been removed we have an incomplete, multiply connected spacetime and we can go to its universal covering space. Having done this we have an embedding of Thurston’s model geometry $`𝐇^2\times 𝐄`$ in an incomplete spacetime that is ”larger” than de Sitter space; nevertheless the Cauchy development of our model geometry is not only simply connected but also complete to the future. This Cauchy development is in itself a spacetime in our sense. To visualize this construction one can draw a series of equal-$`t`$ slices of 3+1 dimensional de Sitter space, that is a series of stereographically projected 3-spheres, and study how the hypersurface $`𝐇^2\times 𝐒^1`$ intersects these slices. In effect all that one has to do is to take equal-$`t`$ slices of fig. 7 and rotate them around a suitable axis. The details are left to the interested reader. We have embedded the model geometry $`𝐇^2\times 𝐄`$ in a locally de Sitter spacetime which cannot itself be found as a subset of de Sitter space. To see that there is no other way to embed this model geometry in de Sitter space we may use eq. (24) to prove that its extrinsic geometry is determined by the requirement that it is embedded in a locally de Sitter space. Therefore its Cauchy development has to be precisely the spacetime that we just constructed, and this spacetime cannot be a subset of de Sitter space since—like its 2+1 dimensional counterpart—it grows from an infinitely long non-closed spacelike geodesic. 7. CONCLUSIONS. Our conclusions can be summarized like this: One can write the de Sitter metric in a form that allows one to choose any allowed locally spherically symmetric geometry on a spacelike slice by choosing initial data for the two dimensional wave equation. A similar form exists for spatial geometries of the form (hyperbolic plane)$`\times `$ (something). Using a pictorial presentation the restriction to local spherical symmetry everywhere can be dropped and the Cauchy development of any spatial slice is easily studied. The model geometry $`𝐇^2\times 𝐄`$ provides an explicit example of a smooth locally de Sitter domain of development whose universal cover cannot be embedded in de Sitter space. On the other hand three flat asymptotic regions connected by two wormholes can be so embedded (pace previous claims). As a final comment we observe that de Sitter space is very different from the other two relativistic space forms, Minkowski space and anti-de Sitter space. In de Sitter space asymptotic regions of spacelike slices occur where the slices approach future infinity. Hence the issue of connecting different asymptotic regions with causal curves does not even arise. A little thought will also convince the reader that our examples of simply connected locally de Sitter spacetimes that cannot be embedded in de Sitter space came about precisely because the conformal boundary (or more accurately the would be conformal boundary) is a spacelike surface. Hence this behaviour cannot occur for locally flat or anti-de Sitter spacetimes—a fact which is perfectly well known , but perhaps we have managed to convey some extra feeling for why it is true. Acknowledgements: We thank Dieter Brill for suggesting this problem, and Stefan Åminneborg and Greg Galloway for discussions and help. IB was supported by the NFR.
no-problem/9906/quant-ph9906079.html
ar5iv
text
# Symmetry of Quantum Phase Space in a Degenerate Hamiltonian System ## I Acknowledgments We are thankful to D.F.V. James and G.D. Doolen for useful discussions. The work of V.Ya.D. and D.I.K. was partly supported by the Russian Foundation for Basic Research (Grants No. 98-02-16412 and No. 98-02-16237). Work at Los Alamos National Laboratory was partly supported by the National Security Agency, and by the Department of Energy under contract W-7405-ENG-36.
no-problem/9906/cond-mat9906195.html
ar5iv
text
# Effect of Dilution on First Order Transitions: The Three Dimensional Three States Potts Model ## I Introduction The study of the impurities effect on the critical behavior of a pure material is an important issue since frequently real systems cannot be considered as pure. Nowadays the effect of dilution (disorder coupled to the energy density) on second order phase transitions is well understood. The Harris’ criterion states that if the specific-heat of the pure system presents a power-like divergence (i.e. $`\alpha _{\mathrm{pure}}>0`$) the disorder induces a new Universality Class. Otherwise ($`\alpha _{\mathrm{pure}}<0`$) the critical behavior of the model remains unchanged. The criterium does not decide in the marginal case $`\alpha _{\mathrm{pure}}=0`$. Moreover, it is possible to show rigorously that for all the continuous phase transitions in presence of disorder, the correlation-length critical exponent $`\nu `$ verifies $`\nu 2/d`$, $`d`$ being the dimensionality of the space . When the pure model shows a first order phase transition the situation is more complicated. However, there are a set of important results, both numerical and analytical. For instance, Aizenman and Wehr showed rigorously that in two dimensions when introducing disorder, its conjugated density becomes a continuous function of the thermodynamic parameters. For instance, in a site diluted model the conjugate density is the energy, while in the Random Field Ising Model (RFIM), it is the magnetization. If the phase transition becomes continuous, one may ask about its Universality Class. A widely studied model in this context has been the $`q`$-states Potts model, whose pure version in $`d=2`$ undergoes a first order phase transition for $`q5`$. In recent numerical simulations using Monte Carlo or transfer matrix methods, the $`\nu `$ exponent has been found compatible within errors with the pure Ising value ($`\nu =1`$) independently of $`q`$, but for the magnetic exponent, $`\beta `$, the numerical results for $`q>2`$ are significantly different from the pure Ising value . Unfortunately, for $`d>2`$ the situation is not so clear. Cardy and Jacobsen (see also Cardy ) have put forward a picture of the general behavior by means of a mapping from the diluted $`q`$-states Potts model to the RFIM. Their mapping being asymptotically exact for large $`q`$, their results are also expected to hold for phase transitions with large latent heat. According to them, when a system that undergoes a first-order phase transition gets weakly diluted, the latent heat decreases. For larger dilutions, it will eventually vanish at a so-called tricritical point. Cardy and Jacobsen relate the latent heat with the order parameter of the RFIM, whose behavior is governed by the zero-temperature fixed point. In this way they are able to predict the critical exponents for the tricritical point $`\beta =\beta _{\mathrm{RFIM}}`$ and $`\nu =\nu _{\mathrm{RFIM}}/(2\alpha _{\mathrm{RFIM}}\beta _{\mathrm{RFIM}})`$. The critical-behavior of the system for larger dilutions remains unaddressed in their work. A physical realization of this scenario is provided by some magnetic semiconductors like Zn<sub>1-x</sub>Mn<sub>x</sub>Te . The magnetic atoms of these materials behave as Heisenberg spins living in a fcc lattice, with antiferromagnetic interactions. In these highly frustrated systems a first order phase transition is found in pure samples that gets second order upon dilution. We finish this overview describing the results obtained by Elderfield and Sherrington for the diluted Potts Model in the Mean Field approximation . They found that the phase transition is first order for $`q>2`$, for all values of spin-concentration. Another interesting and related model is the Potts glass for $`q>4`$ that, according to Mean Field theory, undergoes a first order transition with no latent heat while for $`q=3`$ and $`4`$ the transition is continuous . In this work, we shall consider the effects of site-dilution in the three dimensional three state Potts model, whose pure version presents a weak first order transition (small latent heat or large correlation length at the critical point). Our choice has been motivated by the ubiquity of weak first order transitions in nature. In particular, the pure $`q=3`$ Potts model shows very different experimental realizations appearing in very distant fields. For instance, we can cite the de-confining phase transition in quenched Quantum Chromodynamics. It can also characterize different situations in solid state physics. For instance a cubic ferromagnet with three easy axes of magnetization when a magnetic field in the diagonal of the cubic lattice is turned on (e.g. DyAl<sub>2</sub>), structural phase transitions (e.g. SrTiO<sub>3</sub>) and some fluid mixtures of five (suitable chosen) components . We now describe briefly the phase diagram of the three dimensional three states Potts model in the temperature–concentration plane $`(T,p)`$ (see Fig. 1). The pure model undergoes a (weak) first order phase transition, at a critical temperature $`T_\mathrm{c}(p=1)`$ separating the paramagnetic high-temperature phase from the low temperature ordered one. This first order transition can be, in principle, continued inside the $`(T,p)`$ plane, where the critical temperature $`T_\mathrm{c}(p)`$ will lower for smaller $`p`$. The latent-heat for the first-order phase transition will decrease until the tricritical point. At this point the model suffers a second order phase transition that continues (belonging to another Universality Class) until the $`T_\mathrm{c}(p_\mathrm{c})=0`$ percolation limit. We remark that, in the most economic picture, this phase diagram presents three different Universality Classes: site percolation in three dimensions (which has been studied in the literature, e.g. in Ref. ), the Universality Class of the tricritical point (conjectured in Ref. ) and the Universality Class that controls the critical behavior in the region between the tricritical point and the percolation point. In this paper we will restrict ourselves to the study of the second-order line. Although an experimental realization of the site diluted Potts model is not yet known (disorder tends to couple with the order parameter rather than with the energy), whenever it will appear it will be interesting to have clear theoretical predictions at hand. The techniques used in this paper are well suited for second order transitions, but they should be modified in the concentration range for which the phase transition is first order. Work is in progress to study this region . ## II The Model and Observables We have studied the three dimensional Site Diluted three state Potts Model, whose Hamiltonian defined on a cubic lattice with volume $`V=L^3`$ is $$=\mathrm{R}e\left[\underset{<i,j>}{}ϵ_iϵ_jz_iz_j^{}\right],$$ (1) and periodic boundary conditions are applied. In Eq. (1) $`z_i`$’s are complex roots of $`z^3=1`$, and $`ϵ_i`$’s are uncorrelated quenched random variables which are $`1`$ with probability $`p`$, and $`0`$ with probability $`1p`$. The Boltzmann weight is proportional to $`\mathrm{exp}(\beta )`$. We have used clusters algorithms in order to update the system. In a diluted system, the set of occupied sites can present regions that are lightly connected to the percolating cluster. These regions are very difficult to equilibrate just with a single cluster algorithm . We have found that a single cluster algorithm combined with a Heat Bath sweep per measure is efficient for large concentrations. However, for small concentrations ($`p0.5`$) the previous method is not efficient enough due to the presence of intermediate-sized clusters, and we have used the Swendsen-Wang algorithm . We have simulated at $`p=1.0,0.9,0.8,0.7,0.6,0.5`$ and $`0.4`$ in lattices $`L=8,16,32`$ and $`64`$. For $`p=0.8,0.7,0.4`$ we have also run in $`L=128`$ lattices. We have performed $`N_I=200`$ nearly independent measures in every single disorder realization. For $`p0.8`$ the number of these realizations has been $`N_S=10000`$, except for $`p=0.8`$, $`L=128`$, where we have fixed $`N_S=1000`$. In the $`p=0.9`$ case we have measured in 2000 different disorder realizations. The total amount of CPU time has been the equivalent of 16 years of 200 Mhz Pentium-Pro processor. For small dilutions we have performed the usual $`\beta `$ extrapolation while for $`p0.5`$ we used a $`p`$ extrapolation method . Let us recall that when planning a disordered model simulation, one should balance two competing effects. First, to minimize statistical errors, it is better to work in a $`N_IN_S`$ regime. On the other hand, if $`N_I`$ is too small, the usual calculation of $`\beta `$ derivatives and extrapolations is biased. We follow the same procedure of Ref. to eliminate the bias. With our simulation strategy ($`N_IN_S`$), it is crucial to check that the system is sufficiently thermalized while taking measures. In order to ensure this, we have systematically compared the results coming from hot and cold starts: half of our statistics for the largest lattices have been obtained with hot starts while the other half comes from cold starts. Regarding the observables, in addition to the energy we have measured the complex magnetization $$M=\underset{i}{}ϵ_iz_i,$$ (2) from which we obtain the real susceptibility as $$\chi =\frac{1}{V}\overline{|M|^2}.$$ (3) We have denoted with $`(\mathrm{})`$ the thermodynamical average with fixed disorder and with $`\overline{(\mathrm{})}`$ the average over the disorder. The formulæ for the cumulants read $`g_2`$ $`=`$ $`{\displaystyle \frac{\overline{|M|^2^2}\overline{|M|^2}^2}{\overline{|M|^2}^2}},`$ (4) $`g_3`$ $`=`$ $`{\displaystyle \frac{\overline{M^3}}{\overline{|M|^2^{3/2}}}},`$ (5) $`g_4`$ $`=`$ $`2{\displaystyle \frac{\overline{|M|^4}}{\overline{|M|^2^2}}},`$ (6) $`g_4`$ being the standard Binder cumulant, $`g_2`$ measures whether the susceptibility is or not a self-averaging quantity and $`g_3`$ has been introduced since the three states Potts model is invariant under a global transformation of the $`Z_3`$ group. The other cumulants, $`g_2`$ and $`g_4`$, are also trivially invariant since we have used the modulus of the complex magnetization in their construction. We have used a quotient method , in order to compute the critical exponents. We recall briefly the basis of this method. Let $`O`$ be a quantity diverging as $`t^{x_O}`$ ($`t`$ being the reduced temperature) in the thermodynamical limit. We can write the dependence of $`O`$ on $`L`$ and $`t`$ in the following way $$O(L,t)=L^{x_O/\nu }\left[F_O\left(\frac{L}{\xi (\mathrm{},t)}\right)+𝒪(L^\omega ,\xi ^\omega )\right],$$ (7) where $`F_O`$ is a (smooth) scaling function and $`(\omega )`$ is the biggest non positive eigenvalue of the Renormalization Group transformation (the corrections-to-scaling exponent). This expression contains the not directly measurable term $`\xi (\mathrm{},t)`$, but if we have a good definition of the correlation length in a finite box $`\xi (L,t)`$, Eq. (7) can be written $$O(L,t)=L^{x_O/\nu }\left[G_O\left(\frac{\xi (L,t)}{L}\right)+𝒪(L^\omega )\right],$$ (8) where $`G_O`$ is a smooth function related with $`F_O`$ and $`F_\xi `$ and we have neglected the term $`\xi _{\mathrm{}}^\omega `$ because we are working deep in the scaling region. The definition of the correlation length on a finite box that we use is : $$\xi =\frac{\sqrt{\chi /F1}}{2\mathrm{sin}(\pi /L)},$$ (9) where $`\chi `$ was defined in Eq. (3) and $`F`$ is given by $$F=\frac{V}{3}\underset{𝒌=\frac{2\pi }{L}}{}\overline{|\widehat{z}(𝒌)|^2},$$ (10) $`\widehat{z}(k_1,k_2,k_3)`$ being the discrete Fourier transform of $`ϵ_iz_i`$. We remark that the definition in Eq. (9) makes sense as a correlation length only in the pure paramagnetic phase of the model. The main formula of the quotient method is $$Q_O|_{Q_\xi =s}=\frac{O(sL,t)}{O(L,t)}=s^{x_O/\nu }+𝒪(L^\omega ),$$ (11) e.g. we compute the quotient between $`O(sL,t)`$ and $`O(L,t)`$ at the reduced temperature, $`t`$, in which $`\xi (sL,t)/\xi (L,t)=s`$. As particular cases of interest we cite the susceptibility, $`\chi `$, and the $`\beta `$-derivative of the correlation length, $`_\beta \xi `$, whose associated exponents are: $`x_{_\beta \xi }`$ $`=`$ $`1+\nu ,`$ (12) $`x_\chi `$ $`=`$ $`(2\eta )\nu .`$ (13) A clean measure of scale invariance is provided by $`(\xi /L)|_{Q_\xi =s}`$. Let us recall that $`\xi /L`$ is a monotonically growing function of the inverse temperature. In the ordered phase it grows as $`L^{d/2}`$, while in the disordered phase decreases with growing lattice size. Therefore, for any pair of lattice sizes, there is a crossing temperature where $`Q_\xi =2`$. In a second order transition, $`\xi /L`$ at the crossing point should tend to a non-vanishing universal value. For a first order transition, the crossing temperatures tend to the transition point but $`\xi /L`$ at the crossing diverges due to the coexistence of ordered and disordered phases. We finally analyze the quotient of the cumulants $`g_2`$, $`g_3`$ or $`g_4`$ at two different lattices, $`L`$ and $`sL`$, computed at the temperature where $`Q_\xi =s`$. Notice that for a second order phase transition the asymptotic limit ($`L\mathrm{}`$) of these quotients is $`1`$ corrected by terms like $`L^\omega `$ (see Eq. (11)). The quotient method, Eq. (11), has several interesting features. First, we profit of the large statistical correlation between $`Q_O`$ and $`Q_\xi `$. Next, one does not need a previous estimate of the infinite volume critical point. Finally, it allows a simple control of the scaling corrections. All of this makes the method specially efficient for the measures of anomalous dimensions. ## III Numerical Results The phase diagram that we have obtained numerically is shown in Fig. 1 and it agrees with the standard picture. The dashed line corresponds to the part of the transition line where we do not find a clear second order asymptotic behavior. A clear first order signature is very difficult to see even for dilutions as small as $`p=0.95`$. In this region the individual samples usually present a double peak structure. Consequently the thermalization process is very slow and a different algorithm for updating must be used. Work is in progress to analyze this region . Our scope is now to compute the critical exponents in the region in which the transition is clearly second order, i.e. the study of the Universality Class between the tricritical and percolation limits. The first stage is to determine where an asymptotic second order behavior has been reached with lattice sizes up to $`L=128`$. In Fig. 2 we show the value of $`\xi /L`$, at the points for which $`Q_\xi =2`$ for the different $`(L,2L)`$ lattice pairs and as functions of $`L^\omega `$. We have used for $`\omega `$ the corresponding value of the site diluted Ising model . For $`p0.7`$ we find that $`\xi /L`$ seems to tend to a dilution-independent value. Notice the clear divergence for $`p=1`$, where the transition is known to be first order. For $`p=0.9`$ we find a similar trend that for the pure case while for $`p=0.8`$ we find a transient behavior: for small lattices $`\xi /L`$ grows while in the largest lattices it seems to approach the universal value. Another interesting quantity in order to clarify the second order behavior is the cumulant $`g_3`$. The $`g_3(L)`$ values at the points where $`\xi (2L,t)=2\xi (L,t)`$ are shown in Fig. 3. In this case we see a different scaling behavior for $`p=0.9`$ and $`p=0.8`$ up to the studied lattice sizes. We also guess from this figure that the $`\omega `$ value cannot be much larger than 0.4. We have next considered the quotients of the different cumulants $`g_i`$ at the points where $`Q_\xi =2`$. We recall that these quantities should go to 1 as $`L`$ tends to infinity in a second order phase transition. We present our results in Fig. 4. At concentrations $`p=0.9`$ and $`p=0.8`$ we do not find an asymptotic behavior. For $`p=0.7`$, the behavior is yet not monotonous. Only for $`p=0.4,0.5`$ and $`0.6`$ it seems that the asymptotic behavior is reached. Unfortunately, a reliable estimate of $`\omega `$ cannot be obtained but our results point to a value near $`0.4`$. Moreover, the higher order scaling corrections are rather strong for these quantities. Finally, let us remark that the corrections to scaling and statistical errors are much larger for $`g_2`$ and $`g_4`$ than for $`g_3`$. Therefore, for the study the second order region, we conclude that only for $`p0.6`$ an asymptotic scaling behavior for the considered lattice sizes has been found. We report the results for the critical exponents as functions of $`p`$ and $`L`$ in tables I and II. We have applied Eq. (11) with $`s=2`$ to $`_\beta \xi `$ for computing $`\nu `$ and to $`\chi `$ for extracting $`\eta `$. We can observe that the asymptotic behavior of these estimates for $`p0.7`$ is not clear. We have plotted the $`\nu `$ and $`\eta `$ apparent critical exponents for $`p=0.6,0.5`$ and $`0.4`$ as functions of $`1/L`$ in Figs. 5 and 6 respectively. We recall that we have found a $`\omega 0.4`$ value for the cumulants. From the Figs. 5 and 6 we see that $`\omega =1`$ for the leading scaling-corrections term could be a reasonable choice in this case. A possible explanation of this contradiction could be that for the observables used for computing the critical exponents the leading term ($`\omega 0.4`$) vanishes. In any case we should remark that we have not a precise control over the scaling corrections unlike, for example, in the investigation of the three dimensional site diluted Ising model . Fortunately, the scaling corrections for the critical exponents are rather small. Thus, it is not essential in this model to perform an infinite-volume extrapolation of our estimates. This is in marked contrast with the Ising case, where the extrapolation procedure was crucial to correctly compute the critical exponents. From Figs. 5 and 6 we estimate the critical exponents in the second order region as the displayed in the last row of the table III. We remark that the $`\nu `$ value is indistinguishable from the 3D site diluted Ising model one. However, the $`\eta `$ value is very different from the values found for the rest of the models reported in the table III. ## IV Conclusions We have numerically studied the three dimensional site diluted three states Potts model. The phase diagram in the temperature-concentration plane consists of a ferro-magnetically ordered phase separated from a paramagnetic, high temperature one. Between both regions there is a critical line, which is (weakly) first order in the limit of pure samples. For small concentrations, a clear second-order behavior is found, while the region with $`p0.9`$ shows a different behavior, probably corresponding to a cross-over, more difficult to analyze. The critical exponents of the second order region have been computed using Finite Size Scaling Techniques. We have found that these exponents are dilution independent, and that they show a very mild evolution with the lattice size. That is why a sound estimate of the critical exponents can be given, in spite of the fact that we have been unable of measuring the scaling-corrections exponent $`\omega `$. This is in marked contrast with the situation in the site-diluted Ising model, where the scaling-corrections are severe but $`\omega `$ can be obtained with a $`15\%`$ accuracy. Regarding the variation of the critical exponents with $`q`$ we have compared the results for the Potts model and the Ising case. We have found that the $`\nu `$ exponent varies slowly (or perhaps remains unchanged) with the $`q`$ value whereas the $`\beta `$ or $`\eta `$ exponents show a clear variation on this parameter. This picture strongly reminds the obtained in numerical simulations in two dimensions. The study of the first-order region and of the critical behavior in the neighborhood of the tricritical point requires rather different numerical techniques and will be the matter of future work. ## V Acknowledgments We gratefully acknowledge discussions with D. Belanger, J. Cardy and H. Rieger. We thank partial financial support from CICyT (AEN97-1708 and AEN97-1693). The computations have been carried out using the RTNN machines (Universidad de Zaragoza and Universidad Complutense de Madrid) and the ORIGIN2000 at the Centro de Supercomputación Complutense (CSC).
no-problem/9906/astro-ph9906278.html
ar5iv
text
# Magnitude, color and spectral type of Aql X-1 in quiescence Based on observations obtained at the European Southern Observatory, La Silla, Chile ## 1 Introduction Recent $`K`$-band imaging with Keck I of the transient low-mass X-ray binary Aquila X-1 (= V1333 Aql) has resolved V1333 Aql into two stars lying approximately along the east-west direction and separated by 0.46″ (Callanan et al. Call (1999)), the easterly star contributing 60% of the combined flux at $`K`$. Observations in the $`z`$-band (1.05$`\mu `$) indicated also that the easterly star was ”somewhat bluer” and they speculated that it might be the true optical counterpart of Aql X-1. In 1999 we obtained $`V`$ and $`I`$-band frames of Aql X-1 both in quiescence and during outburst maximum with EFOSC-2 at the ESO 3.6-m telescope. We present here the results of our photometry and of our EFOSC-1 spectroscopy, and we show that the westerly star is the true counterpart. ## 2 Photometry ### 2.1 Quiescence Three dithered 5-min $`V`$-band exposures and 3 dithered 5-min $`I`$-band exposures (using actually a Gunn $`i`$ filter) were obtained on 1999 April 9 with the imaging spectrograph EFOSC-2 at the Cassegrain focus of the ESO 3.6-m telescope at La Silla, using a Loral/Lesser CCD (#40) with 15$`\mu `$ pixels, yielding a projected pixel size of 0.157″. Conditions were clear, the seeing was 1″ for the $`V`$ frames and 0.8″ for the $`I`$ frames. After standard MIDAS reduction procedures, the medianed $`V`$ and $`I`$ frames were measured using the DAOPHOT NSTAR routine of Stetson (Stet (1987)). A clean point spread function (PSF) profile was derived from three nearby isolated stars after three neighbour-removing iterations. Two comparison stars, C1 and C2, situated (12.7″N, 8.7″S) and (14″N, 25″S) respectively from Aql X-1, were used to derive the $`V`$ and $`I`$ magnitudes. Their Johnson-Cousins magnitudes, measured at Observatoire de Haute-Provence with the 1.2-m telescope, are $`V`$ = 17.48, $`I`$ = 16.12 for C1 and $`V`$ = 17.42, $`I`$ = 16.29 for C2 (Chevalier and Ilovaisky 1999c ). We estimate the accuracy of these magnitudes to be 0.05 mag in each band. A close-up view (10″$`\times `$10″) in the $`I`$-band of the stars around Aql X-1 (star $`a`$) is shown in Fig. 1a while Fig. 1b shows the same image after subtraction of the PSF profile at the location of stars $`a`$, $`b`$ and $`c`$, revealing the interloper star $`e`$, in addition to star $`d`$. The NSTAR fitting process, which treats the positions as input parameters to be optimized, was incomplete at this point as the subtracted image shows strong residuals, including star $`e`$. An estimate of the goodness of fit can be derived from the parameters CHI and SHARP produced by the NSTAR routine. CHI is the ratio of the observed pixel-to-pixel scatter in the fitting residuals to the expected scatter based on the values of read-out noise and gain, and should not exceed unity if the fit is good. SHARP measures the difference between the half-width at half-maximum for a star and that for the PSF and is close to zero for isolated stars. When only stars $`a`$, $`b`$ and $`c`$ were included in the fit (Fig 1b), star $`a`$ yielded CHI = 4.62 and SHARP = 0.027, indicating a poor fit. Fig. 1c shows star $`d`$ which is left after subtraction of stars $`a`$, $`e`$, $`b`$, $`c`$ and Fig. 1d shows the cleaned image after removing the profiles fitted to $`a`$, $`b`$, $`c`$, $`d`$, and $`e`$. Inclusion of star $`e`$ (Fig 1d) in the fit yielded CHI = 0.85 and SHARP = 0.007 for both stars $`a`$ and $`e`$, showing a good fit. The same procedure was applied to the medianed $`V`$ frame. The final results of the NSTAR photometry are given in Table 1. The main sources of the errors given in Table 1 are the uncertainty on the absolute calibration (zero-point) of the comparison stars C1 and C2 (color terms are relatively small) and the error on the NSTAR fitting for stars $`e`$ and $`d`$. These determinations are of much better quality than our previous estimates (Chevalier and Ilovaisky 1999a ) based on the 1989 $`V`$-band frames which were obtained with a seeing of 1.2-1.3″ and a projected pixel size of 0.33″. On these frames star $`d`$ was undetected and the magnitude of star $`e`$, which appeared as a faint residual after NSTAR fitting to the $`a`$, $`b`$, $`c`$ group of stars, was underestimated. Under these conditions, the measured magnitude for star $`a`$, $`V`$ = 19.26, is an overestimate by 0.1-0.15 magnitude due to the contamination by star $`e`$ and by the surrounding objects. Measurements obtained with smaller telescopes are affected in a similar fashion, depending on the projected pixel size and seeing. ### 2.2 Outburst During early 1999 May, the source started a new outburst (Jain et al. Jain (1999), Chevalier and Ilovaisky 1999b ) and on May 21, near outburst maximum, we secured 1-min CCD frames in $`V`$ and $`I`$ with EFOSC-2. The FWHM of the image profiles was 1.1″ in $`I`$. Inspection of the frames showed that the barycenter of the variable object image in outburst did not coincide with the position of star $`a`$. We analyzed the outburst frames with DAOPHOT using as a starting point the table of star positions derived by NSTAR from the quiescent frames. From the positions of eight near-by objects we find that the variable object is located $``$0.05 $`\pm `$0.03″ in right-ascension and $``$0.03 $`\pm `$0.06″ in declination from the position of star $`e`$, which is located 0.48″ West of star $`a`$ (Table 1). The lower signal-to-noise ratio in these short exposures did not allow accurate photometry of star $`a`$ but subtraction of a PSF profile fitted to star $`e`$ shows a residual compatible with star $`a`$. On these images, star $`e`$ has $`V`$ = 17.03 and $`VI`$ = 1.03, with a slight contamination from $`a`$. Fig. 2 shows the same field of view as Fig. 1 with the outburst frame shown as a positive print and the quiescent frame overlaid as a negative print. The image in activity appears shifted by 0.42 $`\pm `$0.03″ to the West from the position of star $`a`$, as measured in quiescence. ## 3 Spectroscopy We obtained several spectra of Aql X-1 in quiescence with the imaging-spectrograph EFOSC-1 at the Cassegrain focus of the ESO 3.6-m telescope. They were taken through the B300 grism and cover the range between 3850Å and 6850Å. Two spectra were obtained during the night of 19 May 1988 with 7Å resolution (using CCD #11) and exposure times of 3600s and 4200s. Two more spectra were obtained one year later on 8 May 1989 with 3.5Å resolution (using CCD #8) and exposure times of 3600s and 2700s. The second 1989 spectrum was underexposed and will not be discussed further. Weather conditions were good throughout, with seeing of 1.1″ and 1.3″ respectively, and we used a slit width of 1.5″. Spectra of bright stars of known spectral type from mid-GV to early MV (G3: HD 168402, G5: G 724.1, K0: HD 171982, K3: HD 87521, K7: G 747.3, M0: CD $``$36°6589) were also taken during the 1988 run. All frames were wavelength-calibrated using He-Ar lamp spectra and flat-fielded using internal Tungsten lamp spectra and were reduced using standard MIDAS procedures. The spectra were extracted with the MIDAS long-slit package and corrected for atmospheric extinction using average values for La Silla. The standard star LTT 7987 (Stone and Baldwin Sto (1983)) was used for relative flux calibration. The three spectra of Aql X-1, which did not show any significant differences, were then co-added. Since our observed G8 V star (CD $``$45°12143) turned out to be of a K0 type, we constructed a synthetic G8 V spectrum by averaging our G5 and K0 spectra. The slit included both stars $`a`$ and $`e`$. As determined using Table 1, the flux coming from star $`e`$ is 12% of the total flux from stars $`a`$ and $`e`$ in the $`V`$-band. To derive the spectral type of star $`a`$, we constructed a set of template spectra by adding, to each of our observed spectra of known spectral type, normalized to 0.88 at 5500Å, the observed K7 or M0 spectrum, normalized to 0.12 at 5500Å (see Table 1 and also the Discussion). We applied different corrections for interstellar reddening to the ($`a`$+$`e`$) spectrum, corresponding to $`E(BV)`$ from 0.3 to 0.6 and normalized the de-reddened ($`a`$+$`e`$) spectra to unity at 5500Å. This assumes that both stars are reddened by the same amount (see Discussion). We then compared the set of de-reddened observed ($`a`$+$`e`$) spectra to the set of template spectra. The template spectra including an M0 component were rejected since they exhibited residual TiO bands which do not appear in the observed ($`a`$+$`e`$) spectrum and thus are not shown here. The absorption features of the observed ($`a`$+$`e`$) spectrum, in particular the G-band of CH ($`\lambda \lambda `$4290-4314Å), are only compatible with a late G or early K type for star $`a`$. Fig. 3 shows four different template spectra (K3+K7, K0+K7, G8+K7 and G5+K7) (grey curves) together with the average ($`a`$+$`e`$) spectrum de-reddened with $`E(BV)`$ = 0.3, 0.45, 0.50 and 0.60, values selected to give a best fit to the template spectra. The depth of the Mg$`b`$ band at $`\lambda `$5175Å (which was the only late-type feature in the first quiescent spectrum published by Thorstensen et al. Thor (1978)) and of the TiO band at $`\lambda `$4954Å increase with later spectral types and are too strong in the K3+K7 combination. The ($`a`$+$`e`$) de-reddened energy distributions for $`E(BV)`$ = 0.60, 0.50 and 0.45 match almost equally well the G5+K7, G8+K7 or the K0+K7 composite templates, respectively, although there are differences at the blue and red ends. In Fig. 4 we show the result of subtracting the synthetic G8 spectrum, assigned to star $`a`$, from the ($`a`$+$`e`$) spectrum, de-reddened for $`E(BV)`$ = 0.50. This should approximate the spectrum of star $`e`$. Also shown (grey line) is our observed K7 V spectrum, normalized to 0.12 at 5500 Å. The main features of this difference spectrum are the emission lines at H$`\alpha `$, H$`\beta `$ (weaker) and Ca II K and H(+H$`ϵ`$), which appear strong although the signal-to-noise ratio below 4500Å in our spectra is low. Emission at H$`\gamma `$ is barely detectable and no emission is detected at He ii $`\lambda `$4686Å. The slight deficit between 4400 and 5000 Å may be of instrumental origin. Our spectrum is different from that taken by Garcia et al. (Gar (1999)) when the sum of Aql X-1 plus star $`a`$ was more than half a magnitude above quiescence ($`V`$ = 18.68), and which displays a richer emission-line spectrum, including He ii $`\lambda `$4686Å, a consequence of X-ray heating. ## 4 Discussion The absorption features present in our spectrum of the sum ($`a`$+$`e`$) indicate a spectral type for star $`a`$ between mid-G and K0, reddened by $`E(BV)`$ = 0.5 $`\pm `$0.1. Using $`E(VI)=1.3\times E(BV)`$ (Dean, Warren and Cousins Dean (1978)) and the same amount of reddening $`E(VI)=0.65\pm 0.1`$ for stars $`a`$, $`b`$, $`c`$, $`d`$ and $`e`$, we obtain the estimates, listed in the two right-hand columns of Table 1, for the de-reddened color index $`(VI)`$ and the corresponding ranges of spectral types using the colors computed by Bessell (Bess (1990)) for Vilnius spectra. The magnitudes listed in Table 1 correspond to the flux ratios $`f(e)/f(a)=0.135`$ or $`f(e)/f(e+a)=0.12`$ in the $`V`$-band and $`f(e)/f(a)=0.29`$ or $`f(e)/f(e+a)=0.22`$ in the $`I`$-band. According to Callanan et al. (Call (1999)), star $`a`$ contributes 60% of the combined flux from the pair ($`a`$+$`e`$) at $`K`$, which gives a flux ratio $`f(e)/f(a)=0.67`$ and a magnitude difference $`\mathrm{\Delta }K_{ea}=K(e)K(a)=0.45`$. In the $`V`$-band we find $`\mathrm{\Delta }V_{ea}=V(e)V(a)=2.18`$. Combining both results yields the difference in $`VK`$ color between the two stars, $`\mathrm{\Delta }(VK)_{ea}=1.73`$. Assuming a similar amount of reddening for stars $`e`$ and $`a`$, we derive the intrinsic $`(VK)`$ color and spectral type for star $`e`$ as a function of the intrinsic $`(VK)`$ colour and spectral type assumed for star $`a`$, using the main sequence calibration of Johnson (John (1966)), and these are shown in Table 2. These determinations are compatible with the $`V`$ and $`I`$ photometric results of Table 1 and with a color excess $`E(BV)`$ = 0.5 $`\pm `$0.1. Assuming absolute visual magnitudes of $`M_V=+5.8`$ and $`+8.1`$ for stars $`a`$ and $`e`$ (Gray Gray (1992)) and taking $`A_V=3.1\times E(BV)`$, the magnitudes of Table 1 yield distances of 2.6 and 2.5 kpc, respectively, in agreement with the assumption of equal amounts of reddening for both objects. For a galactic latitude of $`4.1\mathrm{°}`$, such distances put the objects 180 pc below the galactic plane, beyond most of the absorbing layer. The low-energy X-ray absorbing column densities in the line of sight to Aql X-1 reported by Verbunt et al. (Verb (1994)) and Zhang et al. (Zhang (1998)) correspond to color excesses $`E(BV)`$ of 0.59 $`\pm `$0.1 and 0.49 $`\pm `$0.1, respectively (using the relation of Ryter et al. Ryter (1975)). These values are compatible with those used here. ## 5 Conclusions We have shown that the true optical counterpart of the Aquila X-1 recurrent transient is the interloper reported by Callanan et al. (Call (1999)), located 0.48″ West of the star previously assumed to be the candidate, now an ordinary mid to late-G type star. Photometry and low-resolution spectroscopy obtained in quiescence give for the new counterpart $`V`$ = 21.6 and $`VI`$ = +2.2 and suggest a star of K7 V spectral type, reddened by $`E(BV)`$ = 0.5 $`\pm `$0.1 and located at 2.5 kpc. This makes Aql X-1 similar to most other soft X-ray transients which have late K (or M0) companions. We defer examination of the consequences of this new identification on existing optical photometry to a forthcoming article (Chevalier and Ilovaisky1999c ).
no-problem/9906/chao-dyn9906039.html
ar5iv
text
# Evolution of Multispecificity in an Immune Network ## 1 Introduction The ‘lock and key’ concept has been central to understanding the specificity of biochemical molecular interactions, from enzyme-substrate relationships to antigen-antibody matchings. However, it has gradually been realized that such a ‘lock and key’ concept is not strictly valid, particularly in immune systems . Antigen-antibody interactions are found to be plastic or ‘ multispecific’ rather than fixed or single-specific. Namely, antibodies inherently have a flexible recognition capacity. Kearney et al. have confirmed experimentally the existence of such ambiguity of recognition in the antibody binding site of immature B cells. It is generally believed that development from ambiguous to specific recognition is caused by somatic hypermutations . We here propose a new dynamics of specificity evolutions based on Jerne’s network hypothesis . Our model is characterized by a meta-dynamics of idiotype specificity on shape-space . We show here that specific and non-specific responses to an antigen are governed dynamically by a fixed point attractor and a chaotic long-lived transient state of an immune network, respectively. The relevance of such a long-lived transient state is discussed with respect to immune function. ## 2 Modeling with a Meta-dynamics of Specificity We first introduce the standard idiotypic network model. Each idiotype is characterized by a pair of surface sites, called the idiotope and the paratope. If the idiotope site of a lymph cell is bounded by paratopes of other lymph cells, the recognized lymph cells become inactivated, whereas the recognizing cells become activated. Thus the growth dynamics of clone size $`x_{k,j}^n`$ of an idiotype of paratope $`k`$ and idiotope $`j`$ is given as, $$x_{k,j}^{n+1}=x_{k,j}^n+\underset{p}{}\underset{q}{}b_{q,k}x_{p,q}^nx_{k,j}^n\alpha \underset{p}{}\underset{q}{}b_{j,p}x_{k,j}^nx_{p,q}^ndx_{k,j}^n+s,$$ (1) The idiotope-paratope interaction $`b_{i,j}`$ is assumed to have an exponential form: $`\frac{1}{\sigma }e^{\frac{|ij|}{\sigma }}`$. We characterize the ambiguity of the antigen-antibody by the deviation parameter $`\sigma `$. The proposed meta-dynamics controls this parameter. First, as a simple example, we quantize $`\sigma `$ by the power of 2: $`\sigma _m=2^{Mm}`$. The maximum specificity is given by $`m=M`$. Now each idiotype is characterized by three variables: idiotope $`k`$, paratope $`j`$, and the specificity $`m`$. We thus describe the evolution of specificity as follows: $$x_{k,j,m}^{n+1}=(1\mu ^{})x_{k,j,m}^n+\mu ^{}/2\underset{m^{}=m1,m+1}{}x_{k,j,m^{}}^n+s_{m=1},$$ (2) where $`\mu ^{}`$ is the mutation rate of specificity. Here the source term $`s_{m=1}`$ is added for the least specific antibody. This dependency reflects the fact that the premature B-cells are believed to have lower specificities. By combining these equations, we establish the complete clone growth dynamics with mutations among idiotypes and the evolution of the specificities. In our model, there are five different types of idiotopes and of paratopes, so that there are 25 different idiotypes, with $`M=5`$ different levels of specificity. The rest of the system parameters (i.e. $`\mu ^{}=0.3`$, $`s=1.0`$ , $`d=0.1`$, and $`\alpha =2.0`$) are selected so that the size of each clone never diverges. The following results (especially, the natural tolerance at high amount of antigen) are confirmed not to depend on the values of the mutation rate $`\mu ^{}`$ and the source $`s`$. The dependency of system size is still unclear. ## 3 Dynamical Natures of the Network We pay most attention to how the idiotype network responds to persistent antigenic stimulations. A static antigen with a binding site $`k`$ is introduced by adding the constant term $`+A_kb_{k,j}x_{i,j,m}`$ to the above equation. Estimating the mean network specificity $`Sp_k`$ by averaging the specificity of all idiotypes bearing paratope type k, we study the antigenic effect on the network dynamics. An antigen of type 4 is used as an example, but the following result does not depend on the selected antigen type. Because we adopt the periodic boundary condition for the shape-space, each idiotype is equivalent within a network. We show a plot of the averaged specificity ($`\overline{Sp_4}`$) and the maximum Lyapunov exponent under the antigen stimulations over $`10^4`$ steps (see Fig. 1(a), (b)). In Fig.1(a), as expected, the network specificity increases when we increase the amount of antigen. At about 9.5 units of antigen, however, the specificity abruptly diverges to a high value. We say that a specific response has occurred at this antigen level. This specific response is observed until the antigen level reaches 13.5 units. Beyond this critical value, the specific response is no longer observed. Inversely, the specificity is sustained at the lower values. This lower sustained response can be compared to natural tolerance to the antigen. On the other hand, by comparing Fig.1(a) with (b), when the amount of the antigen is set between 9.5 and 13.5 units, we notice that the lower specificity emerges with chaotic dynamics, and the higher specificity emerges with a fixed point dynamics. We shall call the former dynamics a type I attractor and the latter a type II attractor. However, the type I attractor is not a true attractor. It was found to be a long-lived transient state referred to as a super-transient state, which is a common phenomenon in high-dimensional dynamical systems . In Fig.1(c), when the observation period is extended, we observe a transition from type I to type II attractor. There is no inverse-transition from the type II to the type I attractor. The super-transient states are highly dependent on the antigen level. For example, when the antigen level is 11.5 units in Fig.1(c), the transition probability from type I to type II is still less than 12 percent. In such cases, it behaves as an attractor in a practical sense. From a practical viewpoint, response time is also worth noticing. If we say that the relevant time scale for the immune response should be less than 10,000 time steps, in a practical sense there is no specific response even at higher levels of antigen (see Fig.1(c)). Our results suggest that a certain level of antigen causes the super-transient state to suppress fast immune responses under the idiotype network. Besides the response time, much attention has been paid in the field of theoretical immunology to topological changes of the network . Here we argue that the transition from type I (unspecific) to type II (specific) causes a simultaneous change of network topology. The network topology of each of these two states is shown in Fig.2. As we see from the figure, a chaotic super-transient state of type I has a more complex network than does type II. Inversely, higher specificity to the dosed antigen is maintained by a simpler network structure. The maintenance of idiotypic diversity can be attributed to chaotic dynamics. By estimating the amount of specificity of all idiotypes in the type I’s distributed state and the type II’s localized state respectively, it is found that each idiotype in type I’s distributed state has a low specificity on the whole. Namely, each idiotype interacts weakly with many idiotypes in order to have high connectivity. As a result, the stimulation of the network by dosed antigen is distributed over the network, not concentrated only on idiotypes bearing a binding site (paratope with type 4) for the antigen. Thus, the immune response to the antigen has a tendency to be suppressed. This result would support Stewart’s extrapolation that ‘The higher connectivity among idiotypes, the greater the degree of tolerance’ . Recently, a chaotic oscillation was found experimentally in a natural tolerant state. Subsequently, theoretical immunologists have tried to establish ‘natural tolerance under chaotic dynamics’ against a static antigen , though their simulation results show difficulty establishing such a tolerance without assuming a special network topology of an ‘odd-loop structure’ and so-called ‘bell-shaped function’ as an activation function. We have shown how such a tolerance can arise naturally under a chaotic dynamics, without these assumptions, by adding an additional flexibility; i.e., meta-mutation dynamics with specificity of idiotype. We have used a simple idiotypic network model, and have not ventured to use the more complex ’ bell-shaped function model’ because of focusing on capabilities of the meta-dynamics we introduced. Applying the meta-dynamics we introduced here with the bell-shaped function model is left as a future problem. ## 4 Concluding Remarks In this paper, we have expanded the possibilities of theoretical immunology by introducing new meta-dynamics. We believe that the immune response should be seen as having a more dynamic nature than allowed by most current models , and that the specific antigen-response and the dynamical percolation related to natural tolerance are caused by the meta-dynamics controlling the degree of specificity, as introduced here.
no-problem/9906/astro-ph9906195.html
ar5iv
text
# Observational Mishaps: a Database ## 1 Introduction It is not uncommon for an astronomical image obtained after a lengthy integration to reveal that all is not well. As a consequence, telescope time is sacrificed identifying the problem. In an effort to shorten this investigation period, we have created a catalog of astronomical images bearing signatures of a range of mishaps encountered during observing runs. Included with each image is an explanation of the cause of the problem as well as a suggested solution. Since a large number of observatories today are connected to the Internet, the World Wide Web (WWW) was chosen as the ideal medium for presenting this collection of images. Initially, the purpose of such a collection was to assist new graduate student observers at Michigan-Dartmouth-MIT (MDM) Observatory who frequently observe without the benefit of a more experienced observer. The aim was to provide these students with a means of quickly pinpointing the underlying problem affecting the image quality. This idea grew into a WWW accessible database complete with explanations of the “mishaps” responsible for the deterioration of the images, as well as suggested solutions. ## 2 The Format of the Database Every WWW page in this catalog contains an inverted colormap GIF image of the mishap, a table listing relevant information about the image (telescope, date, instrument, filter, exposure time), a brief description of the problem, and, if available, a suggestion of how to fix it. In a few cases, the cause of the problem could not be determined. These were dubbed “Unsolved Mysteries”, and no explanation of the problem or suggestion for a fix are given. Since it is possible for one problem to manifest itself in a variety of ways, multiple images of the same mishap are presented where appropriate, cross referenced with the help of hypertext links. For example, condensation on the dewar window can appear as a filamentary structure or as a bright extended feature with cusps, depending on the locations of light sources in the field of view. For the more common problems of astigmatism, coma, bad guiding/focusing, and poor seeing, we have provided supporting plots/images where applicable via links on the relevant pages. Examples include radial profile plots across a stellar image or multiple images of the same field taken in different seeing conditions. In Fig. 1, we show an example of a typical page in the database, along with the explanation of the problem and a suggestion for the solution. ## 3 The Structure of the Database Much consideration was given to effectively structuring the image catalog. Rather than sorting the images by cause, which is probably unknown to the astronomer accessing the database, we have grouped them by symptom. We provide the following two options for searching the database: 1. The user may browse the complete list of compiled images. This list features links to the various mishap pages as well as a brief description (1 - 2 lines) of the symptoms in the corresponding image. 2. The other option is to first broadly classify the image based on its symptoms and then choose the appropriate web page from a smaller list. This option will likely be more practical with an increasing number of images in the database. Apart from the frequently occuring problems of bad seeing/focusing/guiding, fringing, dust rings, and reflections, the current revision of the database lists the following as the top categories: * Unusual Appearance of Objects in the Image: familiar objects in the image, such as galaxies, stars, etc, have an unexpected appearance (e.g., guider jumps, deflated airbags, etc). * CCD and Electronics Features: features seem to be correlated with the CCD rows or columns, or they are otherwise suspiciously electronic in appearance (e.g., readout errors, shutter failure, etc). * Unexpected Objects in the Image and other External Interference: unexpected features obviously not due to the CCD or the electronics appear in the image (e.g., occulting dropout shutter, condensation on the dewar window, etc). * Unsolved Mysteries: as mentioned above, these are the cases for which we have so far not been able to determine the cause of the problem. Each of the above links leads to a list of mishap pages in that category with a brief description of the corresponding image appearance. ## 4 The Location of the Database The Observational Mishaps Database can be accessed at http:$`//`$www.astro.lsa.umich.edu$`/`$mishaps$`/`$mishaps.html. It is also directly accessible from the University of Michigan Astronomy Department Home Page, whose URL is http:$`//`$www.astro.lsa.umich.edu. ## 5 Additional Remarks We have created a database of images which are deteriorated by the effects of various mishaps encountered during astronomical observing runs. Its structure was designed to help users quickly identify the cause of the poor image quality, thus saving telescope time. In addition to being widely accessible via the WWW, the advantage of such an on-line catalog is its versatility. Unlike a printed catalog, the on-line version can very easily be updated, corrected, and expanded, so that everytime the database is accessed the user will find it in its most up-to-date form. Due to the practically infinite number of possible problems during observing runs, this collection is clearly far from and impossible to complete. Its usefulness, however, is obviously directly related to the number of examples it contains, and therefore we would appreciate any contributions by the astronomical community in the form of examples which might fit into this collection. Instructions for the submission of such images are given in the database. Furthermore, we realize that some of our interpretations of the mishaps, as well as some of our suggestions on how to improve the images, may be incorrect or incomplete. While it is our intention to regularly update and improve this database, we welcome any input about the database in general, its structure, or even individual examples. We would like to express our gratitude to the following people who contributed to this project by supplying examples and/or providing explanations of some of the mishaps: Gary Bernstein, Mario Mateo, Eric Miller, Patricia Knezek, Kelly Holley-Bockelmann, Lynne Allen, Michel Festou, and Doug Welch.
no-problem/9906/hep-ph9906455.html
ar5iv
text
# References CPT–99/P.3845. Improved positivity bound for Deep Inelastic Scattering on transversely polarized nucleon J. Soffer <sup>1</sup><sup>1</sup>1E-mail: soffer@cpt.univ-mrs.fr Centre de Physique Théorique - CNRS - Luminy, Case 907 F-13288 Marseille Cedex 9 - France and O. V. Teryaev<sup>2</sup><sup>2</sup>2 E-mail: teryaev@thsun1.jinr.ru Bogoliubov Laboratory of Theoretical Physics, Joint Institute for Nuclear Research, Dubna, 141980, Russia ## Abstract The positivity bound for the transverse asymmetry $`A_2`$ may be improved by making use of the fact, that the state of a photon and a nucleon with total spin 3/2, does not participate to the interference. The bound is therefore useful in the case of a longitudinal asymmetry small (say, at low $`x`$) or negative (like in the neuteron case). Positivity is playing a very important role in constraining various spin- dependent observables, in particular by providing a bound for the transverse asymmetry in polarized Deep Inelastic Scattering (DIS). It is a well-known condition established long time ago and based on an extensive study by Doncel and de Rafael , written in the form $`|A_2|\sqrt{R},`$ (1) where $`A_2`$ is the usual transverse asymmetry and $`R=\sigma _L/\sigma _T`$ is the standard ratio in DIS of the cross section of longitudinally to transversely polarized off-shell photons. It reflects a non-trivial positivity condition one has on the photon-nucleon helicity amplitudes. By substituting photons for gluons, we found earlier, that the similar bound holds for the various matrix elements for longitudinal gluons in a nucleon $`|\mathrm{\Delta }G_T(x)|\sqrt{1/2G(x)G_L(x)}.`$ (2) However, this bound can be rederived in line with the positivity bound in the quark case, known as Soffer inequality , $`|h_1(x)|q_+(x)={\displaystyle \frac{1}{2}}[q(x)+\mathrm{\Delta }q(x)],`$ (3) by making the substitution in Eq.(2), $`G(x)G_+(x)=\frac{1}{2}[G(x)+\mathrm{\Delta }G(x)]`$, and providing a stronger restriction, especially when the gluon helicity distribution $`\mathrm{\Delta }G(x)`$ is small or even negative. Coming back to the photon case, if $`A_1`$ denotes the asymmetry with longitudinally polarized nucleon, we are led to $`|A_2|\sqrt{R(1+A_1)/2},`$ (4) a stronger bound than Eq.(1). In the present paper we will show that this is really the case, using a transparent physical approach, and we will comment on, why, we think the weaker bound was used up to now. We start with the following expressions for the various photon-nucleon cross-sections in terms of the matrix elements describing the transition from the state $`|H,h>`$ of a nucleon with helicity $`h`$ and a photon with helicity $`H`$, to the unobserved state $`|X>`$ $`\sigma _T^\pm ={\displaystyle \underset{X}{}}|<+1/2,+1|X>|^2\pm |<+1/2,1|X>|^2,`$ $`\sigma _L={\displaystyle \underset{X}{}}|<+1/2,0|X>|^2={\displaystyle \underset{X}{}}|<1/2,0|X>|^2,`$ $`\sigma _{LT}=2Re{\displaystyle \underset{X}{}}<+1/2,+1|X><1/2,0|X>.`$ (5) Note that while longitudinal and transverse cross-sections are symmetric with respect to the reverse of the nucleon and photon helicities, this is not the case for the interference term. The reason is very simple: the opposite helicities of photon and nucleon correspond to their spins parallel , so that the angular momentum of the state $`|X>`$ has its maximum value $`3/2`$. The amplitude, which could possibly interfere with it to produce the transverse asymmetry, should have the same total angular momentum of the state $`|X>`$. This is however impossible, as the flip of the one of the helicities would require another one to exceed its maximal possible value, in order to keep the angular momentum of $`|X>`$ the same. Therefore the interference, responsible for $`A_2`$, does not occur. This is quite a general reason, for the occurence of the $`+`$ helicity configurations in all the cases considered above. We are now ready to write down the Cauchy-Schwarz inequality as $`{\displaystyle \underset{X}{}}|<+1/2,+1|X>\pm a<1/2,0|X>|^20,`$ (6) where $`a`$ is a positive real number. By making use of the definitions (S0.Ex1) and after the standard minimization with respect to the choice of $`a`$, one immediately arrives at $`|\sigma _{LT}|\sqrt{\sigma _L\sigma _T^+},`$ (7) leading directly to (4). The use of the new bound is resolving partially the puzzle, why the measured $`A_2`$ is such a small quantity. The fact, that the bound (1) is far from being saturated is obvious at low $`x`$ in the proton case, because, according to (4), it should be decreased by a factor $`\sqrt{2}`$ due to the small longitudinal asymmetry. The bound under consideration is even more useful with a negative longitudinal asymmetry, like in the neutron case. Recall that for a pure 3/2 configuration we have $`A_1=1`$ which implies $`A_2=0`$ One should note finally, that this result is actually coming from the original papers , while it was somehow weakened and transformed to a more suitable form Eq.(1), because one was willing to exclude $`A_1`$ which was poorly known twenty years ago. To be convince of that, one should look at Eq.(2.40a) in , which was, in fact, already contained in . To conclude, we rederived a known, but so far forgotten stronger bound for the transverse asymmetry in polarized DIS. We are indebted to E. de Rafael for useful dicsussions and to Z-E. Meziani and R. Windmolders for interest in the work.
no-problem/9906/hep-th9906087.html
ar5iv
text
# 1 Introduction ## 1 Introduction There has been a great deal of excitement during this past year following the realization that certain field theories admit concrete realizations as a string theory on a particular background . By now many examples of this type of correspondence for field theories in various dimensions with various field contents have been reported in the literature (for a comprehensive review and list of references, see ). However, attempts to apply these correspondences to study the details of these theories have only met with limited success so far. The problem stems from the fact that our understanding of both sides of the correspondence is limited. On the field theory side, most of what we know comes from perturbation theory where we assume that the coupling is weak. On the string theory side, most of what we know comes from the supergravity approximation where the curvature is small. There are no known situations where both approximations are simultaneously valid. At the present time, comparisons between the dual gauge/string theories have been restricted to either qualitative issues or quantities constrained by symmetry. Any improvement in our understanding of field theories beyond perturbation theory or string theories beyond the supergravity approximation is therefore a welcome development. In this note we raise the Supersymmetric Discrete Light Cone Quantization (SDLCQ) of field theories to the challenge of providing quantitative data which can be compared against the supergravity approximation on the string theory side of the correspondence. We will work in two space-time dimensions where the SDLCQ approach provides a natural non-perturbative solution to the theory. In general, attempts to improve the field theory side beyond perturbation theory seem like a promising approach in two space-time dimensions where a great deal is already known about field theories beyond perturbation theory. We will study the field theory/string theory correspondence motivated by considering the near-horizon decoupling limit of a D1-brane in type IIB string theory . The gauge theory corresponding to this theory is the Yang-Mills theory in two dimensions with 16 supercharges. Its SDLCQ formulation was recently reported in . This is probably the simplest known example of a field theory/string theory correspondence involving a field theory in two dimensions with a concrete Lagrangian formulation. A convenient quantity that can be computed on both sides of the correspondence is the correlation function of gauge invariant operators . We will focus on two point functions of the stress-energy tensor. This turns out to be a very convenient quantity to compute for many reasons that we will explain along the way. Some aspects of this as it pertains to a consideration of black hole entropy was recently discussed in . There are other physical quantities often reported in the literature. In the DLCQ literature, the spectrum of hadrons is often reported. This would be fine for theories in a confining phase. However, we expect the SYM in two dimension to flow to a non-trivial conformal fixed point in the infra-red . The spectrum of states will therefore form a continuum and will be cumbersome to handle. On the string theory side, entropy density and the quark anti-quark potential are frequently reported. The definition of entropy density requires that we place the field theory in a space-like box which seems incommensurate with the discretized light cone. Similarly, a static quark anti-quark configuration does not fit very well inside a discretized light-cone geometry. The correlation function of point-like operators do not suffer from these problems. We should mention that there exists interesting work on computing the QCD string tension directly in the field theory. These authors find that the QCD string tension vanishes in the supersymmetric theories which is consistent with the power law quark anti-quark potential found on the supergravity side. ## 2 Correlation functions from supergravity Let us begin by reviewing the computation of the correlation function of stress energy tensors on the string theory side using the supergravity approximation. The computation is essentially a generalization of . The main conclusion on the supergravity side was reported recently in but we will elaborate further on the details. The near horizon geometry of a D1-brane in string frame takes the form $`ds^2`$ $`=`$ $`\alpha ^{}\left({\displaystyle \frac{U^3}{\sqrt{64\pi ^3g_{YM}^2N}}}dx_{}^2+{\displaystyle \frac{\sqrt{64\pi ^3g_{YM}^2N}}{U^3}}dU^2+\sqrt{64\pi ^3g_{YM}^2N}Ud\mathrm{\Omega }_{8p}^2\right)`$ $`e^\varphi `$ $`=`$ $`2\pi g_{YM}^2\left({\displaystyle \frac{64\pi ^3g_{YM}^2N}{U^6}}\right)^{\frac{1}{2}}.`$ (1) In order to compute the two point function, we need to know the action for the diagonal fluctuations around this background to the quadratic order. What we need is an analogue of for this background which unfortunately is not currently available in the literature. Fortunately, some diagonal fluctuating degrees of freedom can be identified by following the early work on black hole absorption cross-sections . In particular, we can show that the fluctuations parameterized according to $`ds^2`$ $`=`$ $`\left(1+f(x^0,U)+g(x^0,U)\right)g_{00}(dx^0)^2+\left(1+5f(x^0,U)+g(x^0,U)\right)g_{11}(dx^1)^2`$ $`+\left(1+f(x^0,U)+g(x^0,U)\right)g_{UU}dU^2+\left(1+f(x^0,U){\displaystyle \frac{5}{7}}g(x^0,U)\right)g_{\mathrm{\Omega }\mathrm{\Omega }}d\mathrm{\Omega }_7^2`$ $`e^\varphi `$ $`=`$ $`\left(1+3f(x^0,U)g(x^0,U)\right)e^{\varphi _0}`$ (2) will satisfy the equations of motion $`f^{\prime \prime }(U)+{\displaystyle \frac{7}{U}}f^{}(U){\displaystyle \frac{64\pi ^3g_{YM}^2Nk^2}{U^6}}f(U)`$ $`=`$ $`0`$ $`g^{\prime \prime }(U)+{\displaystyle \frac{7}{U}}g^{}(U){\displaystyle \frac{72}{U^2}}g(U){\displaystyle \frac{64\pi ^3g_{YM}^2Nk^2}{U^6}}g(U)`$ $`=`$ $`0`$ (3) by direct substitution into the equations of motion in 10 dimensions. We have assumed without loss of generality that these fluctuation vary only along the $`x^0`$ direction of the world volume coordinates like a plane wave $`e^{ikx^0}`$. The fields $`f(U)`$ and $`g(U)`$ are scalars when the D1-brane is viewed as a black hole in 9 dimensions; in fact there are the minimal and the fixed scalars in this black hole geometry. In 10 dimensions, however, we see that they are really part of the gravitational fluctuation. We expect therefore that they are associated with the stress-energy tensor in the operator field correspondence of . In the case of the correspondence between $`𝒩=4`$ SYM and $`AdS_5\times S_5`$, superconformal invariance allowed the identification of operators and fields in short multiplets . For the D1-brane, we do not have superconformal invariance and this technique is not applicable. In fact, we expect all fields of the theory consistent with the symmetry of a given operator to mix. The large distance behavior should then be dominated by the contribution with the longest range. The field $`f(k^0,U)`$ appears to be the one with the longest range since it is the lightest field. The equation (3) for $`f(U)`$ can be solved explicitly in terms of the Bessel’s function $$f(U)=U^3K_{3/2}(\sqrt{16\pi ^3g_{YM}^2N}U^2k).$$ (4) By thinking of $`f(U)`$ in direct analogy with the minimally coupled scalar as was done in , we can compute the flux factor $$=\underset{U_0\mathrm{}}{lim}\frac{1}{2\kappa _{10}^2}\sqrt{g}g^{UU}e^{2(\varphi \varphi _{\mathrm{}})}_U\mathrm{log}(f(U))|_{U=U_0}=\frac{NU_0^2k^2}{2g_{YM}^2}\frac{N^{3/2}k^3}{4g_{YM}}+\mathrm{}$$ (5) up to a numerical coefficient of order one which we have suppressed. We see that the leading non-analytic (in $`k^2`$) contribution is due to the $`k^3`$ term, whose Fourier transform scales according to<sup>1</sup><sup>1</sup>1It is not difficult to show that for a generic $`p`$-brane, $`𝒪(x)𝒪(0)=N^{\frac{7p}{5p}}g_{YM}^{\frac{2(3p)}{5p}}x^{\frac{19+2pp^2}{5p}}.`$ $$𝒪(x)𝒪(0)=\frac{N^{\frac{3}{2}}}{g_{YM}x^5}.$$ (6) This result passes the following important consistency test. The SYM in 2 dimensions with 16 supercharges have conformal fixed points in both UV and IR with central charges of order $`N^2`$ and $`N`$, respectively. Therefore, we expect the two point function of stress energy tensors to scale like $`N^2/x^4`$ and $`N/x^4`$ in the deep UV and IR, respectively. According to the analysis of , we expect to deviate from these conformal behavior and cross over to a regime where supergravity calculation can be trusted. The cross over occurs at $`x=1/g_{YM}\sqrt{N}`$ and $`x=\sqrt{N}/g_{YM}`$. At these points, the $`N`$ scaling of (6) and the conformal result match in the sense of the correspondence principle . ## 3 Correlation functions from DLCQ The challenge then is to attempt to reproduce the scaling relation (6), fix the numerical coefficient, and determine the detail of the cross-over behavior using SDLCQ. Ever since the original proposal , the question of equivalence between quantizing on a light-cone and on a space-like slice have been discussed extensively. This question is especially critical whenever a massless particle or a zero-mode in the quantization is present. It is generally believed that the massless theories can be described on the light-cone as long as we take $`m0`$ as a limit. The issue of zero mode have been examined by many authors. Some recent accounts can be found in . Generally speaking, supersymmetry seems to save SDLCQ from complicated zero-mode issues. We will not contribute much to these discussions. Instead, we will formulate the computation of the correlation function of stress energy tensor in naive DLCQ. To check that these results are sensible, we will first do the computation for the free fermions and the ’t Hooft model. Extension to SYM with 16 supercharges will be essentially straightforward, except for one caveat. In order to actually evaluate the correlation functions, we must resort to numerical analysis at the last stage of the computation. For the SYM with 16 supercharges, this problem grows too big too fast to be practical on desk top computer where the current calculations were performed. We can only provide an algorithm, which, when executed on an much more powerful computer, should reproduce (6). Nonetheless, the fact that we can define a concrete algorithm seems to be a progress in the right direction. One potential pit-fall is the fact that the computation may not show any sign of convergence. If this is the case, or if it converges to a result at odds with (6), we must go back and re-examine the issue of equivalence of forms and the issue of zero modes. The technique of DLCQ is reviewed by many authors so we will be brief here. The basic idea of light-cone quantization is to parameterize the space using light cone coordinates $`x^+`$ and $`x^{}`$ and to quantize the theory making $`x^+`$ play the role of time. In the discrete light cone approach, we require the momentum $`p_{}=p^+`$ along the $`x^{}`$ direction to take on discrete values in units of $`p^+/k`$ where $`p^+`$ is the conserved total momentum of the system and $`k`$ is an integer commonly referred to as the harmonic resolution. One can think of this discretization as a consequence of compactifying the $`x^{}`$ coordinate on a circle with a period $`2L=2\pi k/p^+`$. The advantage of discretizing the light cone is the fact that the dimension of the Hilbert space becomes finite. Therefore, the Hamiltonian is a finite dimensional matrix and its dynamics can be solved explicitly. In SDLCQ one makes the DLCQ approximation to the supercharges and these discrete representations satisfy the supersymmetry algebra. Therefore SDLCQ enjoys the improved renormalization properties of supersymmetric theories. Of course, to recover the continuum result, we must send $`k`$ to infinity and as luck would have it, we find that SDLCQ usually converges faster than the naive DLCQ. Of course, in the process, the size of the matrices will grow, making the computation harder and harder. Let us now return to the problem at hand. We would like to compute a general expression of the form $$F(x^{},x^+)=𝒪(x^{},x^+)𝒪(0,0).$$ (7) In DLCQ, where we fix the total momentum in the $`x^{}`$ direction, it is more natural to compute its Fourier transform $$\stackrel{~}{F}(P_{},x^+)=\frac{1}{2L}𝒪(P_{},x^+)𝒪(P_{},0).$$ (8) This can naturally be expressed in a spectrally decomposed form $$\stackrel{~}{F}(P_{},x^+)=\underset{i}{}\frac{1}{2L}0|𝒪(P_{})|ie^{iP_+^ix^+}i|𝒪(P_{},0)|0.$$ (9) ### 3.1 Free Dirac Fermions Let us first consider evaluating this expression for the stress-energy tensor in the theory of free Dirac fermions as a simple example. The Lagrangian for this theory is $$=i\overline{\mathrm{\Psi }}/\mathrm{\Psi }m\overline{\mathrm{\Psi }}\mathrm{\Psi }$$ (10) where for concreteness, we take $`\gamma ^0=\sigma ^2,\gamma ^1=i\sigma ^1`$ and we take $`\mathrm{\Psi }=2^{1/4}(\genfrac{}{}{0pt}{}{\psi }{\chi })`$. In terms of the spinor components, the Lagrangian takes the form $$=i\psi ^{}_+\psi +i\chi ^{}_{}\chi \frac{im}{\sqrt{2}}(\chi ^{}\psi \psi ^{}\chi ).$$ (11) Since we treat $`x^+`$ as time and since $`\chi `$ does not have any derivatives with respect to $`x^+`$ in the Lagrangian, it can be eliminated from the equation of motion, leaving a Lagrangian which depends only on $`\psi `$: $$=i\psi ^{}_+\psi +i\frac{m^2}{2}\psi ^{}\frac{1}{_{}}\psi .$$ (12) We can therefore express the canonical momentum and energy as $`P_{}`$ $`=`$ $`{\displaystyle 𝑑x^{}i\psi ^{}_{}\psi }`$ $`P_+`$ $`=`$ $`{\displaystyle 𝑑x^{}}{\displaystyle \frac{im^2}{2}}\psi ^{}{\displaystyle \frac{1}{_{}}}\psi .`$ (13) In DLCQ, we compactify $`x^{}`$ to have period $`2L`$. We can then expand $`\psi `$ and $`\psi ^{}`$ in modes $`\psi ={\displaystyle \frac{1}{\sqrt{2L}}}\left(b(n)e^{\frac{in\pi }{L}x^{}}+d(n)e^{\frac{in\pi }{L}x^{}}\right)`$ $`\psi ^{}={\displaystyle \frac{1}{\sqrt{2L}}}\left(b(n)e^{\frac{in\pi }{L}x^{}}+d(n)e^{\frac{in\pi }{L}x^{}}\right).`$ (14) Operators $`b(n)`$ and $`d(n)`$ with positive and negative $`n`$ are interpreted as a destruction and creation operators, respectively. In a theory with only fermions, it is customary to take anti-periodic boundary condition in order to avoid zero-mode issues. Therefore, $`n`$ will take on half-integer values<sup>2</sup><sup>2</sup>2In SDLCQ one must use periodic boundary condition for all the fields to preserve the supersymmetry.. They satisfy the anticommutation relation $$\{b(n),b(m)\}=\{d(n),d(m)\}=\delta _{n,m}.$$ (15) Now we are ready to evaluate (9) in DLCQ. As a simple and convenient choice, we take $$𝒪(k)=\frac{1}{2}𝑑x^{}\left(i\psi ^{}_{}\psi i(_{}\psi ^{})\psi \right)e^{\frac{ik\pi }{L}x^{}}.$$ (16) which is the Fourier transform of the local expression for $`P_{}`$ with the total derivative contribution adjusted to make this operator Hermitian. Therefore, this should be thought of as the $`T^{++}`$ component of the stress energy tensor. For reasons that will become clear as we go on, this turns out to be one of the simplest things to compute. When acted on the vacuum, this operator creates a state $$T^{++}(k)|0=\frac{\pi }{L}\left(\frac{k}{2}n\right)b(k+n)d(n)|0.$$ (17) Since the fermions in this theory are free, the plane wave states $$|n=b(k+n)d(n)|0$$ (18) constitute an eigenstate. The spectrum can easily be determined by commuting these operators: $$M_n^2|n=2P_{}P_+|n=m^2\left(\frac{k}{n}+\frac{k}{kn}\right)|n$$ (19) which is simply the discretized version of the spectrum of a two body continuum. All that we have to do now is calculate eigenstates of the actual theory we are interested in and to assemble these pieces into (9), but we can do a little more to make the result more presentable. The point is that since (9) is expressed in mixed momentum/position space notation in Minkowski space, the answer is inherently a complex quantity that is cumbersome to display. For the computation of two point function, however, we can go to position space by Fourier transforming with respect to the $`L`$ variable. After Fourier transforming, it is straight forward to Euclideanize and display the two point function as a purely real function without loosing any information. To see how this works, let us write (9) in the form $$\stackrel{~}{F}(P_{},x^+)=\left|\frac{L}{\pi }0|T^{++}(k)|n\right|^2\frac{1}{2L}\frac{\pi ^2}{L^2}e^{\frac{iM_n^2}{2(\frac{k\pi }{L})x^+}}.$$ (20) The quantity inside the absolute value sign is designed to be independent of $`L`$. Now, to recover the position space form of the correlation function, we inverse Fourier transform with respect to $`P_{}=k\pi /L`$. $$F(x^{},x^+)=\left|\frac{L}{\pi }0|T^{++}(k)|n\right|^2\frac{d\left(\frac{k\pi }{L}\right)}{2\pi }\frac{1}{2L}\frac{\pi ^2}{L^2}e^{i\frac{M_n^2}{2(\frac{k\pi }{L})}x^+i\frac{k\pi }{L}x^{}}.$$ (21) The integral over $`L`$ can be done explicitly and gives $$F(x^{},x^+)=\left|\frac{L}{\pi }0|T^{++}(k)|n\right|^2\left(\frac{x^+}{x^{}}\right)^2\frac{M_n^4}{8\pi ^2k^3}K_4\left(M_n\sqrt{2x^+x^{}}\right)$$ (22) where $`K_4(x)`$ is the 4-th modified Bessel’s function. We can now continue to Euclidean space by taking $`r^2=2x^+x^{}`$ to be real and considering the quantity $$\left(\frac{x^{}}{x^+}\right)^2F(x^{},x^+)=\left|\frac{L}{\pi }0|T^{++}(k)|n\right|^2\frac{M_n^4}{8\pi ^2k^3}K_4(M_nr).$$ (23) This is a fundamental result which we will refer to a number of times in this paper. It has explicit dependence on the harmonic resolution parameter $`k`$, but all dependence on unphysical quantities such as the size of the circle in the $`x^{}`$ direction and the momentum along that direction have been canceled. For the free fermion model, (23) evaluates to $$\left(\frac{x^{}}{x^+}\right)^2F(x^{},x^+)=\frac{N}{k}\underset{n}{}\frac{M_n^4}{32\pi ^2}\frac{(k2n)^2}{k^2}K_4(M_nr)$$ (24) with $`M_n^2`$ given by (19). The large $`k`$ limit can be gotten by replacing $`nkx`$ and $`\frac{1}{k}_n_0^1𝑑x`$. We recover the identical result using Feynman rules. For $`rm^1`$, this behaves like $$\left(\frac{x^{}}{x^+}\right)^2F(x^{},x^+)=\frac{N}{k}\underset{n}{}\frac{3(k2n)^2}{2\pi ^2k^2r^4}\frac{N}{2\pi ^2r^4}.$$ (25) ### 3.2 ’t Hooft Model Let us now turn to a slightly more interesting problem of computing the correlation function of $`T^{++}`$ in ’t Hooft’s model of two dimensional QCD in the large $`N`$ limit. This theory has two characteristic scales, one determined by the mass of the quarks and the other by the strength of the gauge coupling $`g_{YM}^2N`$. To a large extent, this is a solvable model. The spectrum and the wave function of the hadrons are encoded in a one parameter integral equation that can be handled in many ways. A thorough analysis of this model including the discussion of asymptotic behavior of certain correlation functions can be found in . This is clearly a very mature subject. Applying DLCQ to the ’t Hooft model is tantamount to placing ’t Hooft’s integral equation for the meson spectrum on a lattice. The lattice is in the light-cone momentum space, which is precisely what is expected when the light-cone is compactified on a circle. The DLCQ of the ’t Hooft model was analyzed in detail by . Our goal here is to show that the computation of (9) is straight forward and that it generates sensible answers. In fact, nothing could be simpler. The ’t Hooft model is nothing more than a gauged version of the free fermion model. The Lagrangian for this theory is simply $$=\frac{1}{4g_{YM}^2}F^2+i\overline{\mathrm{\Psi }}D/\mathrm{\Psi }m\overline{\mathrm{\Psi }}\mathrm{\Psi }.$$ (26) We choose the light cone gauge $`A^+=0`$ which is customary. One then finds that the $`A^{}`$ component of the gauge field is non-dynamical and can be eliminated using the equation of motion, just like the $`\chi `$ component of the spinor in the free fermion model. Expressing everything in terms of the only dynamical field in this theory which is $`\psi `$, one finds the canonical energy and momentum operators to take the form $`P_{}`$ $`=`$ $`{\displaystyle 𝑑x^{}i\psi ^{}_{}\psi }`$ $`P_+`$ $`=`$ $`{\displaystyle 𝑑x^{}\left(\frac{im^2}{2}\psi ^{}\frac{1}{_{}}\psi \frac{g_{YM}^2}{2}J_{}\frac{1}{_{}^2}J_{}\right)}`$ (27) where $`J_{}=\psi \psi ^{}`$. All that changed in comparison to the free fermion model is the addition of a current exchange term in the light-cone Hamiltonian. Therefore, all we have to do here is to perform the identical computation specified by (23), but using the modified Hamiltonian, and letting $`M_n^2`$ and $`|n`$ be the spectrum and the wavefunction of the $`n`$-th meson state in the spectrum. Since in DLCQ we are always working with a finite dimensional representation of the Hamiltonian dynamics, a small change in the form of the Hamiltonian matrix causes no particular difficulty. Let us discuss the result of such a computation. We will consider the case when $`g_{YM}^2Nm^2`$ so that the effect of the gauge interaction is strong. The spectrum can be computed reliably for large $`k`$ and is in agreement with the results reported in (see figure 1.a). Since the state (17) created by operator (16) is odd under parity $`nkn`$, only parity odd states contribute in the spectral decomposition. For $`r1/g_{YM}\sqrt{N}`$, we expect the correlation function to behave just like the free fermion. The lightest meson in the parity odd sector has a mass of order $`g_{YM}\sqrt{N}`$. Due to the presence of this mass-gap, for $`r1/g_{YM}\sqrt{N}`$, we expect to see an exponential damping of the correlation function. This is precisely the behavior we seem to be finding. In figure 1.b, we illustrate the result of computing (23) by first constructing the mass matrix $`M^2`$ symbolically, then evaluating the spectrum and the eigenfunctions numerically and assembling the pieces. We have chosen to set $`g_{YM}\sqrt{N/\pi }=10^1`$ and $`m=10^1`$. We have tried harmonic resolutions $`k=5`$, $`k=10`$, $`k=50`$, and $`k=100`$. Remarkably, the computation at a low harmonic resolution $`k`$ seems not to be so far off from the result found using larger values of $`k`$. The correlation function appears to have more or less converged by the time we reach $`k=100`$. We have also included a plot for $`g_{YM}^2N=0`$ with the same mass for comparison. One of the reasons why the convergence is relatively rapid is the fact that the matrix element $`0|T^{++}(k)|n`$ in (23) is not sensitive to the structure of the ’t Hooft wave function at the boundaries. To see this more clearly, recall that in the continuum limit, we expect the eigenfunctions $`|n`$ to behave as $`x^\beta `$ near $`x=0`$ where $`\beta `$ is determined by $$0=\frac{m^2}{g_{YM}^2N/\pi }+\pi \beta \mathrm{cot}(\pi \beta )1\frac{m^2}{g_{YM}^2N/\pi }\frac{\pi ^2\beta ^2}{3}$$ (28) for small $`m^2/g_{YM}^2N`$ . When we compute matrix elements in DLCQ approximation, we are effectively exchanging integral expression like $$_0^1𝑑xax^{a1}=1$$ (29) by a discretized sum $$\underset{k=1}{\overset{n}{}}\frac{ak^{a1}}{n^a}1+\frac{a\zeta (1a)}{n^a}+𝒪(n^1)$$ (30) whose leading correction for $`a>1`$ is dominated by the terms of order $`1/n`$. For $`a`$ less than one, however, the leading correction is controlled by terms of order $`(\frac{1}{n})^a`$. In computing the form factor for $`T^{++}(k)`$, we were fortunate to have only encountered an integral whose end point behavior went as $`x^{1+\beta }`$. Had we instead chosen to compute two point function of a scalar operators like $`\overline{\psi }\psi `$ or $`\overline{\psi }\gamma ^5\psi `$, we would have considered states $`{\displaystyle 𝑑x^{}\overline{\psi }\psi (x^{})e^{\frac{ik\pi }{L}x^{}}|0}`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle \frac{mL}{2\pi }}\left({\displaystyle \frac{1}{n}}{\displaystyle \frac{1}{kn}}\right)b(n)d(k+n)|0`$ $`{\displaystyle 𝑑x^{}\overline{\psi }\gamma ^5\psi (x^{})e^{\frac{ik\pi }{L}x^{}}|0}`$ $`=`$ $`{\displaystyle \underset{n}{}}{\displaystyle \frac{mL}{2\pi }}\left({\displaystyle \frac{1}{n}}+{\displaystyle \frac{1}{kn}}\right)b(n)d(k+n)|0`$ (31) which gives rise to a pole near $`x=0`$ and $`x=1`$ in the continuum limit. Therefore, the matrix element in the continuum limit behaves near $`x=0`$ as $`x^\beta `$. To control the error in this case, we must take $$\frac{\beta }{n^\beta }ϵ$$ (32) or equivalently $$n\left(\frac{\beta }{ϵ}\right)^{1/\beta }$$ (33) which grows exponentially with respect to $`\beta ^1`$. We would have had to work much harder if we had sent $`m/g_{YM}\sqrt{N}`$ to zero. Luckily, this is not the case for with the $`T^{++}`$ operator. ### 3.3 Supersymmetric Yang-Mills theory with 16 supercharges Finally, let us turn to the problem of computing the two point function of the $`T^{++}`$ operator for the SYM with 16 supercharges. Just as in the ’t Hooft model, adopting light-cone coordinates and choosing the light-cone gauge will eliminate the gauge boson and half of the fermion degrees of freedom. The most significant change comes from the fact that the fields in this theory are in the adjoint rather than the fundamental representations and the theory is supersymmetric. This does not cause any fundamental problem in the DLCQ formulation of these theories. Indeed, the SDLCQ formulation of this as well as many other related models with adjoint fields have been studied in the literature. The main difficulty comes from the fact that in supersymmetric theories low mass states such as $`\mathrm{tr}[b(n_1)b(n_2)b(k+n_1+n_2)]|0`$ with an arbitrary number of excited quanta, or “bits,” appear in the spectrum. This means that for a given harmonic resolution $`k`$, the dimension of the Hilbert space grows like $`\mathrm{exp}(\sqrt{k})`$, which is roughly the number of ways to partition $`k`$ into sums of integers. The fact that the size of the problem grows very fast is somewhat discouraging from a numerical perspective. Nevertheless, it is interesting to note that DLCQ provides a well defined algorithm for computing a physical quantity like the two point function of $`T^{++}`$ that can be compared with the prediction from supergravity. In the following, we will show that this can be computed for the SYM theory by a straight forward application of (23), just as we saw in the case of the ’t Hooft model. The authors of have shown that the momentum operator $`P^+`$ is given by $$P^+=𝑑x^{}\mathrm{tr}\left[(_{}X_I)^2+iu_\alpha _{}u_\alpha \right].$$ (34) The local Hermitian form of this operator is given by $$T^{++}(x)=\mathrm{tr}\left[(_{}X^I)^2+\frac{1}{2}\left(iu^\alpha _{}u^\alpha i(_{}u^\alpha )u^\alpha \right)\right],I=1\mathrm{}8,\alpha =1\mathrm{}8$$ (35) where $`X`$ and $`u`$ are the physical adjoint scalars and fermions respectively, following the notation of . When discretized, these operators have the mode expansion $`X_{i,j}^I`$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{4\pi }}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}{\displaystyle \frac{1}{\sqrt{n}}}\left[A_{ij}^I(n)e^{i\pi nx^{}/L}+A_{ji}^I(n)e^{i\pi nx^{}/L}\right]`$ $`u_{i,j}^\alpha `$ $`=`$ $`{\displaystyle \frac{1}{\sqrt{4L}}}{\displaystyle \underset{n=1}{\overset{\mathrm{}}{}}}\left[B_{ij}^\alpha (n)e^{i\pi nx^{}/L}+B_{ji}^\alpha (n)e^{i\pi nx^{}/L}\right].`$ (36) In terms of these mode operators, we find $$T^{++}(k)|0=\frac{\pi }{2L}\underset{n=1}{\overset{k1}{}}\left[\sqrt{n(kn)}A_{ij}(k+n)A_{ji}(n)+\left(\frac{k}{2}n\right)B_{ij}(k+n)B_{ji}(n)\right]|0.$$ (37) Therefore, $`(L/\pi )0|T^{++}(k)|n`$ is independent of $`L`$ and can be substituted directly into (23) to give an explicit expression for the two point function. We see immediately that (37) has the correct small $`r`$ behavior, for in that limit, (37) asymptotes to (assuming $`n_b=n_f`$) $$\left(\frac{x^{}}{x^+}\right)^2F(x^{},x^+)=\frac{N^2}{k}\underset{n}{}\left(\frac{3(k2n)^2n_f}{4\pi ^2k^2r^4}+\frac{3n(kn)n_b}{\pi ^2k^2r^4}\right)=\frac{N^2(2n_b+n_f)}{4\pi ^2r^4}\left(1\frac{1}{k}\right)$$ (38) which is what we expect for the theory of $`n_b`$ free bosons and $`n_f`$ free fermions in the large $`k`$ limit. Computing this quantity beyond the small $`r`$ asymptotics, however, represents a formidable technical challenge. The authors of were able to construct the mass matrix explicitly and compute the spectrum for $`k=2`$, $`k=3`$, and $`k=4`$. Even for these modest values of the harmonic resolution, the dimension of the Hilbert space was as big as 256, 1632, and 29056 respectively (the symmetries of the theory can be used to reduce the size of the calculation somewhat). In figure 2, we display results that parallel those obtained for the ’t Hooft model earlier (figure 1) with the currently available values of $`k`$, except for the fact that we display the correlation function multiplied by a factor of $`4\pi ^2r^4/N^2(2n_b+n_f)`$, so that it now asymptotes to 1 (or 0 in the logarithmic scale) in the $`k\mathrm{}`$ limit. In this way any deviation from the asymptotic behavior $`1/r^4`$ is made more transparent. Note that with the values of the harmonic resolution $`k`$ obtained at present, the spectrum in figure 2.a is far from resembling a dense continuum near $`M=0`$. Clearly, we must probe much higher values of $`k`$ before we can sensibly compare our field theory results with the prediction from supergravity. ### 3.4 Supersymmetric Yang-Mills theory with less than 16 supercharges The computation of the correlator for the stress energy tensor in the (8,8) model is limited by our inability to carry out the computation for large enough harmonic resolution. It is the (8,8) model which we are ultimately interested in solving in order to compare against the prediction of Maldacena’s conjecture in the supergravity limit. Nevertheless, the computation of the correlation function can just as well be applied to models with less supersymmetry. We will conclude by reporting the results of such a computation. First, let us consider the theory with supercharges (1,1). This theory is argued not to exhibit dynamical supersymmetry breaking in . We can also provide a physicist’s proof that supersymmetry is not spontaneously broken for this theory by adopting the argument of Witten for the 2+1 dimensional SYM with Chern-Simons interaction . In , the index of 2+1 dimensional SYM with gauge group $`SU(N)`$ and 2 supercharges on $`R\times T^2`$ was computed and was found to be non-vanishing for Chern-Simons coupling $`k_3>N/2`$. If the period $`L`$ of one of the circles in $`T^2`$ is sufficiently small, this theory is approximately the 2-dimensional SYM with (1,1) supersymmetry with gauge coupling $`g_2^2=g_3^2/L`$ and BF coupling $`k_2=k_3L`$ . Imagine approaching this theory by taking $`L0`$ keeping $`g_2`$ and $`k_3`$ fixed. In this limit, $`k_20`$ in the units of $`g_2`$ so the limiting theory is that of pure SYM with (1,1) supersymmetry and a vanishing BF coupling. Choosing different values of $`k_3`$ corresponds to a different choice in the path of approach to this limit. If we chose $`k_3>N/2`$, we are guaranteed to have a non-zero index for finite $`L`$. This means that there will be a state with zero mass in the $`L0`$ limit also, indicating that supersymmetry is not spontaneously broken in this limit. On the other hand, the index is not a well defined quantity in the $`L0`$ limit, as a different choice of $`k_3`$ will lead to a different value of the index in the $`L0`$ limit. In fact, the index can be made arbitrarily large by taking $`k_3`$ to be also arbitrarily large. This suggests that there are infinitely many states forming a continuum near $`m=0`$. The index is therefore an ill defined quantity, akin to counting the number of exactly zero energy states on a periodic box as one takes the volume to infinity. This theory is also believed not to be confining and is therefore expected to exhibit non-trivial infra-red dynamics. The SDLCQ of the 1+1 dimensional model with (1,1) supersymmetry was solved in , and we apply these results directly in order to compute (23). For simplicity, we work to leading order in the large $`N`$ expansion. The spectrum of this theory for various values of $`k`$, and the subsequent computation of (23) is illustrated in figure 3.a. The spectrum of this theory at finite $`k`$, illustrated in figure 3.a, consists of $`2k2`$ exactly massless states<sup>3</sup><sup>3</sup>3 i.e. $`k1`$ massless bosons, and their superpartners., accompanied by large numbers of massive states separated by a gap. The gap appears to be closing in the limit of large $`k`$ however. We have tried extrapolating the mass of the lightest massive state as a function of $`1/k`$ by performing a least square fit to a line and a parabola, giving the extrapolated value of $`M^2\pi /g_{YM}^2N=1.7`$ and $`M^2\pi /g_{YM}^2N=0.6`$, suggesting indeed that at large $`k`$, the gap is closed. This is consistent with the expectation that the spectrum is that of a continuum starting at $`M=0`$ discussed earlier, although one must be careful when the order of large $`N`$ and large $`k`$ limits are exchanged. At finite $`N`$, we expect the degeneracy of $`2k2`$ exactly massless states to be broken, giving rise to precisely a continuum of states starting at $`M=0`$ as expected. In the computation of the correlation function illustrated in figure 3.b, we find a curious feature that it asymptotes to the inverse power law $`c/r^4`$ for large $`r`$. This behavior comes about due to the coupling $`0|T^{++}|n`$ with exactly massless states $`|n`$. The contribution to (23) from strictly massless states are given by $$\left(\frac{x^{}}{x^+}\right)^2F(x^{},x^+)=\left|\frac{L}{\pi }0|T^{++}(k)|n\right|^2\frac{M_n^4}{8\pi ^2k^3}K_4(M_nr)|_{M_n=0}=\left|\frac{L}{\pi }0|T^{++}(k)|n\right|_{M_n=0}^2\frac{6}{k^3\pi ^2r^4}.$$ (39) We have computed this quantity as a function of $`1/k`$ and extrapolated to $`1/k0`$ by fitting a line and a parabola to the computed values for finite $`1/k`$. The result of this extrapolation is illustrated in figure 4. The data currently available suggests that the non-zero contribution from these massless states persists in the large $`k`$ limit. Let us now turn to the model with (2,2) supersymmetry. The SDLCQ version of this model was solved in . The result of this computation can be applied to (23). The result is summarized in figure 5. This model appears to exhibit the onset of a gapless continuum of states more rapidly than the (1,1) model as the harmonic resolution $`k`$ is increased. Just as we found in the (1,1) model, this theory contains exactly massless states in the spectrum. These massless states appear to couple to $`T^{++}|0`$ only for $`k`$ even, and the overlap appears to be decreasing as $`k`$ is increased. We believe that this model is likely to exhibit a power law behavior $`c/r^\gamma `$ for $`\gamma >4`$ for the $`T^{++}`$ correlator for $`rg_{YM}\sqrt{N}`$ in the large $`N`$ limit. Unfortunately, the existing numerical data do not permit the reliable computation of the exponent $`\gamma `$. ## 4 Conclusion In this article, we have provided a prescription for computing the correlation function of the stress energy tensor $`T^{++}`$ in the SDLCQ formalism, which may be readily compared with predictions provided by a supergravity analysis following the conjecture of Maldacena. Such a comparison requires non-perturbative methods on the field theory side, and the SDLCQ approach appears at first sight to be particularly well suited to this task. Unfortunately, at the present time, high enough resolution calculation have not been made to evaluate expression (23) accurately enough in the case of SYM with 16 supercharges to reproduce (6). Significant progress is expected however when these calculation are moved from desk top computers to a supercomputer. Ultimately the main obstacle will be that the number of allowed states in the SDLCQ wavefunctions grows exponentially with the resolution. Nevertheless, a concrete well defined algorithm is a great starting point for further investigations, and additional insight may be gained by studying models with less supersymmetry, as we have done here. Thus, the answer to the question posed in the title is a qualified “yes.” Equation (23) may be compared against (6) using SDLCQ, and unless other computational approaches such as the lattice technique catches up, the SDLCQ approach remains the most viable option. ## Acknowledgments We thank D. Gross, S. Hellerman, S. Hirano, N. Itzhaki, N. Nekrassov, Y. Oz, J. Polchinski, and E. Witten for illuminating discussions. The work of AH was supported in part by the National Science Foundation under Grant No. PHY94-07194. The work of FA, OL, and SP was support in part by a grant from the United States Department of Energy.
no-problem/9906/nucl-ex9906005.html
ar5iv
text
# Light Fragment Yields from Central Au + Au Collisions at 11.5 A GeV/c ## Introduction The study of heavy-ion collisions at ultra-relativistic energies has been pursued with the intent of observing matter at extreme temperatures and densities, the eventual goal being the production and study of deconfined and/or chirally restored matter. Although the relativistic heavy-ion program has developed over a number of years, and various hadronic observables have been studied, a complete characterization of the state formed during the high density phase of the collision is still outstanding. Measurements of light nuclei produced in the participant region of heavy-ion collisions are of interest since model calculations can be used to deduce measures of the source volume from the yield of nuclear fragments relative to that of protons . The yield of light nuclei is closely related to the two particle correlation between nucleons induced by final state interactions . This can be incorporated into a theoretical framework from which one can deduce the density of the interaction region at freeze-out. In this publication we will report on investigations of the characteristics of spectra of light nuclei utilizing the E877 spectrometer, located at Brookhaven National Laboratory’s (BNL) Alternating Gradient Synchroton (AGS). Spectra and yields of light nuclei, including deuterons, <sup>3</sup>H, <sup>3</sup>He, and <sup>4</sup>He will be presented and compared to similar measurements at other beam energies and for various colliding systems. ## Experimental Apparatus The E877 apparatus has already been described in . Global observables, single particle spectra of protons, pions and kaons and two particle correlations as well as collective flow have been studied by E877 at the AGS. The experimental setup primarily consists of three components: (1) a set of scintillators and silicon detectors for beam definition; (2) a near $`4\pi `$ calorimetric coverage and a large acceptance charged particle multiplicity measurement, providing global event properties; and (3) a forward spectrometer, allowing measurements of identified charged particles. Figure 1 shows schematically the E877 experimental setup for the 1994 run. The $`z`$ direction is defined to be along the beam and the $`y`$ axis is defined to be out of the page (vertical). The data presented here were collected in the fall of 1994. After event selection cuts were applied, approximately $`5.410^6`$ events for the 4% highest $`E_t`$ collisions were available for analysis. ### Beam Definition The beam in the C5 beam line at the AGS for the 1994 run had approximately $`10^4`$ particles per spill with a 1s spill length. To verify the beam composition and direction, and provide the start time of the experiment a number of detectors were placed upstream of the target. A series of four plastic scintillators (S1, S2, S3, S4) were used for the time reference definition in the experiment. Scintillators S1 and S3 were ring-shaped veto detectors while scintillators S2 and S4 were ellipsoidal detectors placed at $`45^{}`$ relative to the beam. Fast rise time phototubes were utilized on S2 and S4 providing the interaction time with a typical Gaussian standard deviation of $`\sigma =3540`$ ps. Two silicon microstrip detectors (BVER’s) measured the beam angle in the $`xz`$-plane and were located between S2 and S4. The detectors were 300 $`\mu `$m thick and composed of 320 strips with a 50 $`\mu `$m pitch. This provided a measurement of the beam angle with a resolution of about $`40`$ $`\mu `$rad and determined the $`x`$ coordinate of the beam at the target to 300 $`\mu `$m. The BVERs were also used to tag and reject those events in which two or more particles traversed the experiment within the same 1 $`\mu `$s time interval. To eliminate beam components with incorrect nuclear charge number, a silicon semiconductor counter (SILI) 87 microns thick was placed $`2.5`$ cm upstream of the target to measure the energy loss of the beam particles. This detector provided a unit charge resolution of the incoming beam particles. ### Event Characterization For the 1994 run we used a AU target having a thickness corresponding to 1% of an interaction length. Two large, highly segmented calorimeters were used to quantify the centrality and reaction plane of the collisions. Together these detectors cover nearly 4$`\pi `$ in the center-of-mass frame. The Target Calorimeter (TCAL) consists of 832 NaI(Tl) crystals surrounding the target with a polar angle coverage of $`48^{}<\theta <135^{}`$ ($`0.88<\eta <0.81`$). Each crystal has a depth of 13.8 cm, or about 5.4 radiation lengths (0.34 hadronic interaction lengths). The Participant Calorimeter (PCAL) is a finely segmented lead/iron/scintillator calorimeter consisting of 16 azimuthal, 8 radial, and 4 depth sections. It covers a polar angle region of $`1^{}<\theta <47^{}`$ ($`0.83<\eta <4.7`$). There is a small opening in the PCAL ($`136<\theta _x<16`$ mrad; $`11<\theta _y<11`$ mrad) with iron wedges along the inside which define the acceptance of the forward spectrometer. Events are characterized by their centrality as determined by the transverse energy measured by the calorimeters. $$E_t=\underset{i}{}e_isin\theta _i$$ (1) where $`e_i`$ is the energy deposited in one cell of the calorimeter and $`\theta _i`$ is the polar angle of the cell with respect to the beam axis; the sum is performed over all PCAL cells. As was noted in , the overall systematic error associated with the measurement of $`E_t`$ was determined to be less than 4% for all collisions considered in this analysis. The transverse energy is then used as a measure of the centrality of the collision by noting that $$\frac{d\sigma }{dE_t}=\frac{s}{n_tB}\frac{dN}{dE_t}$$ (2) where $`\sigma `$ is the cross-section, $`B`$ is the incident number of beam particles, $`s`$ is the trigger down-scale factor ($`s1`$), and $`n_t`$ is the number of target atoms per unit cross-sectional area. For the experiment we used a Au target with an areal density of 540 mg/cm<sup>2</sup>. This corresponds to a 1% interaction probability for incident Au ions at an calculated inelastic cross section of $`\sigma _{in}`$ = 6.1 b. Central collisions of a certain percentage ($`w`$) of the inelastic cross section were then selected by adjusting the lower limit E$`{}_{}{}^{low}{}_{t}{}^{}`$ such that : $$w=\frac{_{E_t^{low}}^{\mathrm{}}\frac{d\sigma }{dE_t^{}}𝑑E_t^{}}{\sigma _{in}}.$$ (3) ### Forward Spectrometer The momentum and time-of-flight of all charged particles accepted into the forward region are measured using two drift chambers, four multiwire proportional chambers and two time of flight hodoscope walls. These detectors are also shown schematically in Figure 1. In addition, a set of multiwire proportional chambers was placed 2 m from the interaction region; the two chambers were separated from each other by 25cm (VTXA and VTXB). They were not used in this analysis. The analyzing magnet, situated downstream of the vertex chambers and spanning from $`z=260`$ cm to $`z=350`$ cm, could be run in two polarities in order to study possible systematic errors in the momentum measurement and to optimize the acceptance for positive and negative particles. The magnet had a maximum field of $`3,353\pm 4`$ Gauss with a polarity that bent positive particles in the negative $`x`$ direction and an integrated $`BdL`$ of .3487 Tm . Two drift chambers (DC2 and DC3), located at $`z=700`$ cm and $`z=1150`$ cm from the target, respectively, provided track x and y information and were used to determine the rigidity of each track; each drift chamber was composed of six planes of wires which measured the position in the bending plane of the magnet, i.e. the xz-plane, with a resolution of $`250`$ $`\mu `$m ($`350`$ $`\mu `$m) for DC2 (DC3). The anode wires in DC2 and DC3 were parallel to the y-direction, and the cathode planes in both chambers were segmented into chevron shaped pads. The signals from the pads provided y measurements of charged particles with a resolution of typically a few percent of the pad size. Each drift chamber had two segmentation size regions, resulting in $`y`$-resolutions of 2.3 mm (15 mm) in DC2 and 4.3 mm (36 mm) in DC3 depending on the pad size. The multi-wire proportional chambers (MWPC’s) with vertical anode wires space 5.08 mm apart were situated between the two drift chambers. The MWPC’s were used to ease tracking in the high multiplicity environment of Au+Au collisions. An array of 160 plastic scintillators (TOFU) aligned vertically and placed behind DC3 was used to measure time-of-flight with a time resolution in conjunction with the beam start counters of about $`85`$ ps. It also provided an additional y-measurement with a precision of about 1 cm in the y-direction. The FSCI (Forward Scintillator) was a second hodoscope which also measured the time-of-flight at a distance of 31 m from the target with a resolution of 350 ps. It was utilized in this analysis only to study of TOFU systematics since it had a lower resolution and smaller acceptance than the TOFU. ## Data Analysis Particle rigidity was determined from the radius of curvature of a track ($`R`$) through the magnet and the value of the magnetic field $`B`$: $$\mathrm{rigidity}=\frac{\sqrt{p_{x}^{}{}_{}{}^{2}+p_{z}^{}{}_{}{}^{2}}}{Ze}=RB$$ (4) where $`Ze`$ is the charge of the particle. The vertical momentum component, $`p_y`$, is calculated from the vertical position of the track at the TOFU. The reconstructed mass of each particle is then given by $$m^2c^4=\frac{p^2c^2}{\beta ^2\gamma ^2}$$ (5) where $`\beta `$ and $`\gamma `$ are calculated from the time-of-flight measurement coupled with the calculated path length to the associated TOFU scintillator. After all momentum calculations, Fig. 2 shows the E877 acceptance for protons and deuterons as a function of transverse momentum and rapidity for the 1994 run. The charge of each particle is determined from the energy deposition in the TOFU slat. Shown in Figure 3 is the distribution of pulse-heights for those particles corresponding to the deuteron mass peak at beam rapidity. A clear $`Z=1`$ peak is visible as well as a peak corresponding to $`Z=21`$, from two $`Z=1`$ tracks striking the same TOFU slat. In addition, the $`Z=2`$ peak corresponding to <sup>4</sup>He is clearly visible. The resulting distribution is fit to three Landau functions with a non-linear response due to the saturation of the photomultiplier tubes associated with the TOFU. The standard E877 analysis cut on the TOFU pulse-height distribution at $`1.4`$ times the 1 MIP pulse height corresponding to about a $`10`$% loss in the $`Z=1`$ yield with no contamination from $`Z=2`$. ### Deuterons Since $`\delta m/m\delta p/p`$ and since the momentum and time of flight measurements worsens at larger momenta, deuteron selection is complicated by the tail of the proton peak at high momenta. To account for this effect, a systematic background subtraction technique was developed. In each rapidity and transverse momentum bin the yield of deuterons is calculated by fitting the mass distribution in this region to the proton and deuteron mass peaks plus a background between these peaks due to misidentified protons from resonance decays. Shown in Figure 4 is the reconstructed distribution of mass-squared in several $`p_t=\sqrt{(p_{x}^{}{}_{}{}^{2}+p_{y}^{}{}_{}{}^{2})}`$ and $`y_{deut}=0.5\mathrm{ln}(E+p_z)/(Ep_z)`$ bins, where the rapidity was calculated assuming the deuteron mass. At low momenta, the deuteron and proton peaks are clearly distinguishable. In this region of phase space, the mass-squared distribution is described by two Gaussians and either an exponential or linear background<sup>*</sup><sup>*</sup>*The systematic error associated with the choice of the background function has been found to be less than 5% .. At higher momenta, the proton and deuteron mass-squared distributions broaden, due to the momentum resolution of the apparatus, and the background is minimal. In this region of phase-space, shown in Figure 4B, the mass-squared distribution is described by two Gaussian distributions. In the $`2.6<y<3.0`$ region, the momentum resolution smearing coupled with the approximately equal particle yields for protons and deuterons, causes the peaks to overlap (Figure 4C), resulting in a loss of separation of deuterons from protons in this region. However, at deuteron rapidity $`y>3.0`$ the contamination of protons is minimal because the proton would need to be traveling at twice the beam momentum. Thus, a deuteron peak is again identifiable (Figure 4D) and the yield can be determined with a single Gaussian. The widths of the proton and deuteron mass peaks were found to be consistent with the known time of flight and momentum resolution of the experiment. The yield of deuterons ($`N_d`$) in each small bin of phase space is determined from the parameters of a Gaussian fit to the deuteron peak: $$N_d=\sigma _dA_d\sqrt{2\pi }/\mathrm{\Delta }m^2$$ (6) where $`\sigma _{d}^{}{}_{}{}^{2}`$ and $`A_d`$ are the variance and amplitude of the Gaussian distribution fit to the deuteron peak, and $`\mathrm{\Delta }m^2`$ is the bin width. Those regions of momentum space that fall between the first two regions outlined above, Figure 4A and B, respectively, have been fit with both two Gaussians as well as two Gaussians plus a background in order to determine the systematic error due to the choice of fit funcion. The resulting yields of deuterons agree within 5%. ### Tritons The extraction of a triton signal is complicated by the fact that its mass-squared distribution is wider than the deuteron’s due to multiple scattering and its yield is 1-2 orders of magnitude lower than that of the deuteron. However, a technique similar to that used for the deuterons can also be utilized to measure the yield of tritons in the collisions studied here. In Figure 5 is shown the reconstructed mass-squared distribution in various $`y_{trit}`$ bins for $`p_t<0.5`$ GeV/c. A clear triton peak is visible. The yield of tritons was calculated for each rapidity and $`p_t`$ bin by fitting this distribution with a Gaussian at the mass of the triton with an exponential background from the deuteron tail; the choice of an exponential or Gaussian background was found to have less than a 5% effect on the resulting yields. Several examples of these fits are also shown in Figure 5. ### Helium Isotopes Helium isotopes can be separated from $`Z=1`$ species by cutting on energy deposition in the scintillators as described earlier. <sup>3</sup>He are identified by recalculating the mass-squared according to eqn. 5 assuming $`Z=2`$. This will cause all $`Z=1`$ particles which pass the scintillator pulse-height cut to be found at double their actual mass. The <sup>3</sup>He peak in this case is found between, and clearly differentiated from, the proton and deuteron residual peaks, respectively located at $`m^24`$ and $`m^214`$ (GeV/c)<sup>2</sup>, as shown in Figure 6. The <sup>3</sup>He peak is buried under the proton tail for $`y>2.0`$ and, in the other rapidity bins, the yield of <sup>3</sup>He may be extracted by fitting the mass-squared distribution by two Gaussians corresponding to the <sup>3</sup>He and proton peaks. The method described above, i.e. cutting on the energy loss in the TOFU and, subsequently, fitting the peak in the $`m^2/Z^2`$ spectrum, was used to determine the deuteron, triton and <sup>3</sup>He yield. However, this method cannot be used to determine the yield of <sup>4</sup>He because the <sup>4</sup>He peak in $`m^2/Z^2`$ space is buried under the significantly larger deuteron peak. We therefore use a complimentary approach to identify <sup>4</sup>He: we first cut on the $`m^2/Z^2`$ peak and then fit the $`Z=2`$ peak in the TOFU pulse height distribution. As described earlier, this distribution has two peaks, corresponding to one and two $`Z=1`$ particles, at all but beam rapidity where an additional $`Z=2`$ peak is evident as shown in Figure 3. These peaks are fit, at beam rapidity for $`p_t<2`$ GeV/c, to Landau distributions and the yield is then determined from the integral under the fit to the $`Z=2`$ distribution. The overall systematic errors from this procedure is estimated to be less than 10%. ## Results ### Mid-rapidity Shown in Fig. 7 are deuteron invariant multiplicities for the 4% highest $`E_t`$ collisions as a function of $`p_t`$ and $`y`$ with bin widths of 20 MeV in $`p_t`$ and 0.1 in $`y`$. The invariant multiplicity is flat as a function of $`p_t`$ and $`y`$ over most of the measured range. In Figure 8 are shown the measured invariant multiplicities of <sup>3</sup>H and <sup>3</sup>He for the same centrality. Plotted is the differential cross section at the center of the bin assuming an underlying thermal distribution with a shape similar to what is measured The resulting slope and magnitude of the $`p_t`$ spectra was found to be insensitive to a wide range or reasonable assumptions of the true distribution in the acceptance correction.. The resulting measured cross section is not affected significantly by the assumed distribution. The distributions are presented as a function of rapidity for two $`p_t`$ regions. The error bars reflect the statistical uncertainty in the fit procedure and do not include the systematic errors, estimated to be about $`20\%`$ for both <sup>3</sup>He and <sup>3</sup>H, dominated by the uncertainty in the assumed background. ### Beam-rapidity In Figure 9 are shown invariant multiplicities for deuterons at beam rapidity ($`y_{beam}=3.2`$) for the 4% and 10-20% highest $`E_t`$ events. As noted above, deuterons at beam rapidity are easily extracted from the mass-squared distribution. The transverse momentum distribution of beam rapidity deuterons for the 4% highest $`E_t`$ are distinctly harder, i.e. show a larger inverse slope parameter, than deuterons from the lower $`E_t`$ collisions. Over the range in $`p_t`$ in which we measure, a fit to an $`m_t`$ Boltzmann distribution in the 3.0-3.1 rapidity bin gives $`T_B=49.6\pm 1.4`$ MeV and $`T_B=83\pm 2`$ MeV for the 4% and 10-20% $`E_t`$ bins, respectively. However, the reduced $`\chi ^2`$ of each of these fits is greater than 4, implying a Boltzmann shape does not describe the measured distributions well. In beam rapidity <sup>4</sup>He spectra are shown in Figure 10. In the lower $`E_t`$ range (10-20%), the <sup>4</sup>He yield is a factor of 10 higher than for the higher $`E_t`$ (4%), indicating that beam rapidity <sup>4</sup>He are primarily produced by projectile fragmentation. ## Model Comparisons Light nuclei from heavy-ion collisions have historically been studied via a coalescence model which assumes cluster formation takes place at freeze-out. Since the binding energy of the deuteron is $`2.24`$ MeV, it is easily broken apart in the fireball region where current models estimate the temperature to be 120-150 MeV . In the coalescence formulation the probability of forming a deuteron is greatest when a proton and neutron at freeze-out have a small relative momentum. Assuming that the proton and neutron distributions are similar, the deuteron momentum distribution is then given by $$E_d\frac{d^3N_d}{dp^3}(p_d)=B_d\left(E_p\frac{d^3N_p}{dp^3}(p_p)\right)^2$$ (7) where the particle momenta obey $`p_d=2p_p`$ which can be easily generalized to heavier fragments with mass number $`A`$ as $$E_A\frac{d^3N_A}{dp^3}(p_A)=B_A\left(E_p\frac{d^3N_p}{dp^3}(p_p)\right)^A$$ (8) where $`p_A=Ap_p`$. In eqns. 7 and 8, $`B_A`$ ($`B_d`$ in the case of the deuteron) is a phenomenological parameter known as the coalescence constant. In proton-nucleus collisions $`B_A`$ has been related to the interplay between the binding energy of the deuteron and the optical potential of the target nucleus . For nucleus-nucleus collisions it was recognized that the idea of a nuclear optical potential was no longer meaningful and Schwarzschild and Zupančič expressed the coalescence parameter in terms of a momentum difference $`p_0`$. A proton and neutron would coalesce to form a deuteron if, at freeze-out, their momenta differ by less than $`p_0`$. More generally, for the case of a source of Z protons and N neutrons with mass number of the cluster A, $$B_A=\left(\frac{2s_A+1}{2^A}\right)\frac{R_{np}^{}{}_{}{}^{N}}{N!Z!}\left(\frac{4\pi }{3}p_{0}^{}{}_{}{}^{3}\right)^{A1}$$ (9) where $`s_A`$ is the spin of the cluster and $`R_{np}`$ is the neutron to proton ratio in the source. Therefore, the coalescence parameter in these models is simply related to the momentum difference between corresponding protons and neutrons and is described by a simple step function in the coalescence probability at $`p_0`$. Note that no dependence on the colliding system exists in this model which described deuteron and <sup>3</sup>H/<sup>3</sup>He production very well in nucleus-nucleus reactions at Bevalac energies (about $`0.12`$GeV$``$A) and in p-nucleus collisions, at FNAL energies. For all of these energies, only a single constant $`B_A`$ was needed to describe the light fragment distributions based on proton spectra in minimum bias collisions. However, for nucleus-nucleus collisions at AGS energies and above a dramatic drop below the Bevalac values in the coalescence parameter and a strong dependence on the centrality of the collision was observed. For beam energies where nuclear collisions produce a large number of secondary particles, i.e. AGS energies and above, $`B_d`$ was no longer independent of beam energy or composition presumably because the source decoupled at a size larger than the mean size of a deuteron, $`R_{rms}=2.1`$ fm . In this case, $`B_d`$ was theoretically deduced to be inversely proportional to the volume of the source at freeze-out. In the model of Sato and Yazaki , for example, $$B_d=\frac{3}{4}\frac{(8\pi )^{3/2}}{Z!N!}\left[\frac{\nu _d\nu }{(\nu _d+\nu )}\right]^{3/2}$$ (10) where $`\nu _d=.20fm^2`$ is the Gaussian wave function parameter for a deuteron and $`\nu `$ is directly related to the radius of the system at freeze-out by $$R_{rms}=\sqrt{R_{x}^{}{}_{}{}^{2}+R_{y}^{}{}_{}{}^{2}+R_{z}^{}{}_{}{}^{2}}=(3/2\nu )^{1/2}$$ (11) where $`R_x`$, $`R_y`$ and $`R_z`$ are the Gaussian radii. Other models also utilize a wave-function description and a density matrix formalism to describe deuteron formation. With the exception of large cascade codes as well as recent hydrodynamically motivated source parametrizations , however, none of them explictely account for transverse collective expansion. Though such an approach is not excluded from the Sato and Yazaki approach , nothing beyond sources with no space-momentum correlations were considered in their presentation. Further, though there have been some attempts at a fully relativistic treatment of coalescence , this generalization is non-trivial because of difficulties in the proper relativistic treatment of bound states . ### Coalescence Parameter $`B_d`$ The coalescence parameter, as defined in Eq. 7, can be calculated from E877 measurements of protons and deuterons. To properly calculate $`B_d`$ for our data set, the fraction of protons from $`\mathrm{\Lambda }`$ decays must be subtracted from the measured proton spectra in order to obtain the proton distribution at freeze-out. This correction was performed utilizing the measured $`\mathrm{\Lambda }`$ yield in Au+Au collisions and knowledge of the acceptance and reconstruction efficiency of the experimental apparatus. Over the measured acceptance region, the hyperon contribution to the proton spectra was found to be no more than 10%. After correcting for hyperon contributions, $`B_d`$ can be calculated as a function of $`p_t`$ and y. We notice, within the acceptance of the present measurement, no variation of $`B_d`$ as a function of $`p_t`$. However, there is a significant rapidity dependence of the average $`B_d`$ for all measured $`p_t`$ as shown in Figure 11. If $`B_d`$ is related to volume at freeze-out, such a result may imply a changing effective freeze-out volume as a function of rapidity. The measured change in $`B_d`$ of nearly a factor of two from mid-rapidity to $`y=2.6`$ would correspond to a decrease in the effective source radius of $`1.25`$. Figure 12 shows a comparison of the coalescence parameter deduced from experiments at Bevalac energies with those obtained at the AGS in in Si + Al , Si + Pb , Au + Pt , and Au \+ Au collisions. The error on the Au+Au data point corresponds to the minimum and maximum range of $`B_d`$ values shown in Figure 11 including the statistical error bars. All presently measured values of $`B_d`$ for high $`E_t`$ Au+Au collisions are significantly lower than the results for Si+Al and Si+Pb. This is consistent with the interpretation that $`B_d`$ depends on the volume at freeze-out. Two proton and two pion correlations measurements have found that the effective nucleon source size is larger for the Au+Au system than for lighter systems . The coalescence parameter measured in these collisions is nearly one-tenth of that measured in central Au+Au collisions at Bevalac, implying in the models of that the volume at feeze-out in AGS Au+Au collisions is five to ten times that at Bevalac energies. Such a simple geometrical interpretation would be inconsistent with two particle correlations measured at the AGS . A possible reason for such an inconsistency is that the models which deduce an inverse volume dependence do not account for collective motions in the source. As has been shown in , the yield of deuterons is highly dependent on the amount of collective motion in the source. Hence, $`B_d`$ is not simply related to the volume but also the position-momentum correlations at freeze-out. If the source is perfectly correlated, then proton-neutron pairs that freeze-out near one another in space will have similar momentum vectors. The deuteron yield would be larger in this case than for a random source. As such, any consistent model of deuteron production must account for correlations in the source. Therefore, in order to completely describe deuteron production, a more sophisticated model of the source that includes dynamical correlations must be formulated. In large cascade codes it is difficult to study the influence of the source parameters. We thus used a simple dynamical model in the following section to study such effects. In Fig. 13 is the corresponding coalescence parameter for <sup>3</sup>H and <sup>3</sup>He extracted from this data set compared with Bevalac , AGS , and SPS values. Similar to what is observed in $`B_d`$, the three nucleon coalescence parameter drops dramatically at AGS energies, consistent with an increase in the effective source size. ## A Dynamical Model for Deuteron Formation An analytic, hydrodynamically motivated model, devised by Chapman, et al. has been implemented to describe the particle source distribution in heavy ion collisions . This model uses an emission function ($`S`$) for a finite, expanding, locally thermalized source $$E_K\frac{d^3N}{d^3K}=d^4xS(x,K)$$ (12) where $`x`$ and $`K`$ are the position and momentum 4-vectors, respectively. The source includes longitudinal and transverse collective motion with an underlying temperature, given by $$S(x,K)=\frac{M_t\mathrm{cosh}(\eta Y)}{(2\pi )^{7/2}\mathrm{\Delta }\tau }exp\left[\frac{Ku(x)}{T}\frac{(\tau \tau _0)^2}{2(\mathrm{\Delta }\tau )^2}\frac{r^2}{2R^2}\frac{(\eta \eta _0)^2}{2(\mathrm{\Delta }\eta )^2}\right]$$ (13) where $`\eta `$ $`=`$ $`{\displaystyle \frac{1}{2}}ln[(t+z)/(tz)]`$ $`\tau `$ $`=`$ $`\sqrt{(t^2z^2)}`$ $`Y`$ $`=`$ $`{\displaystyle \frac{1}{2}}ln[(1+\beta _l)/(1\beta _l)]`$ $`Ku(x)`$ $`=`$ $`M_t\mathrm{cosh}(\eta Y)\mathrm{cosh}\eta _t(r)K_t{\displaystyle \frac{x}{r}}\mathrm{sinh}\eta _t(r)`$ $`\eta _t(r)`$ $`=`$ $`\eta _f(r/R).`$ In this model $`(x,y,z,t)`$ is the freeze-out position four vector, $`r=\sqrt{x^2+y^2}`$ and $`\beta _l`$ and $`M_t`$ are the longitudinal velocity and transverse mass, respectively. $`K_t`$ is the transverse momentum and $`\eta _f`$ is the transverse flow velocity. Furthermore, $`\mathrm{\Delta }\tau `$, $`\mathrm{\Delta }\eta `$ and $`R`$ describe the respective Gaussian widths of the proper time, longitudinal velocity and one dimensional Gaussian radius. For more information on the source function, see . To determine deuteron yields with this framework, a program devised by R. Mattiello was implemented. It utilizes a method that has been applied to bombarding energies from 1 GeV per nucleon in association with the the intranuclear cascade model and QMD up to AGS and SPS energies with the cascade models RQMD and ARC . In this approach, the number of deuterons is given by a summation over all proton and neutron pairs at freeze-out accounting for the Wigner density of the Hulthén deuteron wave function . This method is equivalent to the density matrix formalism of Sato and Yazaki in the case of a static source. Therefore, by projecting the neutron/proton position-momentum distribution on the deuteron wave-function via this Wigner method, we may determine the distribution of deuterons in phase-space given any particular source of nucleons. Shown in Figure 14 is the distribution of the reduced $`\chi ^2`$ minus the minimum $`\chi ^2`$ ($`\chi _{min}^2=0.96`$) with respect to the measured deuteron yields calculated as a function of $`T`$ and $`R`$ with a resolution of 10 MeV and 0.25 fm respectively. For this study we assume that $`\tau _0=3`$fm/c and $`\mathrm{\Delta }\tau =1`$fm/c in Eqn. 13. For SPS energies, $`\mathrm{\Delta }\tau 1.53.0`$ fm/c from two particle correlations while noted that $`\tau _09`$fm/c. Similar numbers for AGS energies have not been published. The technique we utilize for mapping out this $`\chi ^2`$ space is the following: For each value of temperature $`T`$ and one dimensional Gaussian radius $`R`$, the transverse and longitudinal proton spectra uniquely define the transverse ($`\eta _f`$) and longitudinal ($`\mathrm{\Delta }\eta `$) boost velocities. Every value of $`T`$ and $`R`$ in Figure 14 can describe the proton transverse momentum spectra given a particular $`\eta _f`$ at mid-rapidity. We have found that the proton rapidity distribution in this model poorly reproduces the data , with a reduced $`\chi ^2`$ on the order of 10 for all choices of $`T`$, $`R`$ and $`\mathrm{\Delta }\eta `$. Furthermore, the shape of the transverse momentum distribution of nucleons in this model is assumed to be constant as a function of rapidity, in conflict with experimental data . The assumptions inherent in the model regarding a hydrodynamical source are not entirely valid for nucleons at AGS energies due to these complications in the longitudinal direction. Therefore, in the studies presented here we only consider the nucleon production at mid rapidity and implement this model by fitting $`\eta _f`$ to the proton $`p_t`$ distribution only at mid-rapidity and constrain $`\mathrm{\Delta }\eta `$ to approximate the rapidity distribution. The deficiencies in the model near beam-rapidity should not influence the results we show here for mid-rapidity and indeed the model describes the transverse momentum distribution at mid-rapidity quite well ($`\chi ^21`$). For this analysis, the proton spectra from E866 Au+Au collisions are utilized due to their coverage of protons at mid-rapidity. The model is therefore fit with these two parameters to the proton spectra. After the proton distribution is determined, the deuteron yields at mid-rapidity are calculated with the coalescence algorithm described above. Results are compared to deuterons measured by E877 at mid-rapidity and $`p_t<0.6`$ GeV/c. The resulting $`\chi ^2`$ distribution strongly constrains the temperature and transverse radius in this model of the nucleon source though the two variables are highly correlated. Studies of proton correlations for the Au+Au system at AGS energies concluded there was an increase of the transverse freeze-out distribution of nucleons beyond the Au rms radius of 3.1fm. This is consistent with the current results where all reasonable values of temperature result in a value of the transverse radius significantly larger than the rms radius of a Au nucleus. Furthermore, measures of the temperature at freeze-out in Au+Au collisions at the AGS of 120-150 MeV are also consistent with the results of this study. Note, however, that this method has not constrained $`\mathrm{\Delta }\tau `$ or $`\tau _0`$. Varying these attributes from the values chosen above by a factor of two can vary the yield of deuterons by 10%. An increase in either $`\mathrm{\Delta }\tau `$ or $`\tau _0`$ would decrease the yield of deuterons in the model and, as a result, would shift the $`\chi ^2`$ curve of Figure 14 to lower temperatures and radii. Therefore, deuterons cannot be used to exclusively determine source parameters but must be used in conjunction with complimentary observables to produce a coherent view of the source. ## Conclusions We have measured light fragment yields at both mid and beam rapidities. Invariant multiplicities have been presented for deuterons, <sup>3</sup>H, and <sup>3</sup>He at mid-rapidity. Deuterons at mid-rapidity have been interpreted in a coalescence framework. The resulting coalescence parameter ($`B_d`$) is consistent with an increase in source size from collisions of lighter nuclei at the same energy. However, a clear dependence on rapidity of $`B_d`$ implies the assumption of a constant effective volume as a function of rapidity is incorrect. We have introduced a technique for using deuterons as a sensitive constraint on source parameters in a heavy ion collision by incorporating a simple, hydrodynamically motivated nucleon source function at mid-rapidity coupled with a coalescence algorithm. The results show that deuteron yields are sensitive to a convolution of volume and collective effects, similar to other correlations measurements. These results can be used in conjunction with other measures of the source to tightly constrain theoretical models and assumptions. First measurements of beam rapidity deuterons and <sup>4</sup>He at AGS energies have also been made. The resulting distributions have a strong $`E_t`$ dependence, consistent with a production mechanism dominated by spectator fragmentation. ## Acknowledgements We thank the BNL AGS and tandem operations staff and Dr. H. Brown for providing the beam. This work was supported in part by the U.S. DoE, the NSF, the NSERC, Canada and CNPq, Brazil. We are grateful to Dr. R. Mattiello, who provided the coalescence code utilized in this study and to Drs. J. H. Lee and C. Chasman for providing us with the results of proton distribution measurements made by the E866 collaboration.
no-problem/9906/cond-mat9906172.html
ar5iv
text
# The thermodynamical liquid-glass transition in a Lennard-Jones binary mixture \[ ## Abstract We use the results derived in the framework of the replica approach to study the liquid-glass thermodynamic transition. The main results are derived without using replicas and applied to the study of the Lennard-Jones binary mixture introduced by Kob and Andersen. We find that there is a phase transition due to the entropy crisis. We compute both analytically and numerically the value of the phase transition point $`T_K`$ and the specific heat in the low temperature phase. \] In recent times there have been many progresses in the analytic understanding of the thermodynamics transition of glasses . Beyond technicalities the basic assumption is that a glass is very near to a frozen liquid. This quite old statement can be rephrased as a liquid is very near to a heated glass. In other words the configurations of a glass at low temperature are not far from those of a liquid. Therefore, if we use some smart method to explore the phase space in the liquid phase, we can find the properties of the low temperature glassy phase. This strategy has been put into action using the replica theory. One finds, as output of an explicit computation, that the glass transition is characterized by the vanishing of the configurational entropy (the so called complexity) and the low temperature phase is described by one step replica symmetry breaking with a non-vanishing non-ergodic parameter at the phase transition point. Replica theory is a very powerful tool, but it has the disadvantage that many of the underlying physical hypothesis cannot be seen in a clear way. We will rederive some of the main result of without using the replica formalism . We suppose that below some temperature (to be identified with the mode coupling transition ), the phase space of the system can be approximately divided into regions (which we will call valleys) which are separated by high barriers . This approximation becomes better and better when we decrease the temperature and it becomes exact below the thermodynamics phase transition temperature $`T_K`$ of the glass , where the viscosity should diverges (unfortunately no quantitative predictions have been obtained microscopically on the behaviour of the viscosity beyond the Adams Gibbs argument). In a first approximation each valley can be associated to one inherent structure, i.e. one minimum of the potential energy . Let us consider a generic system with $`N`$ particles with Hamiltonian $`H(C)`$, $`C`$ denoting the generic configuration of the system. The partition function both in the liquid phase at low temperature and in the glass phase can be written as $`Z(\beta )={\displaystyle \underset{a}{}}\mathrm{exp}(\beta Nf_a(\beta ))=`$ (1) $`{\displaystyle 𝑑𝒩(f,\beta )\mathrm{exp}(\beta Nf)},`$ (2) where $`f_a(\beta )`$ is the free energy density of the valley labeled by $`a`$ at the temperature $`\beta ^1`$ and $`𝒩(f,\beta )`$ is the number of valleys with free energy density (per particle) less than $`f`$. In the glassy phase this sum is dominated by the valleys with minimal free energy, while in the liquid phase an exponentially large number of valleys contributes to the partition function. It is normally assumed that in glass forming systems, for large $`N`$ and $`f>f_0(\beta )`$, we have that $$𝒩(f,\beta )=\mathrm{exp}(N\mathrm{\Sigma }(f,\beta )),$$ (3) where the configurational entropy, or complexity, $`\mathrm{\Sigma }(f,\beta )`$ is positive in this region and vanishes at $`f=f_0(\beta )`$. The quantity $`f_0(\beta )`$ is the minimum value of the free energy: $`𝒩(f,\beta )`$ is zero for $`f<f_0(\beta )`$. The key point is the computation of the function $`\mathrm{\Sigma }(f,\beta )`$ for $`f>f_0(\beta )`$. It can be done using liquid theory because in this region we are still in the liquid phase. The location of the zero of $`\mathrm{\Sigma }(f,\beta )`$ will tell us the value of the free energy in the glassy phase. At this end it is convenient to consider the generalized partition function $$Z(\gamma ;\beta )\mathrm{exp}(N\gamma \mathrm{\Phi }(\gamma ;\beta ))=\underset{a}{}\mathrm{exp}(\gamma Nf_a(\beta )).$$ (4) The physical meaning of $`\gamma `$ is will be clearer later. It is evident that $$\gamma \mathrm{\Phi }(\gamma ;\beta )=\gamma f\mathrm{\Sigma }(\beta ,f),f=\frac{(\gamma (\mathrm{\Phi }(\gamma ;\beta )))}{\gamma }.$$ (5) The complexity can be simply obtained from $`\mathrm{\Phi }(\gamma ;\beta )`$ in the same way as the entropy can be obtained from the usual free energy. The crucial step consists in writing $`Z(\gamma ;\beta )=`$ (6) $`{\displaystyle 𝑑C\mathrm{exp}\left(\gamma H(C)N\gamma f(\beta ,C)+N\gamma f(\gamma ,C)\right)}=`$ (7) $`{\displaystyle 𝑑C\mathrm{exp}\left(\gamma H(C)N\gamma \widehat{f}(\beta ,C)+N\gamma \widehat{f}(\gamma ,C)\right)},`$ (8) where $`\widehat{f}(\beta ,C)=f(\beta ,C)f(\mathrm{},C)`$ and $`f(\beta ,C)`$ is a function that is constant in each valley and it is equal to the free energy density of the valley to which the configuration $`C`$ belongs. In deriving eq. (8) we have assumed that the value of $`\gamma `$ is high enough that all the configurations $`C`$, which contribute to the integral for large $`N`$, belong to some valley: We have also assumed that there is a one to one correspondence among the valley at inverse temperatures $`\beta `$ and at $`\gamma `$. Before entering into the computation of $`\widehat{f}(\beta ,C)`$ it is useful to make the so called quenched approximation, i.e. to make the following approximation inside the previous integral: $$\mathrm{exp}(A\widehat{f}(\beta ,C))=\mathrm{exp}(A\widehat{f}_\gamma (\beta )),$$ (9) where $`\widehat{f}_\gamma (\beta )`$ is the expectation value of $`\widehat{f}(\beta ,C)`$ taken with the probability distribution proportional to $`\mathrm{exp}(\gamma H(C))`$. The quenched approximation would be exact if the temperature dependance of the energy of all the valleys would be the same, apart from an overall shift at zero temperature. In other words we assume that the minima of the free energy have different values of the free energy but similar shapes. More refined computations can be done to compute systematically the corrections to the quenched approximation. This approximation would be certainly bad if we were using the free energy $`\widehat{f}(\beta ,C)`$ at the place of $`f(\beta ,C)`$ because the zero temperature energy strongly varies when we change the minimum. We finally find $$\mathrm{\Phi }(\gamma ;\beta )=F_L(\gamma )+\widehat{f}_\gamma (\beta )\widehat{f}_\gamma (\gamma ),$$ (10) $`F_L(\gamma )`$ being the free energy of the liquid ($`S_L(\gamma )`$ will be the entropy of the liquid). A simple algebra shows that $$\mathrm{\Sigma }(\gamma ;\beta )=S_L(\gamma )S_\gamma (\gamma )+\widehat{f}_\gamma ^{}(\gamma )\widehat{f}_\gamma ^{}(\beta ),$$ (11) where $`\widehat{f}_\gamma ^{}(\beta )=\widehat{f}_\gamma (\beta )/\gamma `$. In the liquid phase, we find out that the configurational entropy is given by $$\mathrm{\Sigma }(\beta )=\mathrm{\Sigma }(\beta ;\beta )=S_L(\beta )S_\beta (\beta ),$$ (12) as expected: the entropy of the liquid is the entropy of the typical valley plus the configurational entropy. The thermodynamic transition is characterized by the condition $$\mathrm{\Sigma }(\beta _K)=0.$$ (13) In the glassy phase the free energy can be found by first computing the value of $`\gamma (\beta )`$ such that $$\mathrm{\Sigma }(\gamma (\beta );\beta )=0$$ (14) and evaluating the corresponding free energy. The quantity $`\gamma (\beta )`$ is the inverse of the effective temperature of the valley. It is easy to show (following ) that the previous formulae are completely equivalent to the replica approach. A strong simplification happens if we assume that the entropy of the valley can be evaluated in the harmonic approximation where we only keep the vibrational contributions. In this case we obtain for a system with $`M`$ degrees of freedom that the harmonic entropy of the valley near to a configuration $`C`$ is given by $$S(\beta (C))=\frac{M}{2}\mathrm{ln}\left(\frac{2\pi e}{\beta }\right)\frac{1}{2}\text{Tr}\left(\mathrm{ln}((C))\right),$$ (15) where $`(C)`$ is an $`M\times M`$ Hessian matrix (e.g. if $`H`$ depends on the coordinates $`x_i`$ we have that $`_{i,k}=^2H/x_ix_k`$). If our approximation were fully consistent, we should find that all the eigenvalues of $``$ (the so called INN, Instantaneous Normal Modes ) are positive. This is not the case, however the number of negative eigenvalues becomes very small at low temperature, still in the liquid phase, signaling that valleys can be approximately defined in this region. How can we evaluate the harmonic entropy in this case? Two rather similar possibilities are: (a) We compute $`\text{Tr}\left(|\mathrm{ln}((C))|\right)`$ instead of $`\text{Tr}\left(\mathrm{ln}((C))\right)`$. (b) We find the minimum of the Hamiltonian which is the nearest to $`C`$ (i.e. the corresponding inherent structure) and we use the spectrum of the Hessian at this point. We have checked numerically that the two methods give rather similar results and we will follow the second one in the numerical simulations presented in the paper. The framework has been set. We now segue into an explicit computation. Two possibilities are open: (a) We do an analytic computation of all the quantities which appear in the previous equations; (b) We extract them from numerical simulations. The first possibility is open only in relatively simple systems and further approximations are needed, the second one is viable for all systems. In these letter we shall explore both possibilities in the case of a realistic model for glasses, the binary mixture of particles (80% large particles, 20 % smaller particles) interacting via a Lennard-Jones potential, introduced by Kob and Andersen . This Hamiltonian should mimick the behaviour of some metallic glasses and it is one of the best studied and simplest Hamiltonian which do not lead to crystalization at low temperature. The numerical procedure does not present any serious difficulty. We have studied via Monte Carlo simulations a system of $`N=260`$ particles, in a cubic box with periodic boundary conditions at density $`\rho =1.2`$, starting from $`\beta ^{1/4}=0.02`$ and increasing $`\beta `$ by steps of $`\mathrm{\Delta }\beta ^{1/4}=0.02`$; we performed up to $`\mathrm{4\; 10}^6`$ MC steps at each $`\beta `$. We need to simulate the system at high temperatures in order to compute the entropy of the liquid (we use as reference entropy that of a perfect gas at $`\beta =0`$). The entropy is obtained using the formula $`S(\beta )=S(0)+_0^\beta 𝑑\beta ^{}(E(\beta )E(\beta ^{}))`$. Given an equilibrium configuration the near minimum of the potential is found by steepest descendent and the computation of the 780 eigenvalues of $`(C)`$ does not present any particular difficulty. The results for the total entropy of the liquid and for the harmonic part are shown in fig. (1) as function of $`T^{.4}`$ (a more detailed description of the simulations can be found in ref. ). The entropy of the liquid is remarkable linear when plotted versus $`T^{.4}`$, as it happens in many cases . In fig (2) we show the configurational entropy as function of $`T^{.4}`$. We fit it with a polynomial of second degree in $`T^{.4}`$. The extrapolated configurational entropy becomes zero at a temperature $`T_K=.31\pm .04`$, where the error contains systematic effects due to the extrapolation (similar conclusions have been reached in ref. ). There are many methods to compute analytically the free energy in the liquid phase which lead to integral equations for the correlation functions. Here we follow a simple procedure proposed in ref. . It mixes the HNC (hypernetted chain) and MSA (mean spherical approximation) closures by means of a single parameter $`\alpha (T)`$ that is chosen in order to reduce thermodynamic inconsistencies, minimizing the difference between two different ways of computing the compressibility. This technique allow us to compute the internal energy in the liquid with a reasonable approximation. The resulting integral equations for the correlations function of the two kinds of particles are transformed in a set on non linear differential equations by discretizing them with a space resolution of $`2^5`$ and a large distance cutoff of $`2^4`$ (we have varied these parameters and checked that this choice gives a reasonable accuracy). The resulting equations in $`3512`$ unknown have been solved using a package from the IMSL library. The computation of the spectrum is much more involved. We follow a simple approach which becomes exact if we assume that the range of the Hessian is much larger than the typical interatomic distance and we use the superposition approximation $$g(x_1x_2x_3)=g(x_1x_2)g(x_2x_3)g(x_3x_1)$$ (16) in the cases where it is needed . In order to compute the eigenvalues of the Hessian we compute analytically the moments of the eigenvalue distribution $$M_k=\text{Tr}(^k).$$ (17) Each moment can be written as the appropriate integral over the correlations functions. In this long range approximation some terms are dominant over the others. We keep only those terms and we introduce the superposition approximation into the appropriate places. We use the large range expansion to select the diagrams . After some computations (which we do not report here ) we find a spectral density $`\rho (e)`$ that has support only in the region of positive eigenvalues $`e`$ (real frequencies $`\omega =\sqrt{e}`$) and goes to zero as $`\omega `$ as expected. We can now put everything together: the final analytic predictions for the liquid and harmonic entropy are shown in fig. (1). The liquid entropy turns out to be very good, while there is a minor discrepancy for the harmonic entropy, likely due to the rather strong approximations we have done in the analytic computation. The analytic configurational entropy is shown in fig. (2). It becomes zero at $`T_K=.32`$, which is our analytic prediction for the thermodynamic transition (as a check we have fitted the analytic configurational entropy using the same procedure as for the numerical one and we have found $`T_K=.34`$). We have in our hands all the tools to compute analytically the free energy in the low temperature case. In fig. (3) we show the specific heat coming from the $`x`$ dependent part of the Hamiltonian as a function of the temperature (we must add the momentum contribution $`3/2`$ in order to get the total specific heat). Very similar results are obtained by using the extrapolated numerical entropies at the place of the analytic ones. We notice that the Dulong Petit law is extremely well satisfied in the low temperature region. Indeed in the harmonic approximation the Dulong Petit law would be exact if we neglect the $`\gamma `$ dependence of $`S_\gamma (\beta )`$. The value of $`\gamma (\beta )`$ weakly depends on $`\beta `$: its value in the limit $`\beta \mathrm{}`$ is only about 10% higher that its value (i.e. $`\beta _K`$) at the transition temperature. The results presented here are an explicit realization of a transition driven by an entropy crisis, which has been firstly implemented microscopically in the random energy model . The specific heat has a jump downward, when we decrease the temperature, which is the opposite of the typical behaviour in transitions characterized by the onset of a conventional order parameter (in mean field theory ferromagnets, superconductors, have a jump upward when we decrease the temperature). We stress that it is possible to remove the approximation of using the harmonic entropy. There are no serious difficulties in computing numerically the true entropy of the valleys; this has been done for a binary mixture of soft sphere and the results are very near to the harmonic ones . Analytically we can also expand in the anharmonicity parameters. It is also possible to take care of the fluctuations of the entropy from valley to valley and go beyond the quenched approximation. It is quite likely that these effects will not strongly change the results. The most important step would be to reach a better theoretical control on the spectrum of the INN in the liquid phase, maybe combining the approximations used here with the low density expansion of . Summarizing we have found a method which is able to use liquid theory method in the glasses phase putting in practice the old adage a glass is a frozen liquid. We are able to compute with a reasonable approximation the thermodynamics and with a little more effort we can compute the static and the dynamic structure functions. We owe a lot to M.Mézard and we are very happy to thanks him. We are also very grateful to W.Kob and F.Sciortino for their suggestions and to A.Cavagna and I.Giardina for useful discussions. B.C. would like to thank the Physics Department of Rome University ‘La Sapienza‘ where this work was partially developed during her PhD.
no-problem/9906/cond-mat9906339.html
ar5iv
text
# Untitled Document Poisson Bracket Formulation of Nematic Polymer Dynamics Randall D. Kamien email: kamien@physics.upenn.edu Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA 19104 We formulate the dynamical theory of nematic polymers, starting from a microscopic Poisson bracket approach. We find that the Poisson bracket between the nematic director and momentum depends on the (Maier-Saupe) order parameter of the nematic phase. We use this to derive reactive couplings of the nematic director to the strain rates. Additionally, we find that local dynamics breaks down as the polymers begin to overlap. We offer a physical picture for both results. 21 June 1999; revised 29 June 1999 1. Introduction and Summary The dynamics of line-like objects, from biomolecules to flux-lines in superconductors, has become increasingly important to understand. The processing and manufacture of petroleum products, food products and petroleum-food products requires the control of polymer flow. The remarkable properties of superconductors can be exploited only if flux-flow dynamics is controlled. While the equilibrium statistical mechanics of line-like objects such as directed polymers, flux lines and dipole chains in ferro- and electrorheological fluids has been formulated from the microscopic physics\[1,,2\], a similar theory has not been developed for the dynamics. While polymer entanglement dramatically affects the dynamics by introducing cosmologically long time scales, the statics is not affected. Indeed, it is the dynamics of entanglement that aids in pinning of high-temperature superconductors, in the strength of glue and in other remarkable rheological properties. Previously, dynamics have been formulated on a hydrodynamic basis, which ignores the connectivity of the lines . Other approaches include effective theories which use single-polymer response functions and neglect polymer-polymer interactions . Here, we study the dynamics of polymers which, due to steric or other interactions, are already aligned into a nematic state. By considering a system which is orientationally ordered in equilibrium, the effect of flow on alignment can be separated from the effect of alignment on flow. In this work we formulate the fluctuating hydrodynamics of polymer nematics based on a microscopic, Poisson bracket approach . In this way we will account for polymer degrees of freedom as well as long-lived hydrodynamic variables. Our principle result is a fundamentally new derivation of the elusive reactive coupling $`\lambda `$ between the director field $`𝐧`$ and the velocity field $`𝐯`$: $$\frac{n_\mu }{t}=\left[\delta _{\mu \nu }n_\mu n_\nu \right]\left\{\frac{1+\lambda }{2}n_\gamma _\gamma v_\nu \frac{1\lambda }{2}n_\gamma _\nu v_\gamma \right\}.$$ Unlike most parameters in hydrodynamics $`\lambda `$ is neither set by Kubo formulae nor the fluctuation-dissipation theorem. We will show that to the order we work, our result agrees with the analysis of Archer and Larson which is based upon a phenomenological, Landau-theory dynamics . Thus our result both verifies the validity of both a microscopic and a phenomenological approach. As is typical of the former approach, we develop our theory in steps. First we identify the hydrodynamic variables. In a polymer nematic system we have four fields of interest: the areal polymer density $`\rho (𝐱)`$, the momentum density $`𝐠(𝐱)`$, the fluctuation of the nematic director $`\delta \stackrel{}{n}(𝐱)=𝐧(𝐱)𝐧_0`$, and the density of polymer “heads” and “tails”, $`\rho _{\mathrm{HT}}(𝐱)`$. We then write these fields in terms of the positions and momenta of the individual monomers. We consider, as an example, the areal density field $`\rho (𝐱)`$. Since the polymers are directed we may write the location of the $`\alpha ^{\mathrm{th}}`$ polymer in terms of an affine parameter $`\tau _\alpha `$ which marks the monomers: $$𝐑_\alpha (\tau _\alpha )=[r_\alpha ^x(z)+\tau _\alpha h_\alpha ^x,r_\alpha ^y(z)+\tau _\alpha h_\alpha ^y,\tau _\alpha +s_\alpha ]$$ where $`\stackrel{}{r}_\alpha (z)`$ is the (two-dimensional) displacement of the monomer at height $`z=\tau _\alpha +s_\alpha `$ away from the ground state, straight-line configuration with the two-dimensional tangent $`\stackrel{}{h}_\alpha `$, and the polymers start at a height $`z=s_\alpha `$. The expression for the areal density is: $$\rho (𝐱)=\underset{\alpha }{}\delta ^2\left[\stackrel{}{r}_\alpha (z)+(zs_\alpha )\stackrel{}{h}_\alpha 𝐱_{}\right]\mathrm{\Theta }\left[zs_\alpha \right]\mathrm{\Theta }\left[e_\alpha z\right],$$ where $`e_\alpha `$ is the height of the polymer end. This expression requires further explanation and, at the same time, shows the inherent complexity of this problem. The areal density is a sum of delta-functions in the plane at the positions of the polymers. Moreover, since these polymers have heads and tails the product of Heaviside step functions $`\mathrm{\Theta }()`$ counts the polymer only if we consider heights between the start and end of the polymer at $`z=s_\alpha `$ and $`z=e_\alpha `$ respectively. We will investigate similar expressions for the remaining fields in the next section. We note here that although expressions such as (1.1) are awkward, similar expressions could not easily be formed in isotropic polymer systems. The essential feature of polymer nematics is that chain contour length corresponds to the common $`z`$-axis along which all the polymers are more or less aligned. This allows us to formulate our theory in space, so as to incorporate hydrodynamics, while at the same time including the microscopic polymer degrees of freedom. These degrees of freedom are accounted for via the Poisson brackets of the $`\stackrel{}{r}_\alpha (z)`$ with the conjugate momentum $`\stackrel{}{p}_\alpha (z)`$. From these, we can generate the Poisson brackets of the hydrodynamic fields: if we can get a closed set of brackets then we can construct a consistent hydrodynamic theory. However, we see that the areal density (1.1) not only depends on $`\stackrel{}{r}_\alpha (z)`$ but also on the location of the polymer heads and tails as well as the average polymer tangent $`\stackrel{}{h}_\alpha `$. The head and tail locations are accounted for via the field $`\rho _{\mathrm{HT}}(𝐱)`$ described above. What do we do with the average polymer tangent $`\stackrel{}{h}_\alpha `$? Though the nematic phase has two Nambu-Goldstone modes associated with spontaneously-broken rotational invariance, there is, in addition, a massive mode, the Maier-Saupe order parameter $`S`$ which measures the degree of nematic ordering. The $`\stackrel{}{h}_\alpha `$ are associated with this mode. The order parameter can change locally via changes in the local polymer alignment. Presumably, if the polymers are entangled, the time scale for these rearrangements corresponds to some sort of reptation time required for the release of topological constraints. On the other hand, if the polymers are short (unentangled) then there is a separation of time scales which allows us to approximately decouple the local Maier-Saupe mode. In both these limits, we can decouple these tilt modes from the monomer modes. This will allow us to pre-average the tilt modes to arrive at effective Poisson brackets for the remaining fields. The next step in our formulation of dynamics requires a coarse-grained free energy in terms of our macroscopic variables. In section III we will propose a free energy consistent with the symmetries of the system which also includes the equilibrium physics of the polymer ends. From this we will derive dynamical equations for the hydrodynamic variables. 2. Definition of Hydrodynamic Variables We start by formulating the theory in terms of microscopic, directed trajectories of the individual polymers, $`\{\stackrel{}{r}_\alpha (z)\}`$, where $`\alpha `$ labels the polymer and $`\stackrel{}{r}(z)`$ is the displacement of the polymer from its equilibrium position. Recall that the nematic state is characterized by the Maier-Saupe order parameter, $$S=\frac{3\mathrm{cos}^2\theta 1}{2}$$ where $``$ denotes a thermal, ensemble average, and $`\theta `$ is the angle between the nematic director and the ordering axis (taken throughout to be $`\widehat{z}`$). When $`S1`$, the directors do not all line up along $`\widehat{z}`$. Since the nematic order parameter is set by the details of the isotropic–nematic transition, it is a massive, non-hydrodynamic mode. In order to take into account the ground state value of $`S`$ we will assume that the equilibrium polymer trajectory is $`𝐑_\alpha (z)=[(zs_\alpha )h_\alpha ^x,(zs_\alpha )h_\alpha ^y,z]`$ and that $`\stackrel{}{r}_\alpha (z)=0`$. Thus, the microscopic director is $`d\stackrel{}{r}/dz+\stackrel{}{h}`$. Finally, the $`\mu `$ component of the momentum of polymer $`\alpha `$ at $`z`$ is $`p_\alpha ^\mu (z)`$. Unlike previous formulations , we do not consider the traceless symmetric tensor $`Q_{ij}`$ typically used for nematics. Rather, deep in the nematic state, we are only concentrating on the two Goldstone modes of the broken rotational symmetry. As we shall see this leads to a closed set of Poisson brackets to lowest order in derivatives and the director fluctuation $`\delta \stackrel{}{n}`$. We will incorporate the “up-down” symmetry of nematics via the symmetry of the free energy, just as is done in the usual equilibrium description of nematics in terms of the unit-vector director field $`\widehat{𝐧}`$. 2.1. Canonical Variables and Separation of Time Scales While our description of the chain locations is adequate in terms of $`\stackrel{}{r}_\alpha (z)`$ and $`\stackrel{}{h}_\alpha `$, we should take care in finding correct canonical variables from which to construct macroscopic Poisson brackets. We imagine starting with a Hamiltonian which is a function the actual polymer location $`𝐑_\alpha (\tau _\alpha )`$, as in (1.1), and the conjugate momentum $`𝐩_\alpha (\tau _\alpha )`$. The equations of motion in the $`xy`$-plane are $$\dot{p}_\alpha ^i(\tau _\alpha )=\frac{\delta H}{\delta R_\alpha ^i(\tau _\alpha )}\dot{R}^i(\tau _\alpha )=\frac{\delta H}{\delta p_\alpha ^i(\tau _\alpha )}$$ If $`\stackrel{}{h}_\alpha `$ is constant then these equations of motion are precisely $$\begin{array}{cc}\hfill \dot{p}_\alpha ^i(z)& =\{p_\alpha ^i(z),H\}\hfill \\ \hfill \dot{r}_\alpha ^i(z)& =\{r_\alpha ^i(z),H\}\hfill \end{array}$$ with $$\{r_\alpha ^i(z),r_\beta ^j(z^{})\}=\{p_\alpha ^i(z),p_\beta ^j(z^{})\}=0,$$ and $$\{r_\alpha ^i(z),p_\beta ^j(z^{})\}=\delta _{\alpha \beta }\delta ^{ij}\delta (zz^{}),$$ where $`i`$ and $`j`$ run over the $`xy`$-plane. We have made the identification $`𝐩_\alpha (z)=𝐩_\alpha (\tau _\alpha )`$ with $`z=\tau _\alpha +s_\alpha `$. We must also calculate the Poisson bracket of $`p_\alpha ^z(z)`$ with the spatial coördinate. Motion along the $`z`$-axis should be accounted for in motion of the polymer ends. Indeed, motion of the monomers up or down depends only on the motion of the starting point $`s_\alpha `$. Thus $$\{r_\alpha ^i(z),p_\beta ^z(z^{})\}=0$$ but $$\{z_\alpha ,p_\beta ^z(z^{})\}=\delta _{\alpha \beta }\delta (z_\alpha z^{}),$$ where we use the symbol $`z_\alpha `$ to remind the reader that the commutator does not vanish only for momenta and positions on the same polymer. In other words, $`z_\alpha `$ takes the place of the $`z`$-component of $`\stackrel{}{r}_i(z)`$ and $`z_\alpha =s_\alpha +\tau _\alpha `$. We remind the reader that the fact that the affine parameter $`\tau `$ can be mapped one-to-one with the coördinate height $`z`$ is precisely the simplification which we are exploiting. Of course, when we consider hydrodynamic variables, $`z_\alpha `$ will just become the height coördinate $`z`$. We must now justify our neglect of the dynamics of $`\stackrel{}{h}_\alpha `$. As we discussed in the introduction, we can do this in two limits. One limit is when the chains are long and highly-entangled. In this case we expect that rearrangements of the nematic texture require topological constraint release or, in other words, reptation. We are concentrating here on dynamics on time scales shorter than the reptation time. We could extend this work by including some sort of reptation-based dynamics for the tilt fields $`\stackrel{}{h}_\alpha `$, thus incorporating both hydrodynamics and topological constraints in the same theory. There is another limit of interest. If the polymers are short, we expect that the modes associated with the Maier-Saupe order parameter will be the slowest to relax to an equilibrium distribution. In this case, we expect that on the intermediate time scales we are focusing on, the distribution of $`\stackrel{}{h}_\alpha `$ will also be static, but consistent with the background nematic order. In this case the tilt modes can be averaged over, thus explicitly removing the dependence of the fields on them. We contrast this average to a thermodynamic average: in this case, the average is over an ensemble of different frozen-in textures while in thermodynamics we take the time average of a single molecule. In both these cases we get the same mean-square average of the tilt field, though the correlations will differ. Thus, in both the very long polymer and short polymer case we can pre-average the tilt degrees of freedom. 2.2. Macroscopic Fields We must first define the coarse-grained, macroscopic fields, namely the mass density, the director and the momentum density. The mass density is given by: $$\rho (𝐱)=\underset{\alpha }{}\delta ^2\left(\stackrel{}{r}_\alpha (z)+(zs_\alpha )\stackrel{}{h}_\alpha 𝐱_{}\right)\mathrm{\Theta }\left[zs_\alpha \right]\mathrm{\Theta }\left[e_\alpha z\right]$$ $`(2.1a)`$ As described in the introduction, the terms in the sum come from the following considerations: 1) The Dirac delta function “counts” the areal mass (note that it is a two-dimensional delta function, defined in the $`xy`$ plane). 2) The pair of Heaviside step functions ($`\mathrm{\Theta }[]`$) control the polymer lengths. The $`\alpha ^{\mathrm{th}}`$ polymer begins at $`z=s_\alpha `$ and ends at $`z=e_\alpha `$. The remaining macroscopic fields are defined via $$\begin{array}{ccc}\hfill [\rho \delta n^i](𝐱)& =\underset{\alpha }{}\left(\frac{dr_\alpha ^i(z)}{dz}+h_\alpha ^i\right)\delta ^2\left(\stackrel{}{r}_\alpha (z)+(zs_\alpha )\stackrel{}{h}_\alpha \stackrel{}{x}_{}\right)\frac{\mathrm{\Theta }\left[zs_\alpha \right]\mathrm{\Theta }\left[e_\alpha z\right]}{\sqrt{1+\stackrel{}{h}_\alpha ^2}}\hfill & (2.1b)\hfill \\ \hfill g^\mu (𝐱)& =\underset{\alpha }{}p_\alpha ^\mu (z)\delta ^2\left(\stackrel{}{r}_\alpha (z)+\stackrel{}{h}_\alpha (zs_\alpha )\stackrel{}{x}_{}\right)\mathrm{\Theta }\left[zs_\alpha \right]\mathrm{\Theta }\left[e_\alpha z\right]\hfill & (2.1c)\hfill \end{array}$$ Note that an additional factor of $`\left(1+\stackrel{}{h}_\alpha ^2\right)^{1/2}`$ arises in the definition of $`\delta \stackrel{}{n}(𝐱)`$ since the normalized unit (average) tangent vector is $$𝐓_\alpha =\frac{[h_\alpha ^x,h_\alpha ^y,1]}{\sqrt{1+\stackrel{}{h}^2}}.$$ Note that the naïve definitions of the macroscopic fields, e.g. $$\rho (𝐱)=\underset{\alpha }{}\delta ^2[\stackrel{}{r}_\alpha \stackrel{}{x}_{}]\mathrm{\Theta }\left[zs_\alpha \right]\mathrm{\Theta }\left[e_\alpha z\right],$$ applies only when $`\stackrel{}{h}_\alpha \stackrel{}{0}`$, or, in other words, when the nematic ground state has Maier-Saupe order parameter, $`S=1`$. 3. Pre-Averaging the Nematic Texture As we discussed in the introduction, the Poisson brackets of the hydrodynamic fields cannot close since they all depend on the average polymer tilt $`\stackrel{}{h}_\alpha `$. As discussed in the preceding, we would like to pre-average over these degrees of freedom to arrive at Poisson brackets for the hydrodynamic fields. This procedure can be justified in two limiting cases. Consider the dynamics of a single polymer of length $`L`$. This can be decomposed into normal modes ala Rouse or Zimm . In either case the $`n^{\mathrm{th}}`$ mode has wavenumber $`q_n=2\pi n/L`$ with a relaxation frequency $`\omega _nn^2/L^2`$. For very short polymers the separation in time scales between the lowest mode $`n=1`$ and the other modes is large. In this limit we can decouple the longest mode – the tilt mode – characterized by $`\stackrel{}{h}`$. We could thus average the tilts over the collection of polymers. The mean-square of the tilt fields is simply related to the Maier-Saupe order parameter $`S`$. We make the distinction between the time average of a single polymer’s tilt field which gives equilibrium statistics and the quenched average over the frozen polymer tilts. As we imagine lengthening the polymers, the splitting in timescales becomes less pronounced and their spectrum becomes continuous. However, when the polymer is sufficiently long, a new dynamics dominates the longest wavenumber modes: reptation dynamics. In this regime, the tilt modes can only relax via activated reptation which is an especially apt description of the likely rearrangements in polymer nematics. Thus, when the polymer is long enough to be entangled we also have a large separation of time scales. As a result, in this limit, we may also treat the polymer tilt field as essentially frozen. We have now two distinct averages: the first is an average over a massive degree of freedom that sets the value of $`S`$. In the present formulation, this average is over $`\stackrel{}{h}_\alpha `$. The second average is the thermal average over fluctuations about the ground state, i.e. averages over $`\stackrel{}{r}_\alpha (z)`$. We will treat these two averages separately, which amounts to treating the average over $`\stackrel{}{h}_\alpha `$ as a quenched average. To connect with the Maier-Saupe order parameter, we have $$S=\frac{3}{2}\overline{\left(\frac{1}{1+\stackrel{}{h}_\alpha ^2}\right)}\frac{1}{2}$$ since $`\mathrm{cos}\theta `$ in (2.1) is just the $`z`$ component of the average tangent vector and where we have denoted the quenched average of $`X`$ by $`\overline{X}`$. Expanding for small $`|\stackrel{}{h}_\alpha |`$, we have $`\frac{2}{3}(1S)\overline{\stackrel{}{h}_\alpha ^2}`$. We will then take averages over $`\stackrel{}{h}_\alpha `$ weighted by $$P(\stackrel{}{h}_\alpha )=\frac{1}{2\pi \mathrm{\Delta }}e^{|\stackrel{}{h}|^2/2\mathrm{\Delta }}$$ where $`\mathrm{\Delta }=\frac{1}{3}(1S)`$ so that (3.1) is satisfied. In order to find the average macroscopic field variables, we must regularize the Dirac delta functions appearing in $`(2.1)`$ . To do this we represent them as Gaussians with the widths taken to be the excluded area of each polymer in a fixed $`z`$-plane. We define $$\delta ^2(\stackrel{}{r};a)=\frac{1}{2\pi a}e^{\left|\stackrel{}{r}\right|^2/2a}.$$ Then, when we average over $`\stackrel{}{h}_\alpha `$ weighting by (3.1) we find $$\overline{\delta ^2[\stackrel{}{r}_\alpha (z)+(zs_\alpha )\stackrel{}{h}_\alpha 𝐱_{};a]}=\delta ^2[\stackrel{}{r}_\alpha (z)𝐱_{};a+\mathrm{\Delta }(zs_\alpha )^2].$$ We can thus calculate the tilt-averaged fields. Since we have only included terms for the tangent vector in $`(2.1b)`$ to leading order in $`\stackrel{}{h}_\alpha `$, we calculate average field values to order $`\mathrm{\Delta }`$: $$\begin{array}{ccc}\hfill \overline{\rho }(𝐱)& =\underset{\alpha }{}\delta ^2(\stackrel{}{r}_\alpha (z)𝐱_{};a+\mathrm{\Delta }(zs_\alpha )^2)\mathrm{\Xi }_\alpha (z)\hfill & (3.1a)\hfill \\ \hfill [\overline{\rho }\overline{\delta }n^i](𝐱)& =\underset{\alpha }{}\left(1\mathrm{\Delta }\right)\frac{dr_\alpha ^i(z)}{dz}\delta ^2(\stackrel{}{r}_\alpha (z)\stackrel{}{x}_{};a+\mathrm{\Delta }(zs_\alpha )^2)\mathrm{\Xi }_\alpha (z)\hfill & \\ & \underset{\alpha }{}(zs_\alpha )\mathrm{\Delta }_i\delta ^2(\stackrel{}{r}_\alpha (z)\stackrel{}{x}_{};a+\mathrm{\Delta }(zs_\alpha )^2)\mathrm{\Xi }_\alpha (z)\hfill & (3.1b)\hfill \\ \hfill \overline{g}^\mu (𝐱)& =\underset{\alpha }{}p_\alpha ^\mu (z)\delta ^2(\stackrel{}{r}_\alpha (z)\stackrel{}{x}_{};a+\mathrm{\Delta }(zs_\alpha )^2)\mathrm{\Xi }_\alpha (z)\hfill & (3.1c)\hfill \end{array}$$ where $`\mathrm{\Xi }_\alpha (z)\mathrm{\Theta }\left[zs_\alpha \right]\mathrm{\Theta }\left[e_\alpha z\right]`$. 3.1. The Breakdown of Local Dynamics Typically, Poisson brackets for dynamical variables are of the form $$\{\rho (x),\stackrel{}{g}(x^{})\}=\stackrel{}{}_x\left[\rho (x)\delta ^3(xx^{})\right].$$ The presence of the delta-function in (3.2) along with a local free energy density leads to local dynamics , i.e. every term in the dynamical equations is evaluated at the same point in space and time. Such a result must, of course, be taken with a grain of salt. The delta function is an idealization to a system composed of point-like constituents. In reality, these delta functions should be replaced with a smeared-out distribution. Usually, this presents no problem – calculations may be done with the smeared-out distribution and the limit may be taken at the end. However, this is not the case in this situation. From (3.1), it is easy to see that we may only replace the disorder smeared delta-functions with the original distributions (of width $`a`$) when $`a\mathrm{\Delta }(zs_\alpha )^2`$. In other words, it is only when the original width is wider than the disorder-averaged width that the distributions are delta-function like. Since the presence of the term $`\mathrm{\Theta }\left[zs_\alpha \right]\mathrm{\Theta }\left[e_\alpha z\right]`$ in the expressions for the hydrodynamic variables limits the maximum value of $`z`$, we have $`\mathrm{\Delta }(zs_\alpha )^2<\mathrm{\Delta }(e_\alpha s_\alpha )^2=\mathrm{\Delta }\mathrm{}^2`$, where $`\mathrm{}`$ is the typical polymer length. Thus we see that the original distributions are recovered everywhere whenever $`\mathrm{}\sqrt{\overline{h^2}}\mathrm{}\sqrt{\mathrm{\Delta }}a`$. This is easy to interpret; when the average wandering of the polymer away from its starting point, $`\mathrm{}\sqrt{\mathrm{\Delta }}`$ is smaller than the interpolymer spacing, we may continue to treat the polymers and their interactions as local. Once the polymers have tipped out of their cages the dynamics will become, necessarily, non-local. A force at $`𝐱`$ leads to a reaction at $`𝐱^{}`$ if a single polymer can go from $`𝐱`$ to $`𝐱^{}`$. Presumably, the typical polymer length $`\mathrm{}`$ should be replaced by $`\mathrm{}_P`$, the polymer persistence length, when $`\mathrm{}>\mathrm{}_P`$ and the locality condition becomes $`\mathrm{\Delta }\mathrm{}_P^2a^2`$. In this delocalized regime the formalism will break down. In particular, the Poisson brackets of the coarse-grained fields, calculated in terms of (2.1) and (2.1) will be straightforward but complicated, and will necessarily lead to non-local dynamics, even if the free-energy density arises from local interactions. In the Conclusion we will discuss possible alternatives to a non-local formalism based on different sets of variables. 4. Fluctuating Hydrodynamics We can, however, consider the case of $`\mathrm{\Delta }1`$. This could arise in the description of flux lines in superconductors or polymer nematics in applied fields where the ground state polymer configurations are almost always parallel to $`\widehat{z}`$. In any of these limits, it is appropriate to replace the spread out delta functions $`\delta ^2(;a+\mathrm{\Delta }(zs)^2)`$ with point-like delta functions. In this case, we can calculate the Poisson brackets of the average fields in terms of the canonical, microscopic brackets. We find the following, non-zero brackets, to lowest order in derivatives (of both $`\stackrel{}{r}_\alpha (z)`$ and delta functions) and $`\delta n`$: $$\begin{array}{ccc}\hfill \{\overline{\rho }(𝐱),\overline{g}^\mu (𝐱^{})\}& =\left[\delta _\nu ^\mu \delta _z^\mu \delta _\nu ^z\right]_x^\nu \left[\rho (𝐱)\delta ^3(\stackrel{}{x}\stackrel{}{x}^{})\right]\hfill & \\ & +\left(1+\mathrm{\Delta }\right)\delta _z^\mu \left[\rho _{\mathrm{HT}}(𝐱)\delta ^3(\stackrel{}{x}\stackrel{}{x}^{})_i\left(\rho \delta n^i(𝐱)\delta ^3(\stackrel{}{x}\stackrel{}{x}^{})\right)\right]\hfill & (4.1a)\hfill \\ \hfill \{\overline{\delta }n^i(𝐱),\overline{g}^\mu (𝐱^{})\}& =\left[\delta _\nu ^\mu \delta _z^\mu \delta _\nu ^z\right]\delta ^3(\stackrel{}{x}\stackrel{}{x}^{})_x^\nu \delta n^i(𝐱)\hfill & \\ & +\left[\left(1\mathrm{\Delta }\right)\delta ^{i\mu }_x^z\mathrm{\Delta }\delta ^{\mu z}_x^i\right]\delta ^3(\stackrel{}{x}\stackrel{}{x}^{})\hfill & (4.1b)\hfill \\ \hfill \{\overline{g}^\mu (𝐱),\overline{g}^\nu (𝐱^{})\}& =_x^\nu \left[g^\mu (𝐱)\delta ^3(\stackrel{}{x}\stackrel{}{x}^{})\right]+_x^{}^\mu \left[g^\nu (𝐱^{})\delta ^3(\stackrel{}{x}\stackrel{}{x}^{})\right]\hfill & (4.1c)\hfill \end{array}$$ We have been forced to introduce an additional hydrodynamic field, $`\rho _{\mathrm{HT}}(𝐱)`$, the density of polymer heads and tails, defined as $$\rho _{\mathrm{HT}}(𝐱)\underset{\alpha }{}\left(1\mathrm{\Delta }\right)\left\{\delta ^2(\stackrel{}{r}_\alpha (s_\alpha )\stackrel{}{x}_{})\delta (zs_\alpha )\delta ^2(\stackrel{}{r}_\alpha (e_\alpha )\stackrel{}{x}_{})\delta (e_\alpha z)\right\}$$ Note that the second term in $`(4.1a)`$ is exactly equal to $`_z\overline{\rho }(𝐱)`$. In other words, we have $$\left(1\mathrm{\Delta }\right)_z\overline{\rho }(𝐱)+_i\left[\overline{\rho }\delta \overline{n}_i(𝐱)\right]=\rho _{\mathrm{HT}}(𝐱)$$ This constraint is the familiar constraint of line liquids \[12,,13\] which conserves polymer number but for polymer ends. The prefactor of $`(1\mathrm{\Delta })`$ comes from the average over the $`z`$ direction. In other words, the rotationally invariant conservation law is $`\stackrel{}{n}_0\rho +\stackrel{}{n}=\rho _{\mathrm{HT}}`$. When averaging over the intrinsic randomness of $`\stackrel{}{n}_0`$, we get a factor of $`(1\mathrm{\Delta })`$ in front of $`_z\rho `$. It is not present in other terms because they are higher order in derivatives and powers of $`\delta \stackrel{}{n}`$. 4.1. Equations of Motion With these Poisson brackets we can systematically derive the reactive terms of hydrodynamics. The viscosities will appear in accord with the symmetries of the system. Being uniaxial there will be $`5`$ viscosities , characterized by the viscous stress tensor $`V_{\alpha \beta }=\eta _{\alpha \beta \gamma \beta }A_{\gamma \beta }`$ where $$\begin{array}{cc}\hfill \eta _{\mu \nu \gamma \beta }& =\eta _1\left[n_\mu n_\mu n_\gamma n_\beta \right]+\eta _2\left[\delta _{\mu \gamma }^{}\delta _{\nu \beta }^{}+\delta _{\mu \beta }^{}\delta _{\nu \gamma }^{}\delta _{\mu \nu }^{}\delta _{\gamma \beta }^{}\right]+\eta _4\left[\delta _{\mu \nu }^{}\delta _{\gamma \beta }^{}\right]\hfill \\ & +\eta _3\left[n_\mu n_\gamma \delta _{\nu \beta }^{}+n_\nu n_\gamma \delta _{\mu \beta }^{}+n_\mu n_\beta \delta _{\nu \gamma }^{}+n_\nu n_\gamma \delta _{\mu \beta }^{}\right]+\eta _5\left[\delta _{\mu \nu }^{}n_\gamma n_\beta +n_\mu n_\nu \delta _{\gamma \beta }^{}\right]\hfill \end{array}$$ where $`\delta _{\mu \nu }^{}=\delta _{\mu \nu }n_\mu n_\nu `$. The form of $`V`$ is fixed by symmetries: by nematic inversion it must be even in $`\stackrel{}{n}`$ and by the Kubo formulae it must be symmetric under $`\mu \nu `$, $`\gamma \beta `$ and $`(\mu \nu )(\gamma \beta )`$, and $`A_{\gamma \beta }=\frac{1}{2}\left(_\gamma v_\beta +_\beta v_\gamma \right)`$ is the strain rate tensor constructed out of the fluid velocity $`\stackrel{}{v}`$. Expanding $`\stackrel{}{n}=\widehat{z}+\delta \stackrel{}{n}`$ we can find the viscosity tensor in the nematic phase. The free energy from which we will derive the reactive terms is: $$F^{}=d^3x\left\{\frac{\stackrel{}{g}^2}{2\rho }+F_{\mathrm{equil}}\right\}$$ where the equilibrium free energy is the sum of terms: $$F_{\mathrm{equil}}=F_{\delta \stackrel{}{n}}+F_{\mathrm{pol}}$$ and $$F_{\delta \stackrel{}{n}}=d^3x\left\{\frac{K_1}{2}\left(\delta \stackrel{}{n}\right)^2+\frac{K_2}{2}\left(\times \delta \stackrel{}{n}\right)^2+\frac{K_3}{2}\left(_z\delta \stackrel{}{n}\right)^2\right\}$$ is the usual Frank free energy for a nematic, while $$F_{\mathrm{pol}}=d^3x\left\{\frac{B}{2}\delta \rho ^2+\frac{G}{2}\left[_z\delta \rho +\rho _0\delta \stackrel{}{n}\right]^2\right\}$$ is the free energy of the polymer . In (4.2) $`B`$ is the two-dimensional bulk modulus, $`\rho _0=\rho \delta \rho `$ is the average, areal polymer density and $`G`$ is the fugacity for hairpins and free ends. We have replaced $`\rho _{\mathrm{HT}}`$ with the expression in (4.2), to lowest order in the field fluctuations $`\delta \rho `$ and $`\delta \stackrel{}{n}`$ (which is why we drop the factor of $`\left(1\mathrm{\Delta }\right)`$ in front of $`_z\delta \rho `$). The parameter $`G`$ is given by $`G=k_\mathrm{B}T\mathrm{}/2\rho _0`$ where $`\mathrm{}`$ is the typical polymer length . Terms of the form $`\delta \rho \delta \stackrel{}{n}`$ are not allowed due to the $`\stackrel{}{n}\stackrel{}{n}`$ symmetry of the nematic phase. We are now able to derive the equations of motion for $`\stackrel{}{g}`$, $`\delta \stackrel{}{n}`$ and $`\delta \rho `$ (where we have dropped the bar over the variables denoting the $`\stackrel{}{h}`$ average). They are (with $`\stackrel{}{v}=\stackrel{}{g}/\rho `$): $$\begin{array}{ccc}\hfill _t\rho & =_ig_i+\left(1+\mathrm{\Delta }\right)\rho _{\mathrm{HT}}v_z_i\left(\delta n_ig_z\right)\hfill & (4.2a)\hfill \\ \hfill _t\delta n_i& =_j\left(v_j\delta n_i\right)+\left(1\mathrm{\Delta }\right)_zv_i\mathrm{\Delta }_iv_z\mathrm{\Gamma }\frac{\delta F[\delta n_i]}{\delta (\delta n_i)}+\theta _i\hfill & (4.2b)\hfill \\ \hfill _tg_\mu & =_\nu \left(\frac{g_\mu g_\nu }{\rho }\right)+\left[\delta _\nu ^\mu \delta _z^\mu \delta _\nu ^z\right]\rho _\nu \left(\frac{\delta F[\rho ]}{\delta \rho }\right)+_\alpha V_{\alpha \mu }+\xi _\mu \hfill & (4.2c)\hfill \end{array}$$ where, in order to approach thermodynamic equilibrium, the noise terms $`\xi _i(\stackrel{}{x},t)`$ and $`\theta _i(\stackrel{}{x},t)`$ have the following correlations : $$\begin{array}{ccc}\hfill \xi _\mu (\stackrel{}{x},t)\xi _\beta (\stackrel{}{x}^{},t^{})& =2k_\mathrm{B}T\eta _{\mu \nu \gamma \beta }_\nu _\gamma \delta ^3(\stackrel{}{x}\stackrel{}{x}^{})\delta (tt^{})\hfill & (4.3a)\hfill \\ \hfill \theta _\mu (\stackrel{}{x},t)\theta _\nu (\stackrel{}{x}^{}t^{})& =2k_\mathrm{B}T\mathrm{\Gamma }\delta _{\mu \nu }^{}\delta ^3(\stackrel{}{x}\stackrel{}{x}^{})\delta (tt^{})\hfill & (4.3b)\hfill \\ \hfill \xi _\mu (\stackrel{}{x},t)\theta _\nu (\stackrel{}{x}^{},t^{})& =0\hfill & (4.3c)\hfill \end{array}$$ There are a number of features of these equations worth mention. First, we see that the typical polymer lengths come into the equations through the coupling $`G`$, present in the equations for $`_t\delta n_i`$ and $`_tg_\mu `$. When $`G`$ is large, corresponding to long polymers, we may take the coefficient, $`\rho _{\mathrm{HT}}=0`$. In the short molecule limit, presumably $`G0`$. There is a crossover wavevector $`k_c=\sqrt{2B\rho _0^3/(k_BT\mathrm{})}`$ below which the system behaves as a pure nematic and above which the length of the polymers becomes important. There is a corresponding timescale $`\omega _ck_c^21/\mathrm{}`$ which delineates a similar short-to-long crossover. In the short-polymer limit, care must be taken due to the nature of the approximation under which the Poisson brackets were derived. In particular, since we treat the polymer tangent field as a vector, we are, in principle, distinguishing $`\stackrel{}{n}`$ and $`\stackrel{}{n}`$. However, for small fluctuations around equilibrium, long polymer nematics should not sample the entire space of fluctuations: it is very unlikely that an entire polymer will rotate by $`\pi `$ around the $`x`$\- or $`y`$-axis. This simplification is what enabled us to get a closed set of Poisson brackets order by order in a derivative expansion. The effect of hairpins could be put in explicitly through the introduction of a hairpin-density field . 4.2. Reactive Couplings in the Nematodynamic Limit Finally, we note that $`(4.2b)`$ gives an expression for the elusive reactive parameter $`\lambda `$ in nematodynamics. By rotational invariance, $`\stackrel{}{n}\stackrel{}{n}`$ and the constraint $`\stackrel{}{n}^2=1`$, the reactive part of the equation of motion for $`\stackrel{}{n}`$ has the form: $$\frac{dn_\mu }{dt}=\delta _{\mu \nu }^{}\left[\frac{1+\lambda }{2}n_\gamma _\gamma v_\nu \frac{1\lambda }{2}n_\gamma _\nu v_\gamma \right]$$ We can thus identify $$\lambda =12\mathrm{\Delta }=\frac{1+2S}{3}.$$ This is in general agreement with the results of Forster who calculated the value of $`\lambda `$ within a Poisson bracket formalism for $`Q_{\mu \nu }`$ the symmetric, traceless order parameter for short molecule nematic liquid crystals . He found that $$\lambda =\frac{1+2\alpha S}{3},$$ where $`\alpha =(I_l+2I_t)/(I_lI_t)`$ is a parameter depending on the moments of inertia $`I_l`$ and $`I_t`$ of the nematogens parallel and perpendicular to the nematic axis, respectively. We thus recover Forster’s result when $`\alpha =1`$ or, in other words, when the aspect ratio $`I_l/I_t`$ of the nematogens becomes infinite. Presumably this is a consequence of taking delta-function densities in the transverse plane. It is interesting to compare this result to the work of Archer and Larson . In the same limit of infinite aspect ratio, they found an expression for $`\lambda `$ in terms of the expectations of the second- and fourth-rank order parameters, $`P_2S`$ and $`P_4`$ where $`P_2`$ and $`P_4`$ are the second and fourth Legendre polynomials evaluated at $`x=\mathrm{cos}\theta `$: $$\lambda =\frac{15P_2+48P_4+42}{105P_2}.$$ To be consistent, we should expand (4.4) in powers of $`\mathrm{\Delta }=(1S)/3`$ to compare with (4.4) and compare linear terms. First, to linear order in $`\mathrm{\Delta }`$, $$\mathrm{cos}^4\theta =12\stackrel{}{h}_\alpha ^2\left(1\stackrel{}{h}_\alpha ^2\right)^2\mathrm{cos}^2\theta ^2$$ and so $`P_4(35S^210S7)/18`$ which gives $$\lambda =1+\frac{2}{3}(S1)+\frac{2}{9}(S1)^2+\mathrm{}.$$ Thus to leading order in $`(S1)=3\mathrm{\Delta }`$ (4.4) agrees with (4.4)! Thus within the $`S1`$ limit our result should be consistent with the data as in . We note, moreover, that since the isotropic-to-nematic phase transition is first order, $`S`$ does not grow continuously from $`0`$. Indeed, in Maier-Saupe theory $`S0.44`$ at this transition , and thus $`\mathrm{\Delta }0.2`$. Therefore the deviation between our result and the more exact result (4.4) should be small sufficiently well aligned samples. However, although (4.4) may be quantitatively reasonable, it misses an essential qualitative feature: it is always less than $`1`$ and thus predicts that nematics will always tumble. Though the linear result in $`\mathrm{\Delta }`$ fails to predict the crossover from flow aligning to tumbling behavior, the virtue of our derivation of $`\lambda `$ is that we get a direct interpretation of its origin. In a highly aligned sample with $`S=1`$ only gradients of $`v_{}`$ along the $`z`$-direction can lead to rotations of the molecule (see Figure 1). On the other hand, when $`S<1`$, gradients in the $`x`$-direction of $`v_z`$ can also rotate the molecules and so the polymer nematics would always be in a tumbling mode. A higher order analysis in powers of $`\mathrm{\Delta }`$ would be required to see if a $`\lambda >1`$ could come from our direct approach. 5. Conclusions In summary we have derived the Poisson brackets of the relevant degrees of freedom for a polymer nematic. We have shown that these canonical brackets become highly non-local when the polymers start to overlap. In the limit where the polymers do not overlap, we have presented a microscopic derivation of the reactive coupling $`\lambda `$, a coupling which is not a long-wavelength limit of a correlation function. Our analysis requires a “locality condition” in order to be trustworthy. While this may appear restrictive, one might imagine that, at some level of coarse graining the polymers, the no overlap condition might be met. If we were to clump polymer regions correlated in the $`xy`$-plane together and coarse grain to the scale of this “entanglement correlation length” $`\xi _e`$, our analysis may be applicable. Our analysis is especially applicable to flux lines in superconductors. There, $`\rho _{\mathrm{HT}}0`$ and there is no problem, even in principle, to considering the flux line tangents as vectors. The flux lines, in addition, have $`\mathrm{\Delta }=0`$, simplifying the theory further. It is perhaps in this context that one could hope to make the most theoretical progress. This derivation has shown the inherent, unavoidable non-locality in polymer nematic dynamics. An interesting possibility is to focus on a different set of conserved or almost conserved variables. For instance, it may be possible to reformulate this dynamics in terms of an entanglement density by calculating the Hopf density , a scalar which measures the local curvature of the director configuration – for long polymers curvature is a measure of local entanglement. This density could be used as a starting point for a phenomenological theory of “entanglement dynamics.” Another variable of interest might be the repton density, which measures length per unit length . It may be possible to develop a hydrodynamics for this field as well . 6. Acknowledgments It is a pleasure to acknowledge stimulating discussions with D. Forster, E. Frey, T. Lubensky, S.T. Milner, D. Morse, R. Pelcovits, T. Powers, H. Stark and J. Toner. This work was supported by NSF MRSEC Grant DMR96-32598, the Research Corporation, the Donors of the Petroleum Research Fund, administered by the American Chemical Society and the Alfred P. Sloan Foundation. References relax J.V. Selinger and R.F. Bruinsma, Phys. Rev. A 43, 2910 (1991); Phys. Rev. A 43, 2922 (1991). relax P. Le Doussal and D.R. Nelson, Europhys. Lett. 15, 161 (1991); R.D. Kamien, P. Le Doussal, and D.R. Nelson, Phys. Rev. A 45, 8727 (1992). relax L. Radzihovsky and E. Frey, Phys. Rev. B 48, 10357 (1993). relax D.C. Morse, Phys. Rev. E 58, 1237 (1998); Macromolecules 31, 7030 (1998); 31, 7044 (1998). relax D. Forster, Hydrodynamic Fluctuations, Broken Symmetry, and Correlation Functions (W.A. Benjamin, Reading, MA, 1975). relax D. Forster, Phys. Rev. Lett. 32 (1974) 1161. relax L.A. Archer and R.G. Larson, J. Chem. Phys. 103, 3108 (1995). relax M. Doi and S.F. Edwards, The Theory of Polymer Dynamics, (Oxford University Press, Oxford, 1986). relax T.C. Lubensky, Phys. Rev. A 2 (1970) 2497. relax A.N. Semenov and M. Rubinstein, Eur. Phys. B 1, 87 (1998). relax P.M. Chaikin and T.C. Lubensky, Principles of Condensed Matter Physics, (Cambridge University Press, Cambridge, 1995). relax V.G. Taratura and R.B. Meyer, Liquid Crystals 2 373 (1987). relax R.D. Kamien and D.R. Nelson, J. Stat. Phys. 71, 23 (1993). relax P.G. de Gennes and J. Prost, The Physics of Liquid Crystals, Second Edition, Chap. VII (Oxford University Press, New York, 1993). relax J. Prost and N.A. Clark, in Proceedings of the International Conference on Liquid Crystals, Bangalore, 1979, edited by S. Chandrasekhar (Heyden, Philadelphia, 1980); in Liquid Crystals of One and Two-Dimensional Order, Garmisch-Partenkirchen, Fanvar, 1980 edited by W. Helfrich and F. Hepke (Springer, Berlin, 1980). relax R.D. Kamien, Eur. Phys. J. B 1, 1 (1998). relax P.-G. de Gennes, J. Chem. Phys. 55, 572 (1971). relax R.D. Kamien, unpublished (1999). Figure Captions relaxrelaxFig. 1. Velocity fields and nematic directors with various gradients and molecular orientations. a) and c) Gradient in the $`z`$-direction of $`v_{}`$. This will rotate the director for all values of $`S`$. b) and d) Gradient in the $`x`$-direction of $`v_z`$. Only in d) will this rotate the molecule, i.e. when $`S<1`$.
no-problem/9906/cond-mat9906276.html
ar5iv
text
# 1 Hall conductivity 𝜎_{𝑥⁢𝑦} normalized to its zero-temperature value as a function of temperature normalized to the transition temperature of FISDW. The solid and dashed lines represent the temperature dependences of the condensate density in the dynamic and static limits, respectively, calculated from Eqs. () and (). The left panel shows experimental points for (TMTSF)2PF6 under pressure 8.5 kbar for the 𝑁=1 Hall plateau at 𝐻=15.5 T assuming 𝑇_𝑐=2.52 K, and the right panel for 𝑁=2 at 𝐻=13.25 T assuming 𝑇_𝑐=2.1 K. Temperature evolution of the quantum Hall effect in the FISDW state: Theory vs Experiment Victor M. Yakovenko and Hsi-Sheng Goan Department of Physics, University of Maryland, College Park, MD 20742-4111, USA Jonghwa Eom and Woowon Kang James Franck Institute and Department of Physics, University of Chicago, Chicago, IL 60637, USA Abstract. We discuss the temperature dependence of the Hall conductivity $`\sigma _{xy}`$ in the magnetic-field-induced spin-density-wave (FISDW) state of the quasi-one-dimensional Bechgaard salts (TMTSF)<sub>2</sub>X. Electronic thermal excitations across the FISDW energy gap progressively destroy the quantum Hall effect, so $`\sigma _{xy}(T)`$ interpolates between the quantized value at zero temperature and zero value at the transition temperature $`T_c`$, where FISDW disappears. This temperature dependence is similar to that of the superfluid density in the BCS theory of superconductivity. More precisely, it is the same as the temperature dependence of the Fröhlich condensate density of a regular CDW/SDW. This suggests a two-fluid picture of the quantum Hall effect, where the Hall conductivity of the condensate is quantized, but the condensate fraction of the total electron density decreases with increasing temperature. The theory appears to agree with the experimental results obtained by measuring all three components of the resistivity tensor simultaneously on a (TMTSF)<sub>2</sub>PF<sub>6</sub> sample and then reconstructing the conductivity tensor. In a magnetic field $`H`$, quasi-one-dimensional organic conductors of the (TMTSF)<sub>2</sub>X family experience a cascade of magnetic-field-induced spin-density-wave (FISDW) phase transitions (see review ). At zero temperature, each FISDW phase exhibits an integer quantum Hall effect (see review ). As temperature $`T`$ increases, the Hall effect decreases and virtually vanishes at $`T_c`$, the transition temperature of FISDW. (In the normal state without FISDW, the Hall effect is very small and can be approximated as zero compared to the Hall effect in the FISDW state.) It was shown in Refs. that the temperature evolution of the Hall conductivity (per one layer) is given by the following formula: $$\sigma _{xy}(T)=f(T)\frac{2Ne^2}{h},$$ (1) where $`e`$ is the electron charge, $`h`$ is the Planck constant, $`N`$ is an integer number that characterizes the FISDW, and $`f(T)`$ is the dimensionless condensate density of FISDW. The condensate density $`f(T)`$ interpolates between 1 at zero temperature (all electrons are in the condensate) and 0 at $`T_c`$ (the condensate density vanishes when FISDW disappears). An explicit expression for $`f(T)`$ depends on the order in which the limits of zero frequency $`\omega 0`$ and zero momentum $`q0`$ are taken . In the dynamic limit, where the $`q0`$ is taken first and then $`\omega 0`$ is taken, the condensate density is $$f_d(T)=1_{\mathrm{}}^{\mathrm{}}\frac{dp_x}{v_\mathrm{F}}\left(\frac{E}{p_x}\right)^2\left[\frac{n_\mathrm{F}(E)}{E}\right],$$ (2) where $`E=\sqrt{(v_\mathrm{F}p_x)^2+\mathrm{\Delta }^2}`$ is the electron energy dispersion law in the FISDW state ($`v_F`$ is the Fermi velocity, $`p_x`$ is the electron momentum along the chains, and $`\mathrm{\Delta }`$ is the energy gap), and $`n_\mathrm{F}(ϵ)=(e^{ϵ/k_\mathrm{B}T}+1)^1`$ is the Fermi distribution function ($`k_\mathrm{B}`$ is the Boltzmann constant). In the static limit ($`\omega 0`$ first, then $`q0`$), the condensate density is $$f_s(T)=1v_\mathrm{F}_{\mathrm{}}^{\mathrm{}}𝑑p_x\left[\frac{n_\mathrm{F}(E)}{E}\right].$$ (3) The static limit (3) is appropriate for calculation of the magnetic field penetration depth in superconductors, which comes from the truly static Meissner effect in thermodynamic equilibrium. On the other hand, the Hall effect is kinetic and not thermodynamic, so we believe that the dynamic limit (2) is appropriate for Eq. (1). In Fig. 1, we compare theory and experiment. Hall conductivity $`\sigma _{xy}`$ normalized to its zero-temperature value is shown as a function of temperature normalized to the transition temperature of FISDW. The solid and dashed lines represent the temperature dependences of the condensate density in the dynamic and static limits, respectively, calculated from Eqs. (2) and (3) presuming that the temperature dependence of the energy gap $`\mathrm{\Delta }`$ is the same as in the BCS theory of superconductivity . The points are obtained experimentally by measuring the three components of the resistivity tensor, $`\rho _{xx}`$, $`\rho _{yy}`$, and $`\rho _{xy}`$, simultaneously on a sample of (TMTSF)<sub>2</sub>PF<sub>6</sub> under pressure 8.5 kbar and then calculating the conductivity tensor. The left panel shows the experimental points for the $`N=1`$ Hall plateau at $`H=15.5`$ T assuming $`T_c=2.52`$ K, and the right panel for $`N=2`$ at $`H=13.25`$ T assuming $`T_c=2.1`$ K. The left panel demonstrates a good agreement with the dynamic limit. The right panel seems to agree with the static limit. However, the right panel has only four experimental points, and the discrimination between the static and dynamic fits is effectively controlled by only one point at about $`T/T_c=0.6`$. So, we believe that there is not enough data to make a conclusion for the $`N=2`$ plateau, but for the $`N=1`$ plateau the agreement with the dynamic limit appears to be convincing. Nevertheless, more data is necessary to make a firmer conclusion. We also wish to point out that Eq. (1) provides a unique opportunity to determine the condensate density of FISDW by measuring the Hall conductivity. This is a linear-response measurement, uncomplicated by depinning, phase slips, and other nonlinear effects, which characterize sliding of a regular charge- or spin-density wave. The last terms in Eqs. (2) and (3) reflects the fact that normal quasiparticles, thermally excited above the energy gap, reduce the quantum Hall effect. We can write the condensate density in the form $$f(T)=1\rho _n/\rho ,$$ (4) where the second term is the density of the normal component, $`\rho _n`$, normalized to the total electron density $`\rho `$. Thus, $`f(T)`$ is analogous to the superfluid density in superconductors. Given that normal electrons exhibit virtually zero Hall effect, Eq. (1) can be interpreted in a two-fluid manner: The condensate carries the dissipationless quantum Hall current, and normal electrons carry the longitudinal, dissipative electric current. The Hall response of the condensate remains quantized even at a finite temperature, but the Hall conductivity decreases because the condensate density decreases with increasing temperature. In this picture, the quantum Hall current in the FISDW state is analogous to the supercurrent in a superconductor, which is also carried only by a condensate, whose density decreases with increasing temperature. This analogy continues an amazing sets of parallels between the quantum Hall effect in the FISDW state and superconductivity. The thermodynamics of both states is described by the BCS theory of a mean-field energy gap $`\mathrm{\Delta }`$ opening at the Fermi surface . Both states are characterized by dissipationless currents: the quantum Hall current and the supercurrent. At a finite temperature, both systems have two-fluid electrodynamics. The temperature dependence discussed in this paper was obtained under assumption that the quantum Hall effect originates from the bulk of the crystal. The results have implications for the edge states picture of the quantum Hall effect. Theory says that in the FISDW state there must be $`N`$ gapless chiral electron states at the sample edge . At zero temperature, the quantized Hall conductivity can be equivalently obtained in terms of either bulk or edge states . However, it does not seem possible to describe the Hall conductivity at a finite temperature in terms of only the edge states. Indeed, the edge states are gapless, so they cannot produce Eqs. (2) and (3), which involve the gap $`\mathrm{\Delta }`$. Moreover, because of thermal excitations across the energy gap in the bulk, it does not seem possible to ignore the bulk contribution compared to the edge contribution at a finite temperature. By any means, at $`TT_c`$ distinction between the bulk and edge states must go away. So, the bulk theory of the quantum Hall effect appears to be more general and robust than the edge one. It is worth mentioning that the question where the quantum Hall current actually flows, in the bulk or in the edges, was never studied experimentally in quasi-one-dimensional conductors and does not have a clear answer even in the case of the quantum Hall effect in semiconductors.
no-problem/9906/hep-ex9906006.html
ar5iv
text
# A Polarized HERA Collider ## 1 Introduction It is often said that physics studies involving spin can lead to surprising results. We list some outstanding examples below. Some Surprises With Spin 1. Space quantization associated with quantized spin directions. Stern, Gerlach, 1921. 2. Atomic fine structure and electron spin magnetic moment. Goudsmit, Uhlenbeck, 1926. 3. Proton anomalous magnetic moment; $`\mu _p=2.79\mathrm{nm}`$. Stern, 1933. 4. Electron spin anomalous magnetic moment. $`\mu _e=\mu _0`$ (1.00119); QED. Kusch, 1947. 5. Electroweak interference from $`\stackrel{}{e}_1d`$ DIS parity nonconservation. Prescott & SLAC-Yale Collaboration, 1978. 6. Proton spin structure; puzzle or crisis. EMC, 1989. Hence from an historical viewpoint the spin variable is a promising one for discovery. The first experiment, beginning in the mid-1970’s, on polarized lepton-proton deep inelastic scattering was a series of two measurements at SLAC by a SLAC-Yale group, using an atomic beam polarized electron source built at Yale and an electron beam energy of $``$20 GeV. The virtual photon-proton asymmetry $`A_1^\mathrm{p}`$ was measured in the $`x`$ range from 0.1 to 0.7 and found to be large. The values were consistent with a plausible quark-parton model satisfying the Bjorken sum rule, and also satisfying the Ellis-Jaffe sum rule if Regge extrapolation of $`g_1^\mathrm{p}(x)`$ to $`x`$ = 0 was employed . The European Muon Collaboration (EMC) in the mid-1980’s made a similar measurement of $`A_1^\mathrm{p}`$ using a polarized $`\mu ^+`$ beam with energy up to 200 GeV. In addition to confirming the SLAC results for $`x>0.1`$, their data extended down to $`x=0.01`$. Unexpectedly, the $`A_1^\mathrm{p}`$ data at low $`x`$ fell well below the extrapolated SLAC data. The consequence was violation of the Ellis-Jaffe sum rule at the 3 standard deviation level, and the conclusions were that only the small fraction $`\mathrm{\Delta }=0.12\pm 0.17`$ of the proton spin is due to quark spins and that strange quarks have a substantial negative polarization . This discovery by EMC has led to major new experiments at CERN, SLAC and DESY which were reviewed by R. Windmolders in this workshop. Some outstanding questions relevant to spin structure remain: 1) Behavior of $`g_1^\mathrm{p}(x)`$ at small $`x`$ and the value of the first moment $`\mathrm{\Gamma }_1^\mathrm{p}`$=$`_0^1𝑑x`$$`g_1^\mathrm{p}(x)`$. 2) Contribution of gluons to proton spin and the polarized gluon distribution. 3) Contribution of the orbital angular momentum of quarks and gluons to proton spin. 4) Hadronic spin structure of the photon. 5) Chiral structure of any observed contact interaction or leptoquark beyond the standard model. Experiments such as polarized lepton-nucleon scattering which require a polarized beam and a polarized target are sometimes spoken of as spin physics and are sometimes regarded as not of central interest. However, it is well known that the hadronic tensor $`W^{\mu \nu }`$, which describes the proton in DIS $`ep`$ or $`\mu p`$ scattering, involves four scalar functions: $`F_1(x)`$, $`F_2(x)`$, $`g_1(x)`$ and $`g_2(x)`$, which are required for a complete knowledge of $`W^{\mu \nu }`$. The functions $`F_1`$ and $`F_2`$ do not involve the spin variables, whereas $`g_1`$ and $`g_2`$ do. Perhaps the principal achievements of HERA to date are the measurement of $`F_2(x)`$ at small $`x`$ and also at high $`Q^2`$, the determination of the unpolarized gluon distribution in the proton, study of the hadronic constituents of the photon, and extension of the limits on a contact interaction or leptoquark search. A polarized HERA collider will address these same topics from a different viewpoint. Spin is a fascinating tool but it is not the goal of these experiments. ## 2 Characteristics of a Polarized HERA Collider There has been considerable interest in a possible polarized HERA collider with the characteristics indicated in Tab. 1, in several workshops associated with HERA. At present of course there is a polarized electron beam in HERA which is used in the HERMES experiment. Development of a polarized proton beam is required. Then one or both of the existing collider detectors –ZEUS,H1– could be used for measurements of polarized $`e^\pm p`$ DIS. The huge increase in the $`xQ^2`$ range for measurements of spin variables possible with a polarized HERA collider is shown in Fig.1. Two orders of magnitude increase in both $`x`$ and $`Q^2`$ range is possible for exploring the spin structure of the proton. If a sufficiently intense source of <sup>3</sup>He<sup>-</sup> can be developed, measurements of the spin structure of the neutron could be done. ## 3 Particle & Nuclear Physics with Polarized HERA ### 3.1 Measurement of $`g_1^\mathrm{p}`$ at low $`x`$ The behavior of $`g_1^\mathrm{p}`$ at low $`x`$ is of fundamental interest and the largest uncertainty on the first moment of $`g_1^\mathrm{p}`$ now comes from the unmeasured low $`x`$ region, $`x<0.003`$ . The statistical uncertainties for a measurement of $`g_1^\mathrm{p}`$ with a polarized HERA collider (Tab. 1), using the H1 or ZEUS detector, are shown in Fig. 2. Even though the predicted asymmetries are as small as $`3\times 10^4`$ at low $`x`$, false asymmetries associated with correlations of proton beam intensity or bunch crossing angle with proton polarization, or with time variation of detection efficiencies can be kept still smaller by modulation of the spin direction. From the polarized H<sup>-</sup> source any desired polarization can be provided for a proton bunch so that rapid modulation for successive interactions is achieved. Also in the HERA ring at high energy, all of the spin directions of the proton bunches can be flipped periodically by a microwave pulse. Recently it has been checked using a fast detector simulation that the detector smearing and event migration does not hinder the measurement $`g_1`$. Also shown in Fig.2 are the various low $`x`$ predictions for the structure function arising from models and QCD fits consistent with the presently available data. ### 3.2 The Polarized Gluon Distribution Determination of the polarized gluon distribution $`\mathrm{\Delta }G(x,Q^2)`$ inside a nucleon and its first moment $`\mathrm{\Delta }G`$ have become important goals of all experiments proposed and planned in the next decade. Polarized HERA can contribute significantly, through the various different and independent ways in which it can measure $`\mathrm{\Delta }G(x,Q^2)`$. Perturbative QCD analysis of $`g_1^\mathrm{p}`$ allows a determination of $`\mathrm{\Delta }G(x,Q^2)`$ through a next-to-leading order analysis. The published value of $`\mathrm{\Delta }G=1.0_{0.3}^{+1.2}(\mathrm{stat})_{0.2}^{+0.4}(\mathrm{syst})_{0.5}^{+1.4}(\mathrm{theo})`$ indicates that statistical and theoretical uncertainties dominate our lack of knowledge of $`\mathrm{\Delta }G(x,Q^2)`$. A study using simulated HERA data and the presently available fixed target data indicates that the statistical & theoretical uncertainties could be reduced by factors $`2`$ & $`3`$, respectively. In the Photon Gluon Fusion (PGF) process the gluon is involved at leading order (LO). The measurement of such a process through detection of 2 high–p<sub>T</sub> jets and the scattered electron allows access to the polarized gluon distribution. Figure 3 shows the statistical uncertainty achieved by polarized HERA collider (Tab. 1). Also shown are three widely different predictions for $`\mathrm{\Delta }G(x,Q^2)`$ at LO, all consistent with the fixed target data, which HERA data can easily distinguish between. Expected uncertainty on $`\mathrm{\Delta }G/G`$ after HERA measurements of 2-jets is $`\pm 0.1`$. It has also been checked that the detector smearing and migration effects due to the measurement process do not affect the measurability of $`\mathrm{\Delta }G`$. Recently, NLO corrections to this process were evaluated and found to be small. Both H1 and ZEUS collaborations have published results on the parton distributions inside the unpolarized photon $`q^\gamma `$. With polarized HERA one could investigate the structure of the polarized photon $`\mathrm{\Delta }q^\gamma `$. A study, using single and 2 high-p<sub>T</sub> jets or hadron tracks from the PGF process in photoproduction, showed that a luminosity of 100 pb<sup>-1</sup> was sufficient to resolve the polarized photon and the gluon structure. ## 4 Polarized HERA: Accelerator Aspects Production of a high energy polarized proton beam in HERA requires first a polarized H<sup>-</sup> source and then acceleration to high energy through the DESY acceleration chain with retention of polarization (Fig. 4). The full initial study of the requirements for a high energy polarized beam in HERA identified two crucial problems. One is the development of an adequate polarized H<sup>-</sup> source, and the second is the acceleration and retention of proton polarization in HERA. ### 4.1 Polarized Source for HERA The design specifications for a polarized H<sup>-</sup> source for HERA are: I=20 mA, in 100$`\mu s`$ pulses at 0.25 Hz with emittance $`2\pi `$ mm$``$mr, and $`P_p=0.8`$. At present the most promising approach is the Optically Pumped Polarized Ion Source (OPPIS) development at TRIUMF by A. Zelenski et al. . The overall source arrangement is shown in Fig. 5. A polarized H<sup>-</sup> source is now being developed at TRIUMF for the Relativistic Heavy Ion Collider (RHIC) at BNL for their RHIC-Spin program. Development of an OPPIS source suitable for HERA is progressing with optimism. ### 4.2 Acceleration of Polarized Protons The acceleration and then storage of polarized protons from the ion source to the high energy HERA ring is a major problem dominated principally by depolarizing resonances. Because of the large anomalous magnetic moment of the proton ($`\mu _\mathrm{p}=2.79\mathrm{nm}`$), the relativistic equations of spin motion show that at high energy the number of spin precessions per orbit –the spin tune– is large, indeed equal to $`G\gamma `$ which is $``$1530 at E=800 GeV. Hence the spin motion is very sensitive to the magnetic field, and in particular the many intrinsic and imperfection resonances can lead to depolarization. Avoidance of depolarization involves the use of the Siberian Snake principle. Extensive simulation calculations by spin tracking codes have been done. Figure 6 shows for two energies the equilibrium polarization distribution or the spin vector for the stored proton beam in HERA, using 8 snakes. Although such results are encouraging, much further intensive study of spin motion is required to assure adequate proton polarization at high energy. Electron cooling in the DESYIII ring is being considered to reduce the beam emittance and thus make it simpler to obtain high proton polarization. ### 4.3 Proton Beam Polarimetry Measurement of proton beam polarization at high energy is under active development at BNL, mainly for the RHIC Spin program in which polarized proton beams up to 250 GeV will collide. Three types of polarimeter are presently considered: 1) Inclusive pion production $`\stackrel{}{p}+C\pi ^++X`$, 2) $`\stackrel{}{p}+C`$ elastic scattering in the Coulomb Nuclear Interference (CNI) region, capable of $`8\%`$ absolute accuracy determined by the theoretical uncertainty of the hadronic spin flip amplitude, and 3) $`\stackrel{}{p}+p`$ elastic scattering involving a polarized jet target. This method equates the asymmetry $`A`$ in the elastic scattering $`(p_\mathrm{b}+\stackrel{}{p}_{\mathrm{jet}})`$, for which the polarization of the jet $`P_{\mathrm{jet}}`$ is known, to the beam polarization $`P_\mathrm{b}`$ for $`(\stackrel{}{p}_\mathrm{b}+p_{\mathrm{jet}})`$ scattering for the same scattering kinematics. It is capable of $`3\%`$ absolute accuracy. ## 5 Summary Important physics results can be expected from $`\stackrel{}{e}+\stackrel{}{p}`$ collisions with the polarized HERA collider. However, increased effort is needed now to solve the challenging accelerator physics problems associated with achieving high energy polarized protons in the HERA ring.
no-problem/9906/hep-ph9906267.html
ar5iv
text
# The 𝑎₀⁢(980), 𝑎₀⁢(1450) and 𝐾₀^∗⁢(1430) Scalar Decay Constants and the Isovector Scalar Spectrum ## I Introduction Despite considerable recent theoretical activity, there exists, at present, no concensus on the nature of the $`f_0(980)`$ and $`a_0(980)`$ mesons. Interpretations still advocated in the literature include (1) the $`q^2\overline{q}^2`$ (“four-quark”) cryptoexotic interpretation originally proposed by Jaffe (see Refs. ); (2) the $`K\overline{K}`$ bound state, or “molecule”, picture; (3) the unitarized quark model picture; and (4) the “minion” picture of Gribov. The $`K\overline{K}`$ molecule assignment for the $`f_0(980)`$ and $`a_0(980)`$ is attractive (and, perhaps for this reason, often cited as already established) because it naturally explains the proximity to $`K\overline{K}`$ threshold and strong $`K\overline{K}`$ couplings of the states. Such bound states occur naturally in the quark model treatment of Ref. , as well as several coupled channel treatments using model meson-meson interactions with parameters fit to experimental data. The picture predicts $`\gamma \gamma `$ decay widths compatible with experiment, and considerably smaller than those expected, at least in the simplest version of the quark model, for $`{}_{}{}^{3}P_{0}^{}`$ $`q\overline{q}`$ mesons. One should bear in mind, however, that an alternate quark model approach, whose $`\gamma \gamma `$ model dynamics are constrained by an analogous treatment of $`\pi ^0,\eta \gamma \gamma `$, allows widths compatible with experiment for a conventional quark model assignment, and that small $`\gamma \gamma `$ widths are also claimed in the four-quark picture, based on a schematic estimate loosely motivated by the MIT bag model. An alternate means of testing the molecule scenario is via the processes $`\varphi f_0\gamma ,a_0\gamma `$. Recent experimental determinations give branching ratios larger than predicted in the molecule picture but, since the predictions of Refs. and differ by a factor of 4, realistic theoretical uncertainties may be large, making the discrepancy with experiment difficult to interpret reliably. One of the problems with the interpretation of the light scalar states is that their narrow experimental widths, which appear unnatural for a conventional $`q\overline{q}`$ meson assignment (but natural in the $`K\overline{K}`$ molecule picture), may not, in fact, reflect the true intrinsic widths, since strong coupling of an intrinsically broad state to a nearby $`s`$-wave threshold can produce significant narrowing through the effects of channel coupling and unitarity. Given the empirically observed strong $`f_0(980)`$ and $`a_0(980)`$ couplings to $`K\overline{K}`$, the possibility arises that these states are, not molecules, but rather “unitarized quark model” (UQM) states, i.e. conventional $`q\overline{q}`$ mesons with properties strongly distorted by coupling to the nearby $`K\overline{K}`$ threshold. In such a scenario, one expects significant $`K\overline{K}`$ content in the scalar states, and hence most likely a significant reduction in the $`\gamma \gamma `$ widths (though these widths have yet to be calculated in the UQM picture). The distortion caused by the strong coupling will presumably also alter expectations for $`\varphi f_0\gamma ,a_0\gamma `$, though one again awaits an explicit calculation of this effect. A number of observations made in the context of the UQM, and discussions surrounding it, bear repeating at this point. First, in the presence of strong coupling to nearby decay channels, a single underlying state can manifest itself as more than just a single pole close to the physical region (Törnqvist and Roos, for example, claim that both the $`a_0(980)`$ and $`a_0(1450)`$ are likely to be manifestations of the same underlying $`q\overline{q}`$ state). Second, existing experimental data are apparently insufficient to distinguish between the molecule and UQM pictures, since coupled channel models exist, both with and without an underlying conventional quark model state, which provide good fits to existing data in the coupled $`\pi \pi `$, $`K\overline{K}`$ and $`\pi \eta `$, $`K\overline{K}`$ channels, once the data has been used to fit the free parameters of the models. Third, although the method for distinguishing conventional resonances from bound states posited by Morgan and Pennington (a conventional resonance having poles near the physical region on both the second and third sheets, a bound state having such a pole on only the second sheet) is frequently borne out in existing coupled channel calculations, this is not universally the case. In particular, a situation in which there are two nearby poles, one on the second and one on the third sheet, can arise in a model which provides a good fit to existing data, but for which one of the nearby poles is most naturally thought of as the coupled-channel remnant of a $`K\overline{K}`$ bound state (in the sense that, as the channel coupling is dialed up towards its final fitted value, what was a $`K\overline{K}`$ bound state in the absence of channel coupling moves continuously to become one of the two final nearby poles). Thus, even if data could distinguish between the one-nearby-pole and two-nearby-pole scenarios (see Refs. for arguments that this may not be possible at present), it appears that it would not allow us to distinguish between the UQM and molecule scenarios. In this paper, we discuss the feasibility of using meson decay constants describing the coupling of the $`I=1/2`$ and $`I=1`$ scalar mesons to (pointlike) flavor-non-diagonal scalar densities, as a means of further clarifying the nature of the isovector scalar spectrum. The basic idea is the same as that underlying attempts to make use of the processes $`f_0,a_0\gamma \gamma `$ and $`\varphi f_0\gamma ,a_0\gamma `$, namely that the spatial extent of a loosely bound $`K\overline{K}`$ molecule is significantly larger than that of a conventional $`q\overline{q}`$ meson, which is, in turn, significantly larger than that of the very compact Gribov minions. The scalar densities, being pointlike, provide a direct probe of such scale differences, and one which is, moreover, less prone to obscuration by intervening dynamics. In other channels, analogous decay constants provide useful information on the classification of states. For example, the fact that, using recent ALEPH data, one has $$g_K^{}=1.1g_\rho ^{}g_\rho ^{},$$ (1) (where $`g_{V^{ab}}`$, with $`a,b=u,d,s`$, is defined by $`0|J_\mu ^{ab}|V^{ab}g_{V^{ab}}ϵ_\mu `$, with $`J_\mu ^{ab}=\overline{q_a}\gamma _\mu q_b`$, $`V^{ab}`$ the corresponding vector meson, and $`ϵ_\mu `$ the vector meson polarization vector) supports the assignment of the $`\rho `$ and $`K^{}`$ to the same $`SU(3)_F`$ multiplet. The same-multiplet assignment of the $`\pi `$ and $`K`$ is similarly supported by the result $$f_K=1.2f_\pi f_\pi ,$$ (2) while the fact that $$g_\omega ^{EM}/g_{\rho ^0}^{EM}1/3,$$ (3) (where $`g_\omega ^{EM}`$, $`g_{\rho ^0}^{EM}`$ are the electromagnetic decay constants of the $`\omega `$ and $`\rho ^0`$), rather than the value $`1/\sqrt{3}`$ (expected if the $`\omega `$ is the flavor $`8`$ member of a vector meson octet), gives direct evidence for ideal mixing in the vector meson nonet. Expectations for the relative sizes of the scalar decay constants in the various scenarios are obvious: they should be small for a weakly bound $`K\overline{K}`$ molecule, large for the very compact ($`.1.2\mathrm{fm}`$) Gribov minions and, for conventional $`q\overline{q}`$ states, should obey approximate $`SU(3)_F`$ relations among the the decay constants of different members of the same multiplet. In the four-quark picture, since, in the bag model, a four-quark state is larger than a two-quark state and, in addition, the hidden strange pair would have to be annihilated, one would also expect the decay constant to be suppressed. We will show below that experimental data allows us to determine the scalar decay constant of the $`K_0^{}(1430)`$, while a QCD sum rule analysis fixes the analogous $`a_0(980)`$ and $`a_0(1450)`$ decay constants. We will then show that (1) the weakly bound $`K\overline{K}`$ molecule and minion scenarios are ruled out by the sum rule analysis and (2) that the relations between the various decay constants do not support the assignment of either $`a_0`$ resonance as the $`SU(3)_F`$ partner of the $`K_0^{}(1430)`$, but rather suggest a UQM-like scenario. ## II The $`K_0^{}(1430)`$ Scalar Decay Constant We define the decay constants of interest to us in this paper as follows: $`0|^\mu J_\mu ^{su}|K_0^{}(1430)`$ $``$ $`f_{K_0^{}(1430)}m_{K_0^{}(1430)}^2`$ (4) $`0|^\mu J_\mu ^{du}|a_0`$ $``$ $`f_{a_0}^{}m_{a_0}^2`$ (5) where, in Eq. (5), $`a_0`$ refers to either of the two $`a_0`$ resonances. In QCD, one has $$^\mu J_\mu ^{ab}=i\left(m_am_b\right)S^{ab},$$ (6) with $`S^{ab}=\overline{q_a}q_b`$. Since the scalar densities, $`S^{ab}`$, are members of an $`SU(3)_F`$ octet, $`SU(3)_F`$ relations are simplified by redefining the $`a_0`$ decay constants in such a way that a common mass factor occurs on the LHS’s of Eqs. (4) and (5), i.e., $$\left(\frac{m_sm_u}{m_dm_u}\right)0|^\mu J_\mu ^{du}|a_0f_{a_0}m_{a_0}^2.$$ (7) For an $`a_0`$ state lying in the same $`SU(3)_F`$ multiplet as the $`K_0^{}(1430)`$, one should then have $$f_{a_0}m_{a_0}^2f_{K_0^{}(1430)}m_{K_0^{}(1430)}^2,$$ (8) while, in the $`K\overline{K}`$ molecule/four-quark/minion pictures, the LHS of Eq. (8) should be, respectively, much smaller/smaller/larger than the RHS. To make use of the expectation provided by Eq. (8), one first requires the decay constant of the “reference” quark model state, the $`K_0^{}(1430)`$. This may be obtained from the spectral function, $`\rho _s(s)`$, of the correlator, $`\mathrm{\Pi }_s(q^2)`$, defined by (with $`J_s=^\mu J_\mu ^{su}`$), $$\mathrm{\Pi }_s(q^2)=id^4xe^{iqx}0|T\left(J_s(x)J_s^{}(0)\right)|0,$$ (9) since one has, at the $`K_0^{}(1430)`$ peak, neglecting background, $$\frac{f_{K_0^{}(1430)}^2m_{K_0^{}(1430)}^3}{\pi \mathrm{\Gamma }_{K_0^{}(1430)}}=\rho _s(m_{K_0^{}(1430)}^2).$$ (10) $`\rho _s(s)`$ may, in turn, be constructed from experimental $`K\pi `$ phase shifts and the value of the timelike scalar $`K\pi `$ form factor, $`d(s)`$, at some convenient kinematic point since (1) unitarity relates the $`K\pi `$ component of the spectral function to $`d(s)`$ via $$\left[\rho _s(s)\right]_{(K\pi )}=\frac{3}{32\pi ^2}\sqrt{\frac{\left(ss_+\right)\left(ss_{}\right)}{s^2}}|d(s)|,$$ (11) (where $`s_\pm =\left(m_K\pm m_\pi \right)^2`$) and, (2) $`d(s)`$ satisfies an Omnes relation, $$d(s)=d(0)exp\left[\frac{s}{\pi }_{th}^{\mathrm{}}𝑑s^{}\frac{\delta (s^{})}{s^{}(s^{}siϵ)}\right],$$ (12) where $`\delta (s)`$ is the phase of $`d(s)`$, and the normalization $`d(0)`$ is known from ChPT and $`K_{e3}`$. (In writing this equation, we have, as elsewhere in the literature, ignored a possible polynomial pre-factor; see Ref. for an empirical justification.) Experimentally, $`K\pi `$ scattering is known to be purely elastic up to $`s2.5\mathrm{GeV}^2`$. Thus, $`\delta (s)`$ is identical to the $`I=1/2`$ $`K\pi `$ phase shift up to this point. At the edge of the experimental region, $`s=(1.7\mathrm{GeV})^2`$, moreover, the measured $`K\pi `$ phase has essentially reached the asymptotic value, $`\pi `$, required by quark counting rules for $`d(s)`$, allowing one to rather safely assume $`\delta (s)=\pi `$ for $`s>(1.7\mathrm{GeV})^2`$. With these assumptions, $`\left[\rho _s(s)\right]_{K\pi }`$ is determined by Eqs. (11) and (12). The self-consistency of these assumptions has been checked by the sum rule analysis of Ref. . Finally, the $`K_0^{}(1430)`$ $`K\pi `$ branching fraction is known to be compatible with $`100\%`$, so the $`K\pi `$ component represents the full spectral function in the $`K_0^{}(1430)`$ region. (Note that Particle Data Group values for the mass and width reflect mis-transcriptions in Ref. ; the corrected values are $`m_{K_0^{}(1430)}=1.412\mathrm{GeV}`$ and $`\mathrm{\Gamma }_{K_0^{}(1430)}=0.294\mathrm{GeV}`$. The corresponding values of the effective range parameters appearing in the LASS parametrization of the $`K\pi `$ phase are $`a=2.19\mathrm{GeV}^1`$ and $`b=3.74\mathrm{GeV}^1`$.) Combining experimental data and Eqs. (10), (11) and (12), one obtains $$f_{K_0^{}(1430)}m_{K_0^{}(1430)}^2=.0842\pm .0045\mathrm{GeV}^3.$$ (13) The errors in Eq. (13) reflect the range of values obtained when the $`K_0^{}(1430)`$ resonance parameters are varied within the errors quoted for the phase shift fit of Ref. , and also the difference between the values obtained using the corrected LASS fit and the fit of Ref. . ## III The Isovector Scalar Decay Constants The spectral function, $`\rho (s)`$, of the isovector scalar correlator, $`\mathrm{\Pi }(q^2)`$, receives contributions from the two $`a_0`$ resonances proportional to $`f_{a_0(980)}^2`$ and $`f_{a_0(1450)}^2`$, respectively. This makes a QCD sum rule treatment of $`\mathrm{\Pi }(q^2)`$, in which the OPE of $`\mathrm{\Pi }(q^2)`$ is used to fix unknown parameters of the hadronic spectral function, such as $`f_{a_0}^2`$, an especially favorable way to distinguish between “large decay constant” and “small decay constant” scenarios. Previous attempts to determine $`f_{a_0(980)}`$ using QCD sum rules employed the conventional Borel transformed (SVZ) sum rule method (see Section 7.3 of Ref. for details). Borel transformation of the original dispersion relation converts the weight in the hadronic spectral integral into an exponentially falling one, $`\mathrm{exp}(s/M^2)`$, where the Borel mass, $`M`$, is a parameter of the transformation, and, simultaneously, both kills subtraction constants and provides factorial suppression of higher dimension condensate contributions on the OPE side ($`1/\left(Q^2\right)^n1/(n1)!M^{2n}`$, with $`Q^2=q^2=s`$). Small values of $`M`$ suppress dependence of the transformed spectral integral on the unknown high-$`s`$ portion of the spectral function, while large values suppress dependence of the OPE side on unknown higher dimension condensates. Ideally, one finds a “stability plateau (or window)”, i.e., a region of $`M`$ values for which both suppressions are reasonably operable, and within which extracted spectral parameters are roughly constant. Since, typically, the high-$`s`$ contribution to the transformed spectral integral is not negligible in the stability window, one requires a model for this portion of the spectral function. The standard approach is to use the “continuum ansatz”, or local duality approximation, for $`\rho (s)`$, for all $`s`$ beyond some “continuum threshold”, $`s_0`$. This is known to be a crude approximation in regions where typical resonance widths are not much greater than typical resonance separations, and hence leads to systematic uncertainties if continuum contributions to the spectral integral are significant. In the case of the earlier sum rule analyses of $`f_{a_0(980)}`$, one sees, from Fig. 7.6 of Ref. , both the absence of a stability plateau and a strong sensitivity to the choice of $`s_0`$. The latter observation signals the importance of contributions from the continuum region. An alternate implementation of the QCD sum rule approach, which avoids the use of the local duality approximation, is the finite energy sum rule (FESR) method. Consider the so-called “PAC-man” contour, which traverses both sides of the physical cut between threshold, $`s_{th}`$, and $`s_0`$, on the timelike $`q^2=s`$ axis, and is closed by the circle of radius $`s_0`$ in the complex $`s`$-plane. Cauchy’s theorem and analyticity then ensure that $$_{s_{th}}^{s_0}𝑑s\rho (s)w(s)=\frac{1}{2\pi i}_{|s|=s_0}𝑑sw(s)\mathrm{\Pi }(s),$$ (14) for any function $`w(s)`$ analytic in the region of the contour. Typically one wishes to use spectral data and/or a spectral ansatz on the LHS of Eq. (14), and the OPE on the RHS. The isovector scalar channel is very well adapted to the FESR approach, as far as the hadronic side of the sum rule is concerned, since the first two resonances in the channel are rather well-separated. Thus, if one chooses $`s_0`$ so as to include only the first two resonance regions, a well-motivated, resonance-dominated spectral ansatz is possible. Since the resonance masses and widths are known, only the decay constants remain as unknown parameters. The potential problem with the FESR approach is that those $`s_0`$ for which one can reliably employ a resonance-dominated ansatz for $`\rho (s)`$ correspond, by definition, to scales for which local duality is not a good approximation. This means, in particular, that the OPE representation for $`\mathrm{\Pi }(s)`$ on the circle $`|s|=s_0`$ must, at the very least, break down over some region near the timelike real axis. Fortunately, an argument by Poggio, Quinn and Weinberg, suggests that, for moderate $`s_0`$, this is the only region over which this breakdown should occur. This is confirmed empirically by the success of the conventional FESR treatment of hadronic $`\tau `$ decays (which involves a weight, determined by kinematics, having a double zero at $`s=s_0=m_\tau ^2`$), and by the investigation of Ref. , which shows that FESR’s involving weights $`w(s)`$ having either a single or double zero at $`s=s_0`$ are all extremely well satisfied in the isovector vector channel (where one can use the experimentally determined spectral function), even at scales, $`s_0`$, significantly below $`m_\tau ^2`$. In Figure 1 we show that such “pinch-weighted” FESR’s can be used to very accurately extract hadronic spectral parameters. In the Figure, the dots (with error bars) are the experimental ALEPH data; the dashed line shows the result of using the OPE side of a continuous family of such sum rules to fix the decay constants appearing in a simple spectral ansatz consisting of an incoherent sum of three Breit-Wigner resonances; and the solid line represents a least squares fit of the same spectral ansatz directly to the data. Note the very accurate determination of the $`\rho (770)`$ decay constant. This determination is dominated by the perturbative ($`D=0`$) contributions to the OPE, and hence by the value of the running coupling, $`\alpha _s`$. While a determination of a non-perturbative quantity like $`g_\rho `$ in terms of a perturbative one like $`\alpha _s`$ might sound implausible, we remind the reader that this is possible because we have used empirical non-perturbative information (in the form of resonance masses and widths) as input to the sum rules. Once this information has been input, analyticity relates the running coupling and the resonance decay constants via the basic FESR relation Eq. (14). In light of above discussion, we chose to study the isovector scalar channel using the pinch-weighted FESR method. For the spectral function we employ a sum of $`a_0(980)`$ and $`a_0(1450)`$ contributions, using experimental input for the masses and widths. $`s_0`$ is then restricted to lie at or below the upper edge of the second resonance region (actually, somewhat higher, since the zero at $`s=s_0`$ means that the region just below $`s_0`$ in the spectral integral has negligible weight). We take the maximum value of $`s_0`$ to be $`3.0\mathrm{GeV}^2`$. On the OPE side, the expressions for the dimension $`D=0,2,4`$ and $`6`$ contributions are given in Refs. , and will not be reproduced here. The $`D=0`$ (perturbative) contribution is known to four-loop order, and is determined by the running light quark masses and coupling. The light quark mass scale is set by any one of $`m_u,m_d,m_s`$, since the mass ratios are determined by ChPT. We employ, as our basic input, the value $`m_s(2\mathrm{GeV})=115\pm 8\mathrm{MeV}`$ (quoted in the $`\overline{MS}`$ scheme) determined in Ref. . Ratios of decay constants are, of course, independent of this choice. As input for the running coupling we take $`\alpha _s(m_\tau ^2)=0.334\pm 0.022`$. The masses and couplings at other scales are obtained from the exact solutions of the RG equations generated using the four-loop truncated versions of the $`\beta `$ and $`\gamma `$ functions. As input for the higher dimension condensate contributions, we employ conventional values: $`\alpha _sG^2=0.07\pm .01\mathrm{GeV}^4`$, $`\left(m_u+m_d\right)\overline{u}u=f_\pi ^2m_\pi ^2`$, and $`g\overline{q}\sigma Fq=\left(0.8\pm 0.2\mathrm{GeV}^2\right)\overline{q}q`$. The four-quark $`D=6`$ condensate terms are taken to have their vacuum saturation values, modified by an overall multiplicative factor, $`\rho _{VSA}`$. To be conservative, we consider the range $`\rho _{VSA}=5\pm 5`$, allowing, therefore, up to an order of magnitude violation of vacuum saturation. Integrals around the circle $`|s|=s_0`$ on the OPE side of Eq. (14) are all performed using the contour improvement prescription of LeDiberber and Pich, which is known to improve convergence, and reduce residual scale dependence of the truncated perturbative series. Since the convergence of the integrated $`D=0`$ contributions worsens as $`s_0`$ is lowered, we have chosen $`s_0=2.4\mathrm{GeV}^2`$ as a minimum value for our analysis. Ideally, to improve convergence, one would like to know the mass and width of the third $`a_0`$ resonance, and then work at larger scales $`s_0`$, but this is not possible at present. We discuss our estimate of the resulting truncation errors below. The last point in need of discussion is the treatment of instanton effects. These effects can be important in scalar and pseudoscalar channels, especially at lower scales such as those we have been forced to work at here. The effect of instantons on $`\mathrm{\Pi }(s)`$ is known exactly only in the approximation in which one treats the single instanton configuration in the background of the perturbative vacuum. It is known, however, that quark and gluon condensates strongly affect the density for large scale instantons. Since contributions in Refs. , from instantons large enough to be subject to this effect are known to be significant, the exact results are of only schematic use (as pointed out by the authors themselves). An alternate representation of instanton effects is provided by the instanton liquid model, in which the effective instanton density is assumed to be sharply peaked around a single effective average size. Although this sharp peaking leads to a slower fall-off with $`q^2`$ than would be obtained using a broader distribution of sizes, the model has the advantage of being phenomenologically constrained. We have, therefore, employed the instanton liquid model. Ref. gives the form of the corresponding contributions to FESR’s with $`w(s)=s^k`$. In view of the the phenomenological nature of the model, we chose, to be safe, to restrict ourselves to weights $`w(s)`$ for which instanton effects are not large. In order that this suppression not be specific to the particular $`q^2`$-dependence of the instanton liquid model, we further restrict ourselves to those weights for which an evaluation using the much more strongly $`q^2`$-dependent form obtained in Ref. are also small. This turns out to restrict us to weights of the form $$w(s)=\left(1\frac{s}{s_0}\right)\left(A\frac{s}{s_0}\right),$$ (15) with the parameter $`A`$ lying in the range $`2\pm 1`$. The suppression of the incompletely known instanton contributions is optimal for $`A=2`$, and we display all results below employing this value. (Note that the weight $`w(s)`$ has been constructed so as to have a zero at $`s=s_0`$, which is required in order that the resulting FESR’s be reliable at scales such as that considered here.) We will use the difference between the results obtained using the two different instanton implementations as a (hopefully conservative) measure of the uncertainty associated with our use of the instanton liquid model. Fitting the $`a_0(980)`$ and $`a_0(1450)`$ decay constants by matching the hadronic and OPE sides of the FESR above in the fit window $`2.4\mathrm{GeV}^2<s_0<3.0\mathrm{GeV}^2`$, we then obtain, adding errors from all sources in quadrature, $`f_{a_0(980)}m_{a_0(980)}^2`$ $`=`$ $`0.0447\pm 0.0085\mathrm{GeV}^3`$ (16) $`f_{a_0(1450)}m_{a_0(1450)}^2`$ $`=`$ $`0.0647\pm 0.0123\mathrm{GeV}^3.`$ (17) Uncertainties due to those on the resonance masses and widths are completely negligible; the errors, therefore, reflect uncertainties in the input to the OPE side of the sum rules. The major sources of error on $`f_{a_0(980)}m_{a_0(980)}^2`$ are as follows: (1) from that on $`\alpha _s(m_\tau ^2)`$: $`\pm 0.0068\mathrm{GeV}^3`$, (2) from that on the overall mass scale, $`\left(m_sm_u\right)^2`$: $`\pm 0.0032\mathrm{GeV}^3`$, (3) from that on $`\rho _{VSA}`$: $`\pm 0.0023\mathrm{GeV}^3`$ and (4) due to truncation of the perturbative series: $`\pm 0.0029\mathrm{GeV}^3`$. Those on $`f_{a_0(1450)}m_{a_0(1450)}^2`$ are, similarly: (1) from that on $`\alpha _s(m_\tau ^2)`$: $`\pm 0.0112\mathrm{GeV}^3`$, (2) from that on the overall mass scale: $`\pm 0.0046\mathrm{GeV}^3`$, (3) from that on $`\rho _{VSA}`$: $`\pm 0.0015\mathrm{GeV}^3`$ and (4) due to truncation of the perturbative series: $`\pm 0.0011\mathrm{GeV}^3`$. If one employs, instead of the instanton liquid model, the results of Ref. , $`f_{a_0(980)}m_{a_0(980)}^2`$ is increased by $`0.0044\mathrm{GeV}^3`$ and $`f_{a_0(1450)}m_{a_0(1450)}^2`$ by $`0.0023\mathrm{GeV}^3`$. We remind the reader that the instanton liquid model is subject to phenomenological constraints, while the results of Ref. are not. One final source of uncertainty, not included in the errors quoted above, is relevant in the case of the $`a_0(1450)`$. Because the $`a_0(980)`$ is well separated from subsequent resonances in the channel, its extracted decay constant is stable with respect to assumptions about the behavior of the spectral function beyond $`3\mathrm{GeV}^2`$. This is, however, not necessarily true for the $`a_0(1450)`$, which might, for example, overlap to some extent, and hence interfere with, the next higher resonance. Even were this interference to be incoherent, there could still be a non-trivial contribution from the tail of the next $`a_0`$ under the $`a_0(1450)`$. This would lead to an overestimate of the $`a_0(1450)`$ decay constant. We cannot meaningfully investigate this uncertainty since we do not know the width, or location, of the next $`a_0`$ resonance. In order to get a feel for the resulting uncertainty, however, we have investigated the effect of including a third resonance in the spectral ansatz, taking, for illustration, its mass and width to be $`1.9\mathrm{GeV}`$ and $`0.4\mathrm{GeV}`$, respectively. Assuming the contributions add incoherently, one finds that the $`a_0(1450)`$ decay constant is decreased by $`40\%`$. That level of uncertainty in the $`a_0(1450)`$ decay constant is, therefore, unavoidable without further experimental input. In contrast, the $`a_0(980)`$ decay constant is, as expected, essentially unaffected by the use of the three-resonance ansatz. ## IV Discussion For results obtained using QCD sum rules to be considered reliable, it is necessary that the form of the spectral ansatz employed be physically sensible. Even the most ridiculous spectral ansatz will have some choice of parameters which “optimizes” the match between the OPE and hadronic sides. Of course, if the ansatz is not a good one, this “optimal” match will be poor. While at the scales we have employed, resonance dominance seems a very safe assumption, it is worthwhile checking this statement. This is done in Figure 2. The Figure displays the results for the $`A=2`$ pinch-weighted FESR discussed above, both for the two-resonance ansatz on which the results above are based, and for two other ansätze, which serve as the basis for further discussions below. In the Figure, the dashed-dotted line represents the OPE side of the sum rule, the dotted line the optimized match of the hadronic to the OPE side, obtained by adjusting the two $`a_0`$ decay constants. The agreement is obviously excellent. The solid line, in contrast, represents the match obtained when one forces $`f_{a_0(980)}^20`$ by hand (as would be expected for a loosely-bound $`K\overline{K}`$ molecule) and adjusts $`f_{a_0(1450)}^2`$ to optimize the match. The resulting fit is very poor, ruling out the loosely-bound $`K\overline{K}`$ interpretation of the $`a_0(980)`$. Similarly, the fact that no good match exists with $`f_{a_0(980)}`$ much larger than $`f_{K_0^{}(1430)}`$ rules out the minion interpretation. A sceptical reader might object that the poor fit between the OPE and hadronic sides in the $`a_0(1450)`$-only spectral ansatz might be cured, not only by the addition of a narrow $`a_0`$ contribution, but also by the inclusion of a broader non-resonant background, even though such a possibility appears somewhat perverse, in view of the extremely good match between the dashed-dotted and dotted curves. The dashed line in the Figure demonstrates that this possibility is also ruled out, for reasons we will now explain. Although we do not know, in detail, what to expect for the background $`\pi \eta `$ and $`K\overline{K}`$ contributions to the isovector scalar spectral function below the $`a_0(1450)`$ region, it is rather easy to make sensible rough estimates. Indeed, once one knows the scalar form factors describing the couplings of the scalar densities to the $`\pi \eta `$ and $`K\overline{K}`$ states, one obtains the corresponding contributions to the spectral function by unitarity, as in Eq. (11). Near threshold, the timelike form factors may be trivially computed using ChPT. If one performs this exercise for the $`\pi \eta `$ contribution, using the tree-level ChPT expression for the timelike form-factor, the optimized background-plus-$`a_0(1450)`$ fit is very close to that of the solid line in Figure 2. The dashed line is the result of multiplying this background contribution by a factor of $`5`$, and then adjusting $`f_{a_0(1450)}`$ so as to optimize the match between the OPE and hadronic sides (but still with no explicit $`a_0(980)`$ contribution). Again one sees that the fit is very poor, demonstrating that it is a narrow state, and not a broad background which is required in the spectrum at low $`s`$. We, thus, conclude that the results obtained above are, indeed, reliable, subject to the quoted errors and the caveats regarding the $`a_0(1450)`$ decay constant already discussed. We see that (1) the products $`f_{a_0}m_{a_0}^2`$ for the two $`a_0`$ resonances are comparable and (2) that both are somewhat smaller than the corresponding product for the $`K_0^{}(1430)`$. In addition to ruling out the $`K\overline{K}`$ molecule and minion pictures, these results suggest a UQM-like scenario. One ambiguity which is unavoidable concerns the way in which one interprets the expected $`SU(3)_F`$ relations amongst the decay constants for the states lying in the same multiplet as the $`K_0^{}(1430)`$. In the $`SU(3)_F`$ limit, of course, all states in the same multiplet would have the same mass, so whether one compared the values of $`fm^2`$ or the values of $`f`$ would make no difference. In attempting to decide whether or not the $`a_0(980)`$ should be interpreted as the $`I=1`$ partner of the $`K_0^{}(1430)`$, however, this ambiguity plays a potentially significant role: if we compare decay constants, then we have $$\frac{f_{a_0(980)}}{f_{K_0^{}(1430)}}=1.1$$ (18) whereas, if we compare the products of the decay constants and the squared masses, we have $$\frac{f_{a_0(980)}m_{a_0(980)}^2}{f_{K_0^{}(1430)}m_{K_0^{}(1430)}^2}=0.53.$$ (19) Since, in other channels, it is the strange state which has the larger decay constant, however, neither of these results, in fact, corresponds to what one might expect based on the pattern from other meson nonets. In conclusion, we have determined the scalar decay constants of the $`a_0(980)`$, $`a_0(1450)`$ and $`K_0^{}(1430)`$ mesons. The relations between them suggest a UQM-like scenario for the isovector scalar states. At present, we do cannot be certain that the results rule out the four-quark interpretation of the $`a_0(980)`$, though it appears likely that one would expect a much smaller value for $`f_{a_0(980)}`$ in this scenario. A calculation of the $`f_{a_0(980)}`$ in the four-quark picture would thus be highly desirable. ###### Acknowledgements. The author would like to thank M. Pennington for a number of very useful conversations, and for bringing the existence of the transcription errors in the published version of the LASS parametrization to his attention; W. Dunwoodie for confirming the revised values of the $`K_0^{}(1430)`$ fit parameters; and A. Höcker for a number of discussions concerning the ALEPH spectral data and their interpretation. The ongoing support of the Natural Sciences and Engineering Research Council of Canada, and the hospitality of the Special Research Centre for the Subatomic Structure of Matter at the University of Adelaide, where this work was performed, are also gratefully acknowledged.
no-problem/9906/hep-ph9906499.html
ar5iv
text
# 1 An example of the ladder diagram contributing to the differential spin structure function. ## Acknowledgments This research has been supported in part by the Polish Committee for Scientific Research grants 2 P03B 184 10, 2 P03B 89 13 and 2P03B 04214 and by the EU Fourth Framework Programme ’Training and Mobility of Researchers’, Network ’Quantum Chromodynamics and the Deep Structure of Elementary Particles’, contract FMRX–CT98–0194.
no-problem/9906/astro-ph9906282.html
ar5iv
text
# The x-ray background-foreground galaxy cross-correlation: evidence for weak lensing? ## 1 Introduction After many years of observational work and theoretical investigations, the nature and origin of the unresolved component of the cosmic X-ray background (XRB) still remains an unsolved problem. The deep X-ray imaging data, combined with optical spectroscopic observations, now suggest that up to $``$ 70% of the soft XRB observed with ROSAT in the 0.5 to 2.0 keV energy band is resolved to individual galaxies, mainly active galactic nuclei (AGN), out to redshifts of $``$ 4 and greater (e.g., Miyaji et al. 1998a; Hasinger 1999 contains a recent review). Other than various possibilities that have been suggested in the literature, the exact nature of the remaining contributors to the soft XRB has not been clearly established. The possibilities for candidates so far include a population of low-luminosity galaxies and AGNs, and an optically obscured population of moderate to high redshift and high luminosity galaxies and AGNs. The strong isotropy of the unresolved component of the XRB, as measured by its auto-correlation function, requires that most of the sources responsible are at high redshifts and constraints models involving a population of low redshift and low luminosity AGNs. Returning to an obscured population at optical wavelengths, the hard XRB requires a ratio of obscured to unobscured populations of AGNs that amount to a factor as high as $``$ 3; As discussed in Almaini et al. (1999), the implications for such an obscured population is wide ranging. Recent experimental developments now allow some of these possibilities to be observationally tested. For example, the obscured population at optical wavelengths is expected to be visible at submm and far-infrared (FIR) wavelengths, through reemission of absorbed UV radiation by dust at longer wavelengths. Such sources should now be detected through deep observations with Submm Common User Bolometer Camera (SCUBA; Holland et al. 1998) on the James Clerk Maxwell Telescope. The current ongoing deep surveys with SCUBA will eventually test the exact fraction of obscured AGNs (see, Smail et al. 1999 for a recent review), with initial results suggesting that a dominant AGN fraction as high as 30% may be contributing to current SCUBA number counts (e.g., Cooray 1999a). At hard X-ray wavelengths, most of the Compton-thick AGNs which are absent at soft X-ray bands are expected to be present. Such populations have now been searched with ASCA and the Italian-Dutch BeppoSax satellite (Piro et al. 1995) in the 2 to 10 keV energy band. Contrary to expectations, however, these surveys are finding that all hard X-ray sources have soft X-ray counterparts (Hasinger 1999; however, see, Fiore et al. 1999). As most of these FIR/submm and hard X-ray observational programs are still ongoing, it is unlikely that an exact answer on the sources responsible for the unresolved component will soon be available. Recently, the existence of a high redshift population of low luminous X-ray emitting sources has been suggested by Haiman & Loeb (1999). These sources are present in cosmological models of hierarchical structure formation and are associated with the first generation of quasars. The presence of a high redshift population of X-ray emitting sources is also suggested by the possibility that there is no clear evidence for a decline in X-ray AGN number counts beyond a redshift of 2.5 (e.g., Miyaji et al. 1998b), which is contrary to optical quasar surveys where a decline has been inferred at high redshifts (e.g., Schmidt et al. 1995). According to the expected number counts of high redshift AGNs from Haiman & Loeb (1999), the contribution to current unresolved XRB from a high redshift AGN population is greater than 90%. Thus, almost all of the present unresolved XRB can be explained with such a low X-ray luminous population and without invoking the presence of optically obscured or Compton-thick sources. In addition to analytical calculations presented in Haiman & Loeb (1999), a population of high redshift low luminous quasars is also present in Monte Carlo realizations of merger histories of dark matter halos based on extended Press-Schechter theory (see, e.g. Cole 1991; Kauffmann & White 1993; Somerville & Kolatt 1998) combined with semi-analytical models of galaxy and quasar formations (Cooray & Haiman, in preparation). Given that the direct detection of such low luminous AGNs at X-ray wavelengths is not likely to be possible with current observational programs, the evidence for such high redshift X-ray sources should be inferred through indirect methods. It is likely that this situation will soon change with upcoming X-ray satellites such as the Chandra X-ray Observatory (CXO)<sup>1</sup><sup>1</sup>1http://asc.harvard.edu and the X-ray Multiple Mirror (XMM) Telescope <sup>2</sup><sup>2</sup>2http://astro.estec.esa.nl/XMM/. In Almaini et al. (1997), a cross-correlation between the unresolved XRB at soft X-ray energies, based on three $``$ 50 ksec ROSAT deep wide-field deep observations, and foreground bright galaxies, down to B-band magnitude of 23, has been presented. Such a correlation has been previously investigated in various studies involving the nature of XRB and sources responsible for it (e.g., Lahav et al. 1993; Miyaji et al. 1994; Carrera et al. 1995; Roche et al. 1996; Refregier et al. 1997; Soltan et al. 1997). The cross-correlation showed a highly significant signal and has been interpreted as evidence for a population of low redshift sources, traced by bright optical galaxies, as contributors to the unresolved XRB. Such an interpretation is based on the fact that detected cross-correlation is due to clustering between sources responsible for the unresolved component of the XRB and optical galaxies. If clustering were not to be present, in a case in which sources responsible for the XRB and optical galaxies were physically distinct in redshift space - or at least at scales greater than $``$ 100 Mpc - one would not normally expect any cross-correlation signal to be present. Apart from clustering, however, physically distinct populations can produce detectable cross-correlation if the flux-limited number counts and/or spatial distribution of one population was affected by the other. A well known possibility is that gravitational lensing by foreground sources modifies the distribution and number counts of background sources. Thus, an alternative possibility for the unresolved XRB is a population of sources at high redshifts provided that their X-ray emission is gravitationally lensed through foreground large scale structure. The cross-correlation between such sources and foreground optical galaxies results from the fact that foreground galaxies are a biased tracer of the large scale structure. The presented cross-correlation effect here is similar to the one involving high redshift optical quasars and foreground galaxies as discussed in Bartelmann (1995) and Dolag & Bartelmann (1997). A more general treatment of the cross-correlation between foreground and background samples due to weak gravitational lensing could be found in Sanz et al. (1997) and Moessner & Jain (1998). In both these studies, cross-correlation between two distinct populations in redshift was suggested as a probe of weak lensing due to large scale structure. In Sect. 2, we further investigate this possibility by modeling the X-ray emission from background sources and considering weak lensing effects of X-ray number counts. We use recent results from Haiman & Loeb (1999) to describe the background X-ray population. The general framework for the weak lensing calculation follows Cooray (1999b). We refer the reader to Mellier (1998) for a recent review on weak gravitational lensing, its applications and observations. We follow the conventions that the Hubble constant, $`H_0`$, is 100 $`h`$ kms<sup>-1</sup>Mpc<sup>-1</sup> and $`\mathrm{\Omega }_i`$ is the fraction of the critical density contributed by the $`i`$th energy component: $`b`$ baryons, $`\nu `$ neutrinos, $`m`$ all matter species (including baryons and neutrinos) and $`\mathrm{\Lambda }`$ cosmological constant. ## 2 The x-ray background - foreground galaxy cross-correlation Here, we briefly describe the expected signal between a foreground galaxy population with number density $`n_g`$ and X-ray sources responsible for the unresolved XRB, $`n_x`$. The angular cross-correlation function between the two samples is: $$w(\theta )=\delta n_g(\widehat{\varphi })\delta n_x(\widehat{\varphi }^{})$$ (1) where $`\delta n`$ is the excess fluctuations at a given line of sight. The cross-correlation between two physically distinctive samples contain four terms (Moessner & Jain 1998): $`w(\theta )=`$ $`\delta n_g^c(\widehat{\varphi })\delta n_x^c(\widehat{\varphi }^{})+\delta n_g^c(\widehat{\varphi })\delta n_x^\mu (\widehat{\varphi }^{})`$ $`+\delta n_g^\mu (\widehat{\varphi })\delta n_x^c(\widehat{\varphi }^{})+\delta n_g^\mu (\widehat{\varphi })\delta n_x^\mu (\widehat{\varphi }^{}),`$ where $`\delta n^c`$ is the fluctuations due to clustering of the sources while $`\delta n^\mu `$ is fluctuations due to gravitational lensing. These two terms can be written as, $$\delta n^c(\widehat{\varphi })=_0^{\chi _H}𝑑\chi b(r(\chi )\widehat{\varphi },\chi )W(\chi )\delta (r(\chi )\widehat{\varphi },\chi )$$ (3) and, $$\delta n^\mu (\widehat{\varphi })=3(\alpha 1)\mathrm{\Omega }_m_0^{\chi _H}𝑑\chi g(\chi )\delta (r(\chi )\widehat{\varphi },\chi ),$$ (4) respectively. Here, $`\chi _H`$ is the comoving distance to the horizon, $`W(\chi )`$ is the radial distribution of sources, $`\alpha `$ is the slope of number counts of these sources, $`nS^\alpha `$ with flux $`S`$, $`b(r(\chi )\widehat{\varphi },\chi )`$ is the source bias with respect to matter distribution, assuming to be both scale and time dependent, and $`g(\chi )`$ is a weight function: $$g(\chi )=r(\chi )_\chi ^{\chi _H}\frac{r(\chi ^{}\chi )}{r(\chi ^{})}W(\chi ^{})𝑑\chi ^{}.$$ (5) In Eq.(3), (4) and (5), $`r(\chi )`$ is the comoving angular diameter distance written as $`r(\chi )=1/\sqrt{K}\mathrm{sin}\sqrt{K}\chi ,\chi ,1/\sqrt{K}\mathrm{sinh}\sqrt{K}\chi `$ for closed, flat and open models respectively with $`K=(1\mathrm{\Omega }_{\mathrm{tot}})H_0^2/c^2`$ and $`\chi `$ is the radial comoving distance related to redshift $`z`$ through: $$\chi (z)=\frac{c}{H_0}_0^z𝑑z^{}\left[\mathrm{\Omega }_m(1+z^{})^3+\mathrm{\Omega }_k(1+z^{})^2+\mathrm{\Omega }_\mathrm{\Lambda }\right]^{1/2}.$$ (6) The lensing term in the cross-correlation is due to the fact that number counts of lensed background sources are affected in two ways: magnification by a factor $`\mu `$ so that lensed counts reach a fainter flux level ($`S/\mu `$) and distortion of the observed area such that solid angle observed is reduced by a factor $`1/\mu `$. Thus, lensed number counts change to $`n^{}\mu ^{\alpha 1}S^\alpha `$ from unlensed counts of $`nS^\alpha `$. In the weak lensing limit, magnification $`\mu =1+2\kappa `$, where $`\kappa `$ is the convergence and is equivalent to a weighted projection, via $`g(\chi )`$, of the matter distribution along the line of sight to background sources (see, e.g., Jain & Seljak 1997; Kaiser 1998; Schneider et al. 1998). The four terms in the cross-correlation are respectively: (1) clustering of sources in the two samples, when their redshift distributions overlap (2) lensing of background sources by large scale structure front of them traced by foreground galaxies (3) lensing of foreground sources by large scale structure traced by background galaxies; this term is non-zero only if there is an overlap in redshift distribution between the two samples, and (4) lensing of both foreground and background sources by large scale structure. When there is no overlap in redshift between the two samples, terms (1) and (3) are zero, while the last term can be ignored as its contribution is an order of magnitude lower than the 2nd term involving lensing of background sources by foreground large scale structure. The gravitational lensing effect results from two effects: (1) magnification due to lensing such that sources too faint to be included due to flux limit are now introduced and (2) modification of the observed solid angle, or volume, such that number counts are diluted. Considering these two well known effects, finally, the cross-correlation between two samples separated in redshift space can be written in the weak lensing limit as: $`w_{gx}(\theta )=`$ $`3b_g\mathrm{\Omega }_m(\alpha _x1){\displaystyle _0^{\chi _H}}W_g(\chi ){\displaystyle \frac{g_x(\chi )}{a(\chi )}}`$ $`{\displaystyle _0^{\mathrm{}}}{\displaystyle \frac{dkk}{2\pi }}P(k,\chi )J_0\left[kr(\chi )\theta \right]`$ where $`W_g(\chi )`$ and $`W_x(\chi )`$ are the radial distributions of foreground galaxies and background X-ray sources, $`\alpha _x`$ is the slope of number counts of background X-ray emitting sources at the limit of the unresolved background, and $`b_g`$ is the galaxy bias, assuming a linear bias independent of scale and time. The detailed derivation of Eq.(7), can be found in Bartelmann (1995) for an Einstein-de Sitter Universe with an extension to general cosmologies and nonlinear evolution of the power spectrum in Dolag & Bartelmann (1997), Sanz et al. (1997) and Moessner & Jain (1998). Here, we have introduced slope of the number counts, $`\alpha _x`$, for background X-ray sources, while, for example, in Moessner & Jain (1998) background sources were considered to be galaxies with a logarithmic number count slope of $`s`$ in magnitudes. ### 2.1 Expected contribution from weak lensing In order to estimate the expected level of contribution from weak lensing effects, we describe the background X-ray sources following calculations presented in Haiman & Loeb (1999). The foreground sources are described following Almaini et al. (1997), with a redshift distribution that peaks at a redshift of $``$ 0.5 and decreases to zero by redshift around $``$ 2.0. Such a redshift distribution for galaxies down to a magnitude limit of 23 in B-band is consistent with observations. We assume that galaxies are biased such that $`b_g=1/\sigma _8`$, which should adequate for the present calculation. Since most of the galaxies are at low redshifts, our predictions are insensitive to the exact redshift distribution of background sources as long as their redshifts are greater than 2.0. For the purpose of this calculation, we consider a background redshift distribution in which X-ray sources are distributed around a mean redshift of $``$ 3.5. In Fig.1, we show the two foreground and background redshift distributions. There is a slight overlap in redshift between the two distributions, but we have ignored it for the purpose of this calculation. Our input dark matter power spectrum and its non-linear evolution is calculated following Cooray (1999) using the fitting formulae given in Hu & Eisenstein (1998) to obtain the transfer function and Peacock & Dodds (1996) to obtain the nonlinear evolution. We consider cosmologies in which $`\mathrm{\Omega }_b=0.05`$, $`\mathrm{\Omega }_\nu =0.0`$, $`h=0.65`$. The power spectrum is normalized to $`\sigma _8(=0.56\mathrm{\Omega }_m^{0.47})`$ as determined by number density of galaxy clusters (Viana & Liddle 1996). Following calculations presented in Haiman & Loeb (1999), we determined the slope of X-ray number counts, $`\alpha `$, at the limit of the unresolved XRB to be $``$ 1.2. This number, however, is not well determined and is highly sensitive to how one models the X-ray emission from high redshift low luminous sources and number counts of such sources, as derived based on the Press-Schechter theory. We note that a value for $`\alpha <1.0`$ produces a cross-correlation which is negative, while $`\alpha =1.0`$ produces no contribution to cross-correlation from weak lensing. Finally, in order to account for the finite point spread function (PSF) of the PSPC detector, we convolve the expected lensing contribution with a parametric form of the PSF given by Hasinger et al. (1992). In Fig.2, we show the expected contribution from weak lensing to be observed cross-correlation. The data and associated errors are from Almaini et al. (1997). The two curves show the expected contribution for two cosmological models involving $`\mathrm{\Omega }_m=1.0`$ and $`\mathrm{\Omega }_m=0.3`$, $`\mathrm{\Omega }_\mathrm{\Lambda }=0.7`$. ## 3 Discussion & summary As shown in Fig.2, the weak lensing contribution to the observed cross-correlation between the XRB and foreground bright galaxies is substantial. The fractional contribution in the simple model considered here amount up to and more than 50%. Ignoring such a contribution is likely to produce biased estimates on the amplitude of clustering or the luminosity density of X-ray sources. In this respect, we note that previous estimates on the number density and luminosities of sources responsible for the unresolved XRB, using the cross-correlation, is certainly overestimated. In addition to changing cosmological parameters, we can increase the lensing contribution by increasing the foreground galaxy bias or increasing the slope of the X-ray number counts at the limit of the unresolved XRB. Currently, both these quantities, more importantly the slope of the number counts, are unknown. Therefore, it is premature to consider detailed models to explain the XRB using weak lensing effects completely. Since galaxy bias, however, is not expected to be much larger than $`1/\sigma _8`$, especially at low redshifts considered here, and that the slope of number counts is not likely to be very steep, it is unlikely that weak lensing alone can be used to fully explain the observed cross-correlation signal. As shown in Fig.2, weak gravitational lensing and Poisson fluctuations can easily account for almost all of the detected cross-correlation. However, we note that, in addition to weak lensing by large scale structure, strong lensing by individual galaxies and clusters of galaxies can contribute to the observed signal at small lag angles. Such a contribution is likely to be smaller than the weak lensing effect; still, it is likely that we have underestimated the complete lensing contribution to cross-correlation between the unresolved XRB and foreground galaxies by only considering weak lensing effects. The most likely scenario is that the observed cross-correlation is both due to clustering, from a low redshift population overlapping with the galaxy distribution and weak lensing effects of a high redshift population. The isotropy of the XRB, from its auto-correlation function, requires that bulk of the sources are at redshifts greater than 1. The clustering analysis of the observed XRB-galaxy cross-correlation suggests that up to $``$ 40% of the unresolved XRB is due to faint low-redshift X-ray sources (e.g., Almaini et al. 1997; Roche et al. 1996; Soltan et al. 1997). The additional contribution could arise from the high redshift X-ray emitting sources, however, we note that intracluster medium of galaxy clusters and groups as well $``$ 10<sup>6</sup> Kelvin gas in outskirts of galaxies, where most of the baryons at low redshifts are now believed to be present (Cen & Ostriker 1999), can contribute to the unresolved XRB. In addition to clustering and lensing terms, an additional term is present in the cross-correlation at zero lag or when $`\theta =0`$ due to the Poisson behavior of the background. Even though this term only arises for $`\theta =0`$, the finite PSF produces a substantial contribution at angular separations out to $``$ 30 arcsecs; the contribution is proportional to the integrated luminosity density of X-ray sources. Following Almaini et al. (1997) and using the Miyaji et al. (1998a) luminosity function for X-ray AGNs at a redshift of $``$ 3.5, we have estimated such a Poisson noise contribution to the cross-correlation. In Fig.2, we show this term with a dotted line. A Poisson fluctuation contribution level similar to the one calculated and a weak lensing contribution similar to the one calculated for $`\mathrm{\Omega }_m=1.0`$, when added, can easily explain the observed cross-correlation signal. As stated earlier, given that we have no reliable knowledge on the number counts and foreground galaxy bias, such a fit to the observed data is meaningless. We leave the task of a detailed comparison between the observed XRB and galaxy cross-correlation and various models involving lensing, clustering and Poisson contributions to a later paper. In fact, if the contribution to cross-correlation from latter two terms can be independently determined, then the lensing contribution can be used as a probe of the high redshift low luminosity X-ray source population, in addition to possibilities as a cosmological probe and a method to determine foreground galaxy bias. For now, we strongly suggest that there is adequate evidence for a weak lensing contribution to the observed unresolved XRB - foreground galaxy cross-correlation. Here, we have presented a hypothesis for the observed cross-correlation between the unresolved XRB and foreground bright galaxies using a population of high redshift X-ray sources. The upcoming surveys with CXO and XMM will allow the detection of such high redshift low luminosity sources, as discussed in Haiman & Loeb (1999) for the case of CXO. The followup observations of such deep and planned X-ray imaging of wide fields will eventually test the presence of such a population. In fact, the planned Guaranteed Time Observations (GTO) of several deep fields with CXO, such as the Hubble Deep Field (HDF; Williams et al. 1996), can easily be used to test the hypothesis whether remaining contributors to the unresolved XRB are a low redshift or high redshift population. The possibility that whether the cross-correlation is due to clustering of low redshift sources or lensing of high redshift sources can then be statistically studied based on the observed redshift distribution and luminosity function of X-ray emitting sources. * We acknowledge useful discussions and correspondences with Omar Almaini, Zoltan Haiman and Lloyd Knox. Omar Almaini is also thanked for communicating results from his analysis on the XRB-Galaxy cross-correlation and for answering various questions on the nature of XRB in general. We also thank an anonymous referee for helpful comments and suggestions on the manuscript and acknowledge partial support from a McCormick Fellowship at University of Chicago.
no-problem/9906/nucl-th9906004.html
ar5iv
text
# Comment on “Tensor force in doubly odd deformed nuclei” ## Abstract An article by Covello, Gargano and Itaco $`[`$Phys. Rev. C 56, 3092 (1997)$`]`$ tries to find evidence for the important role of the residual tensor force between the valence proton and neutron in doubly odd deformed nuclei. It is shown that observable effects discussed by these authors do not fully justify their rather strong conclusions. Recently, the role of the effective proton–neutron (p–n) interaction with particular attention focused on the tensor force operating between the unpaired proton and neutron in odd–odd deformed nuclei was studied by Covello et al. . Their results obtained for <sup>176</sup>Lu were interpreted as evidence for the originally proposed central+tensor force which “may be profitably used for a systematic study of doubly odd deformed nuclei in the rare–earth region” . In this Comment we critically analyze rather strong conclusions articulated by Covello et al. and also by Itaco et al. . Our method and results , that was also questioned in Ref. , will be defended elsewhere. In their study, Covello et al. carried out calculations of the spectrum of the first K<sup>π</sup>=0<sup>-</sup> band built upon the {$`\frac{7}{2}`$+ $`\frac{7}{2}`$–} intrinsic configuration in <sup>176</sup>Lu and found out that very good agreement with experiment is achieved only if the central+tensor forces with the Gaussian radial shape are taken into account when the Newby (N) shift in this band is estimated. They stated that its empirical value is exactly reproduced when the p–n parameters as recommended by Boisson et al. are used. Then, not surprisingly, also the whole rotational band is very well reproduced since there is no evidence of mixing with other bands. But, it is interesting to note that this N shift was not satisfactorily calculated in Ref. even if the tensor terms of the Gaussian force were set active, see Table I. On the other hand, our finite–range values of this N shift the estimates of which are based on different empirical sets of the p–n parameters agree well with experiment as shown in Table I. Consequently, also our empirical sets of the p–n parameters with not well determined tensor strengths yield a consistent explanation of the spectrum of the lowest K<sup>π</sup>=0<sup>-</sup> band. Unfortunately, the authors of Ref. did not provide any comparison of their theoretical B<sub>N</sub> values in this band with our results . In our opinion, there is more serious difficulty concerning the predictive power of the experimental spectrum of the first K<sup>π</sup>=0<sup>-</sup> rotational band in <sup>176</sup>Lu. Since the relevant N shift is of the central type (NC shift ), it is not expected to possess significant tensor contributions . In our analysis the results of which are given in Table I, the central terms of the Gaussian force are sufficient to predict its reasonable theoretical value even though small tensor contributions with the right signs are calculated. Thus, this particular example have a little to do with the importance of the tensor–force effects. It is worth noting that the N shift measured in the K<sup>π</sup>=0<sup>+</sup> {$`\frac{7}{2}`$+ $`\frac{7}{2}`$+} band in <sup>174</sup>Lu and also in <sup>170,172</sup>Lu and <sup>176</sup>Ta which is of the tensor type (NT shift ) gives a better picture of the tensor–force effects. Although the strengths of the tensor forces are not well determined in our experimental set of the N shifts , their effect is well visible in Table I where different values of this N shift are collected. (Notice that our theoretical B<sub>N</sub> values are very different from those calculated in Ref. .) In our calculations, similarly as in the previous example, the finite–range tensor forces operate in the right direction. Their contributions are, however, more significant since the central forces alone do not provide an acceptable B<sub>N</sub> value. Nonetheless, the tensor–force effects are rather small in order that one can deduce a definite conclusion. The second point we want to discuss is the role of irregularities which are known to be present in rotational bands in <sup>176</sup>Lu. In Ref. and a subsequent preprint , the authors analyzed an odd–even staggering observed experimentally in the two lowest K<sup>π</sup>=1<sup>+</sup> rotational bands built upon the {$`\frac{9}{2}`$– $`\frac{7}{2}`$–} and {$`\frac{7}{2}`$+ $`\frac{9}{2}`$+} configurations, respectively. They suggested that this rather large staggering may be caused by direct Coriolis coupling with Newby–shifted K<sup>π</sup>=0<sup>+</sup> bands assigned as {$`\frac{7}{2}`$– $`\frac{7}{2}`$–} and {$`\frac{7}{2}`$+ $`\frac{7}{2}`$+}, respectively. But these effects are small; typically 5–15 % admixtures were reported in Refs. . In the former case, the main trends were reproduced with the central+tensor Gaussian force . The latter effect, that is equally well developed in the experimental spectrum, failed to be described satisfactorily. The best picture, even if it is hardly acceptable, was obtained with the same type of the residual p–n force . All these results were then interpreted as “clear evidence of the importance of the tensor–force effects” , see also Ref. . In our previous study , we have assumed an old interpretation for the two lowest K<sup>π</sup>=1<sup>+</sup> rotational bands with the band heads at 194 keV and 338 keV, respectively. The revised interpretation of these bands given by Klay et al. yields theoretical values of the Gallagher–Moszkowski (GM) splitting energies calculated with our sets of the p–n parameters which are in even better agreement with experiment than our previous results , see Table II. We have carried out preliminary calculations of the spectrum of <sup>176</sup>Lu assuming 42 low–lying rotational bands of positive parities and including the Coriolis interaction, intrinsic rotational contributions, recoil terms, diagonal terms of the residual p–n interaction, and also non–diagonal p–n mixing ($`\mathrm{\Delta }`$K=0 interaction); the latter was not considered in Refs. . We have used the same mean–field parameters of the Nilsson potential as in Ref. . The G<sub>T</sub> sets of the p–n parameters have been adopted from the same study. The most striking feature of our calculations is that there is strong $`\mathrm{\Delta }`$K=0 mixing between both K<sup>π</sup>=1<sup>+</sup> rotational bands as well as between their K<sup>π</sup>=8<sup>+</sup> GM partners with non–diagonal matrix elements $`<V_{pn}>5060`$ keV. When included, this interaction, for example, affects significantly empirical values of the GM splitting energies. This is demonstrated in Table II where corrected empirical values, that are obtained assuming the G<sub>T</sub>(GM) set of the p–n parameters for the $`\mathrm{\Delta }`$K=0 interaction, are written in the ninth column. This finding suggests that the odd–even staggering can be transferred from one band to the other; such a picture was not confirmed in Refs. . Further, since the K<sup>π</sup>=0<sup>+</sup> {$`\frac{7}{2}`$– $`\frac{7}{2}`$–} band is expected to lie very high in energy (its band head was tentatively placed at 1057 keV ), its influence on the low–lying K<sup>π</sup>=1<sup>+</sup> band is found to be smaller in our calculations than in those performed in Ref. . In particular, the odd–even staggering that is discussed in Ref. , is equally well explained in our calculations only if very strong Coriolis mixing with the relevant K<sup>π</sup>=0<sup>+</sup> band exists. It holds if our theoretical G<sub>T</sub> value of the corresponding N shift is taken from Table I. Let us note that also in this case the tensor–force effects are not negligible. However, due to tentative assignment of this K<sup>π</sup>=0<sup>+</sup> band , an extremely large absolute empirical B<sub>N</sub> value has to be regarded as very uncertain and can hardly be compared with our predictions. Unfortunately, the authors of Refs. did not provide any theoretical value of this quantity. On the other hand, the odd–even staggering in the K<sup>π</sup>=1<sup>+</sup> {$`\frac{7}{2}`$+ $`\frac{9}{2}`$+} band, that is badly described in Ref. , is very well reproduced in our calculations based on our well determined value of the N shift in the K<sup>π</sup>=0<sup>+</sup> {$`\frac{7}{2}`$+ $`\frac{7}{2}`$+} band in <sup>174</sup>Lu, see Table I. Moreover, we have found that, due to the $`\mathrm{\Delta }`$K=0 interaction, the latter staggering is partially transformed into the K<sup>π</sup>=1<sup>+</sup> band lying lower in energy. Nevertheless, a better analysis of this effect is required. In conclusion, we would like to stress that our previous statement that the p–n parameters of the tensor forces are not well determined in our set of presently known N shifts does not imply that their role should be negligible. Here, we are forced to infer that, although probably right in principle, the conclusions concerning the importance of the tensor–force effects drawn in Refs. are not sufficiently supported. The reason is that a particular example (the lowest K<sup>π</sup>=0<sup>-</sup> band in <sup>176</sup>Lu) cannot indicate general features which are known to be extremely subtle. Nearly the same picture is obtained when the space–exchange and spin–spin space–exchange forces for the description of the N shifts are assumed to be the most important. In such a way, the spectrum of the lowest K<sup>π</sup>=0<sup>-</sup> band does not provide any argument against our previous conclusions . It should be finally pointed out that there remains a place for a different explanation of the observed odd–even staggering in the low–lying K<sup>π</sup>=1<sup>+</sup> bands discussed in Refs. . The crucial point for its understanding lies in a proper estimate of non–diagonal mixing which is caused by the Coriolis coupling, as correctly suggested in Refs. , but also by the $`\mathrm{\Delta }`$K=0 interaction. Thus, we conclude that the odd-even staggering and its theoretical description including the tensor terms does not directly imply that “only this force is able to reproduce a sizable N shift” for both K<sup>π</sup>=0<sup>+</sup> bands under consideration. This work was supported by the Grant Agency in the Czech Republic under Contract No. GAČR 202/99/1718.
no-problem/9906/hep-ph9906443.html
ar5iv
text
# What have we learned and want to learn from heavy ion collisions at CERN SPS? ## 1 The QCD Phase Diagram ### 1.1 Theoretical progress The QCD phase diagram version circa 1999 , shown in Fig 1(a), looks rather different from what was shown at previous Quark Matter conferences. Most significant progress is seen in the large-density low-T region, where two new Color Super-conducting phases have appeared. Unfortunately heavy ion collisions do not cross this part of the phase diagrams: it belongs to neutron star physics. The subject is covered in talks , so I only make few comments here. The 2-flavor-like color superconductor CSC2 phase was known before , but realization that it should be induced by instantons has increased the gaps (and $`T_c`$) from a few MeV scale to $``$100 MeV (50 MeV). It is hardly surprising, since the $`same`$ interaction in $`\overline{q}q`$ channel is responsible for chiral symmetry breaking, with the gap (a constituent quark mass) 350-400 MeV. The symmetries of the CSC2 phase are similar to the electroweak part of the Standard Model, with scalar isoscalar $`ud`$ diquark as Higgs. The colored condensate breaks the color group, making 5 out of 8 gluons massive. The 3-flavor-like phase, CSC3, is brand new: it was proposed in based on one gluon exchange interaction, but in fact it is favored by instantons as well . Its unusual features include color-flavor locking and coexistence of both types of condensates, $`<qq>`$ and $`<\overline{q}q>`$. It combines features of the Higgs phase (8 massive gluons) and of the usual hadronic phase (8 massless “pions”). Surprisingly, at very large densities Cooper pairs are bound magnetically , leading to growing gaps at extremely large $`\mu `$. Another important new element is the (remnant of) the QCD tricritical point . Although we do not know where it is<sup>1</sup><sup>1</sup>1Its position is very sensitive to precise value of the strange quark mass $`m_s`$, we hope we know how to find it, see . All ideas proposed rotate around the fact that the order parameter, the famous sigma meson, is at this point truly massless, and creates a kind of a “critical opalecence”. ### 1.2 The Phase diagram as our Map The major goal of the heavy ion program in general is to reproduce as much as possible properties of dense/hot matter in bulk, and study qualitatively new phenomena like phase transitions. Now, with accurate statistical description of the particle composition, small and Gaussian event-by-event (EBE) fluctuations, and detailed data on collective flow, people are using macroscopic language more than before, and so it is entirely appropriate to start with a discussion of where we are on the phase diagram. Looking at the fireball by a detector is like looking at the Sun: one can only see the photo-sphere, with $`T6000^o`$, not the hotter interior. Because we can see many hadronic species, we can also trace down conditions at which the composition is formed. The result can be summarized as two separate freeze-out points, $`thermal`$ or kinetic one, and $`chemical`$: a recent sample of those coming from SPS, AGS and SIS data is shown in Fig.1(b). The chemical point are about the same for all A, but the thermal ones are “cooler” for larger A. For a discussion of freeze-out systematics see . Only at SPS we find that the two freeze-outs are well separated, and so we indeed found a new phase of matter at SPS, never seen before: namely a chemically frozen thermally equilibrated “resonance” gas. The zigzag shaped paths on the phase diagram are adiabatic curves for a particular EOS (b) which I have discussed at QM97 (c): one can see that measured freeze-out points at SPS and AGS happen to be close to the predicted paths. Do chemical and thermal points indeed follow the same adiabates of “resonance gas”? The points in Fig.1(b) are not accurate enough to say that, but it can be studied in models. In the URQMD cascade was studied<sup>2</sup><sup>2</sup>2This model of course has no zigzag and QGP, only the hadronic part of the path is the same. , and although the non-equilibrium effects in some observables can be large, the exact and equilibrium expressions for entropy differ by only about 6%. Also the kinetic models agree that most of the entropy is produced early, and then it is only slightly affected by re-scattering. In summary: the adiabatic paths seem to be the paths to follow! ## 2 Hadronic stage ### 2.1 The radial flow The main news from the last few years is that at AGS/SPS energy domain the heavy ion collisions really produce a Little Bang rather than a fizzle, with strong explosive transverse flow converting a significant part of thermal energy into that of collective motion. Looking two decades ago at the pp ISR data I have found no trace of transverse radial flow: the $`m_t`$ slopes for various secondaries were identical. Only in mid-90s a significant difference in slopes was observed, first for light- and then heavy-ion collisions. At all previous QM conferences the origin of these slopes was debated. Are differences in AA and pp due to “initial state” parton re-scattering or the final state re-scattering of hadrons? Now the debate is mostly over since the amount of evidences which proves the latter is overwhelming. Let me mention two of them. (i) Initial state scattering may broaden the nucleon $`m_t`$ spectra, but it then predicts the same slope for deuterons, contrary to observations. Only a correlation between the two nucleons can explain data, and the calculated flow reproduces it well. (ii) slopes for $`\pi ,K,N,d`$ depend linearly on particle mass (common flow velocity $`v_t`$), except for strange baryons. The largest deviation is found for $`\mathrm{\Omega }^{}`$: this is nicely explained by its early freeze-out due to smaller cross sections<sup>3</sup><sup>3</sup>3Why does the $`\varphi `$ not have a small slope as well? See discussion below. . How large is the transverse flow velocity $`v_t`$? The fits to “hydro-motivated” formulae have produced widely varying values, which also were strongly A- and rapidity dependent. The explanation was worked out in the hydro-kinetic framework , which I also described at QM97. The $`motto`$ of this work, “the larger the system, the further it cools”, explains both strong A and y dependence of flow. Very large $`v_t.6`$ and very low $`T_{thermal}=110120MeV`$ in PbPb were predicted<sup>4</sup><sup>4</sup>4As only very elementary kinetics of low energy pion and nucleon re-scattering is actually involved, any event generator like RQMD also “knew” about it. Just it was not put in the proper words.. Later analysis based heavily on NA44,NA49 HBT radii $`and`$ spectra have confirmed such selection . The value of $`v_t`$ is important because it tells us about the EOS of hadronic matter. The observed values are mostly generated by “resonance gas”, which at SPS has simple EOS $`p.2ϵ`$, $`p,ϵT^6`$. $`v_t`$ is large because “resonance gas” at $`T=120160MeV`$ has no Hagedorn “softening”<sup>5</sup><sup>5</sup>5The Hagedorn phenomenon caused by the exponential rize of spectral density was in fact found to exist in QCD without quarks. It accurately explains lattice data on the value of the deconfinement transition for $`N_c=2,3,4`$. So, the Hagedorn softening would have happened in QCD at $`T_{Hagedorn}260MeV`$, if not for the light-quark dominated transition at lower T.. The mixed phase is however softer, with the minimum of $`p/ϵ`$ (the the softest point) corresponding to all matter converted to QGP. It was predicted (a) that the collisions which start from this condition should last longer. We are waiting for the next SPS run, at 40 GeV, to see if this prediction is indeed confirmed. ### 2.2 Elliptic flow In high energy collisions the shape of the “initial almond” for non-central collisions leads to enhanced “in-plane” flow, in the direction of the impact parameter . It is very important because (as pointed out in (a)) it is developed $`earlier`$ than the radial one, and thus it may shed light on whether we do or do not have QGP at such time. Now it is measured by the asymmetry of the particle $`number`$, or $`v_i`$ harmonics defined as $`\frac{dN}{d\varphi }=\frac{v_0}{2\pi }+\frac{v_2}{\pi }\mathrm{cos}(2\varphi )+\frac{v_4}{\pi }\mathrm{cos}(4\varphi )+\mathrm{}`$ rather than asymmetry of the momentum distribution. Furthermore, $`v_i`$ can be additionally normalized to the spatial asymmetry of the initial state (the “almond”) at the same $`b`$, in order to cancel out this kinematic factor and to see the response to asymmetry. Let us first look at the energy dependence: a compilation of measured nucleon $`v_2`$ is shown in Fig.2(a). First news is indication for softening of EOS at $`E>6AGeV`$ observed by EOS detector at AGS, exactly where the initial conditions are expected to hit for the first time the critical line (see ). At the same energy $`K/\pi `$ ratio and other strangeness enhancement signals change rapidly. Both indicate that the mixed phase is actually reached already at such low energies. The dash-dotted curve for energies above the SPS is from hydro calculations . In agreement with , large and near constant ellipticity in this region is found, driven by the QGP push at early times. Strong decrease of this effect is expected in the SPS domain, as the “QGP push” disappears. The dashed curve is my speculation of what the excitation function of $`v_2`$ may be in the SPS domain. The existence of a $`plateau`$ is not obvious, but an inflection point seems inevitable<sup>6</sup><sup>6</sup>6 This prediction looks similar to that made by van Hove for radial flow. His argument was however too naive, because it ignored time development: soft EOS leads to longer expansion time. Therefore even larger velocity can be obtained by the end. But ellipticity is a small effect which does $`not`$ determine the expansion time: so in this case the argument should be valid.. Let us now turn to SPS data, which depict $`centrality`$ dependence of $`v_2`$. This issue is more complicated, because by increasing $`b`$ we make the “almond” more elliptic, but also much smaller and thinner: eventually finite size corrections reduce the pressure build-up<sup>7</sup><sup>7</sup>7And this is why the guessed dashed curve in Fig.2(a) is below the hydro prediction. . One needs cascade codes to study this effect: see . Last year RQMD has been radically changed, including the possibility to vary the EOS and to include the “QGP push”. New version (b) predicts a relative growth of (properly normalized to spatial asymmetry) $`v_2`$ at small $`b`$, to values close to those predicted by hydro. Preliminary NA49 data presented at this meeting indeed observe such enhancement at small $`b`$. Much more work is clearly needed to understand this complex interplay of EOS and finite size effects. U collisions discussed below in principle provide the means to decouple finite size and deformation issues. The next run (40 AGeV) data are also of great importance here: no “QGP push” is expected there. If a clear difference in $`v_2(b)`$ between 40 and 158 GeV is observed, it would really be the first sign of the QGP push at SPS. Whether it is seen at SPS or not, it certainly is expected to lead to quite dramatic phenomena at RHIC. A very non-trivial and highly EOS-sensitive expansion pattern is predicted in : The so called “nutcracker” scenario (for non-central collisions) includes formation of two “shells”, which are physically separated from each other by freeze-out, with only a small “nut” at the center. This behavior predicts higher moments $`v_4,v_6`$ of specific kind growing with energy, and spectacular HBT radii. We are looking forward to first RHIC data to see if this is true. ### 2.3 The event-by-event fluctuations: all events are (about) equal! Search for “unusual” events (e.g. disoriented chiral condensate) had attracted significant attention in the past, but they were not found. NA49 has shown that fluctuations in such observables as $`<p_t>`$ are rather small and Gaussian, without unusual tails. Their widths can be measured rather accurately. What can we learn from them? It is tempting to apply thermodynamical approach based on entropy here<sup>8</sup><sup>8</sup>8 Such type of analysis of multi-hadron production reactions goes back to 70’s, see e.g. my own work where probabilities of exclusive channels for low energy $`\overline{p}p`$ annihilation were calculated from ideal gas entropy.. Furthermore, as the temperature fluctuation is related to specific heat $`\mathrm{\Delta }T^2/T^2=1/C_v`$, it was proposed in to use it to measure $`C_v`$ of a hadronic matter at freeze-out. For example, at the critical point mentioned above, $`C_v`$ diverges and $`\mathrm{\Delta }T`$ must vanish. Can this dramatic prediction be somehow observed? Unfortunately, as explained in , this argument is too naive. Pions can indeed be used as a “thermometer”, in contact with a fluctuating medium (the sigmas), but fluctuations of its T measures mostly the thermometer’s own $`C_v`$. The critical fluctuations can be seen, but only as a correction at the 10% level or so. The fluctuations for intensive and extensive observables are generally of different nature. An example of the former case is $`<p_t>`$: the deviation of the measured width from pure statistics (mixed events) is in this case surprisingly small<sup>9</sup><sup>9</sup>9Even Bose enhancement which definitely is there in HBT measurements is canceled out.. Fluctuations of particle composition are also intensive, and they are sensitive to resonances. New point: if one would be able to get any hold on multiple production of multi-strange objects, one can tell if such exotic objects as “color ropes” do or do not exist. An example of an extensive variable is total multiplicity: even restricted to central 5% NA49 finds a Gaussian width $`twice`$ that for random (Poisson) emission. As shown in , only about half of the effect comes from resonances at freeze-out. The rest must come from fluctuations in initial conditions: see discussion in . I think the most interesting part of the development is attention to mixture of the two, such as $`<\mathrm{\Delta }p_t\mathrm{\Delta }N>`$. As explained in , those do not appear in simple statistics, and so carry non-trivial information. ## 3 The initial (or pre-hadronic) stage ### 3.1 Two main scenarios Although we have now reached some understanding of the hadronic stage of the evolution, from chemical to kinetic freeze-out, we know very little about what happens before it. Ignoring details<sup>10</sup><sup>10</sup>10 E.g. disregarding purely hadronic models, because particle composition cannot be explained by reactions in hadronic phase. , let me emphasize two major points of view present on the market: (i) The QGP scenario (hopefully our future Standard Model), based on a picture of quick matter equilibration; and (ii) string scenario, represented by most traditional Lund-type event generators like Venus and RQMD, or reggion-based approaches like DPM. A positive feature of the former approach is obviously its conceptual simplicity, while the latter has direct connection to pp and pA phenomenology. $`Both`$ can explain why the initial EOS is soft enough not to contribute much to the radial flow. Particle composition would be a stronghold of thermodynamics, if not for the fact that in pp and $`e^+e^{}`$ collisions it is also possible to use successfully statistical models. The difference is however seen in the strangeness production, much suppressed in pp and $`e^+e^{}`$ but not in AA<sup>11</sup><sup>11</sup>11 Some studies even find that strangeness is slightly enhanced compared to the equilibrium value, partly due to Coulomb effects. . The QGP scenario explains it naturally, while the string one needs “color ropes” (or other exotic devices) to explain the strangeness. Significant experimental efforts have been made to locate the transition between two extremes, the pp-type and heavy-ion-type regimes of strangeness production. Excellent data from WA97,NA49 for $`\mathrm{\Lambda },\mathrm{\Xi },\mathrm{\Omega }`$ and their anti-particles found that strangeness has little dependence on centrality. Lighter ion data obtained before also confirm the impression that strangeness production is basically constant throughout the SPS domain, with the transition being somewhere in the AGS domain. To me it indicates that strangeness “un-suppression” is clearly related with the approach of the hadronic phase boundary, production of the mixed phase with at least $`some`$ QGP. We will separately discuss $`\rho `$ melting and $`J/\psi `$ suppression below: but let me now emphasize their contrasting A dependence. CERES data on dilepton enhancement below $`\rho `$ show comparable effect in SAu and PbAu, while the NA50 $`J/\psi `$ suppression is drastically different. Why is it so? It is a must to see if the energy dependence of two effects is different as well. By adding $`\psi ^{}`$ in between, and decomposing $`J/\psi `$ into $`\chi `$ and proper $`J/\psi `$ parts, one would get a whole sequence of “melting” phenomena, happening as matter becomes hotter/denser. We will see at RHIC how members of the $`\mathrm{{\rm Y}}`$ family do the same later on. Which of them is the QGP signal then? Well, this depends on details which we still have to work out. ### 3.2 Dileptons As “penetrating probes” , dileptons suppose to tell us the story of “melting” of all vector mesons ($`\rho ,\omega ,\varphi ,J/\psi ,\mathrm{{\rm Y}}`$), replaced by radiation from thermal quarks. All SPS dilepton experiments (HELIOS-3,CERES,NA50) see significant dilepton enhancement (compared to “trivial sources”), being stronger at small $`p_t`$, indicating matter effects. The first important issue is the origin of the enhancement observed by NA50 at $`M_{\mu ^+\mu ^{}}2GeV`$. It was suggested that it is due to enhanced charm production, but another (and, in my mind, much more probable) explanation is the thermal QGP emission . The QGP-like rate reproduces HELIOS3 data , and preliminary estimates show it works for NA50 data as well. The origin of a qualitative change of the shape of the vector spectral density $`\rho _v(M)`$ for $`M<1GeV`$ observed by CERES was discussed here in detail , let me therefore address only one central question: to what extent does the observed “$`\rho `$ melting” indicate an approach to chiral symmetry restoration? In-matter $`\rho `$ width is significantly increased, mostly by re-scattering on nucleons. The non-trivial fact is: even for small masses $`M=.3.6GeV`$ the rate (obtained in a complicated hadronic calculation ) happen to be rather close to the “partonic” or QGP rate, corresponding to free $`\overline{q}q`$ annihilation in the heat bath. It tells us that the interaction between $`\overline{q}`$ and $`q`$ in the vector channel is becoming weak. $`If`$ axial spectral density is modified similarly as well, the finite-T Weinberg-like sum rules demand that the chirally-odd quantities in its r.h.s. become small, which means chiral symmetry restoration. $`If`$ the axial spectral density remains different, chiral symmetry is still broken. Although we cannot access it directly, one may still look for Dalitz-type decays of $`a_1`$ . ### 3.3 $`J/\psi `$, $`\psi ^{}`$ suppression Let me start with a brief comment on the $`\psi ^{}`$ suppression. The NA50 data show that the $`\psi ^{}/\psi `$ ratio stop falling and is stabilized ata small value, about 4%. An explanation suggested in is: all $`\psi ^{}`$ are killed first, but then are re-created from $`J/\psi `$. The central NA50 finding is of course a statement that $`J/\psi `$ suppression for central (b$`<`$ 8 fm) PbPb collisions is different from extrapolations based on pA and SA collisions. Several mechanisms of this suppression were proposed: (i) gluonic photo-effect ; (ii) these states simply do not exist in QGP due to Debye screening ; (iii) hadronic co-movers kill them ; (iv) non-monotonous variation of the QGP lifetime, due to the “softest point” . New data reported at this meeting have clarified the situation for the most central collisions: using now a very thin target, it was found that the 1996 data suffered from multiple interactions. In fact there is a significantly stronger suppression at small $`b`$, making the two component picture with separate $`\chi `$ and $`\psi `$ thresholds more probable<sup>12</sup><sup>12</sup>12And the idea (iv), according to which suppression may even become weaker at smaller b less likely.. It is desirable to make another step, increasing the density and/or the famous variable L: only deformed U provides an opportunity here. What else can be done to discriminate experimentally these ideas? In particularly, how can we tell whether suppression happened quickly or took a longer time? The old idea is to study suppression dependence on $`p_t`$. Unfortunately changing $`p_t`$ we also change the kinematics: e.g. destruction by gluons or hadrons goes better if $`p_t`$ grows. Maybe a better idea is to use the azimuthal dependence of the suppression. Instantaneous suppression should show $`no`$ asymmetry, but if it takes a few fm/c the anisotropy should show up. The problem is the initial “almond” at $`b<`$ 8 fm is not very anisotropic, and for larger $`b`$ there is no anomalous suppression. (Here too the deformed U can help.) ### 3.4 $`\varphi `$-related puzzles $`\varphi `$ is a little brother of $`J/\psi `$, but its production in AA is $`enhanced`$ rather than suppressed, as compared to NN. Whatever effects are killing $`J/\psi `$’s, $`\varphi `$ is re-created because strangeness is close to equilibrium at chemical freeze-out. The first puzzle is the apparent absence of $`\varphi `$ in-matter modification. As argued , even modest modification of kaons should strongly (by factor 2 or so) increase the width of $`\varphi `$ decay inside the fireball. Non-negligible fraction of $`\varphi `$, up to a half, should decay in-matter, while experimentally (see e.g. excellent NA49 data presented here by C.Hohne) $`no`$ $`\varphi `$ modification is seen at all! The second puzzle (already mentioned in the flow section) is that the $`m_t`$ slopes of $`\varphi `$ spectra measured by NA49 in KK channel are large, close to those of the $`p,\overline{p}`$ (particles of similar mass). It suggests that somehow $`\varphi `$ participates fully in the radial flow<sup>13</sup><sup>13</sup>13Again, ellipticity may give a clue here.. How is it possible, with its small re-scattering cross section? New NA50 data on $`\varphi `$ reported at this conference have a different slope than NA49, only about 220 MeV. If extrapolated from larger $`p_t`$ (where NA50 data are) to small ones, they go well above the NA49 ones. All puzzles may be explained if absorption destroy K from most in-matter decays<sup>14</sup><sup>14</sup>14Or their $`refraction`$ in a collective potential., more so at low $`p_t`$. Obviously, there should be no missing $`\varphi `$ in dilepton channel, and so such experiments should see the $`\varphi `$ missing from KK channel at low $`p_t`$. (Phenix at RHIC will have good resolution in both channels, so it should eventually clarify the issue.) In summary: as a first (indirect) sign of missing (modified?) $`\varphi `$ we have evidences for the unusual change of its $`m_t`$ slope. If able to cover smaller $`p_t`$, NA50 should see $`\varphi `$ spectra which are different from NA49, including those which decay inside the fireball. ## 4 Conclusions and suggestions ### 4.1 Little Bang versus the Big one It is always fun to notice parallels to cosmology. There are many methodic similarities, as well as amusingly close timing of some recent developments. The first obvious connection between the “Little Bangs” in AA collisions and the “Big Bang” is that both are violent explosions. Expansion of the created hadronic fireball approximately follows the Hubble law, $`vr`$, although anisotropic one. The $`final`$ velocities, the Hubble constant and $`v_t`$, have been a matter of debates few years ago, but now are believed to be fixed (at say 10% level). The next important issue in both cases is the acceleration history, needed to shed some light on the fundamental EOS. Cosmologists use distant supernovae to access flow long ago: we use $`\mathrm{\Omega }^{}`$ to learn what was the flow at “mid-time” (5-10 fm/c). The last point: angular anisotropy of flow and its fluctuations. Amazingly accurate measurements of the microwave background have found a dipole anisotropy (due to our motion relative to “ether”) and tiny ($`10^5`$) fluctuations with angular momentum $`l100`$ due to frozen plasma oscillations, from the freeze-out stage in which the QED plasma was neutralized into ordinary atomic matter. In the Little Bang we have $`average`$ dipole and elliptic flows of a few percent. $`Fluctuating`$ higher harmonics are not analyzed yet: there must be some trace of “frozen QGP oscillations” as well. True, cosmologists have much more photons, but they are restricted to only one event, while we have millions of them! ### 4.2 Using deformed nuclei (U): An old idea with a new twist An old idea<sup>15</sup><sup>15</sup>15Kind of a folklore of our field, the only written version I found is a memo written by P.Braun-Munzinger to BNL. is to select head-on (long-long) collisions, by triggering on maximal number of participants $`N_p`$. Because of larger A and $`deformation`$, the gain in energy density can realistically reach 35-40% , which is important e.g. for the $`J/\psi `$ suppression studies. The main finding of my recent studies of UU collisions is however a possible virtue of “parallel” collisions, in which both long axis are orthogonal to the beam. Using $`two`$ control parameters, the number of participants and ellipticity<sup>16</sup><sup>16</sup>16The measured deformation of spectra, $`v_2`$, is proportional to this initial deformation (with EOS-depending coefficient!) and should have similar distribution. , one can effectively separate those. As one can see from Fig.2(b), unlike PbPb collisions, the UU ones provide a range of deformations, with long-to-short ratio reaching about 1.3. (Those correspond to collisions with two long directions parallel to each other and orthogonal to the beam). It is comparable to the deformation reached for mid-central collisions of spherical nuclei, but now for much larger and denser system. As mentioned several times in this talk, it may help to clarify many issues, such as presence of the QGP push in elliptic flow at SPS, a time scale of the $`J/\psi `$ suppression, etc. One more example are corrections to hard processes, like “shadowing” due to initial state re-scattering or “jet quenching” due to final state. Selecting two geometries, head-to-head and “parallel”, one can change the longitudinal to transverse size ratio from 1.3 to 1/1.3, a significant level arm to tell the difference. ### 4.3 Experimental goals It is clear that there can only be a finite time-span for the SPS heavy ion program, so we have to be “picky”. I think the following list includes only experiments which are a complete “must”. * – It is not likely this energy region would be studied later, so we better be sure no qualitative phenomenon is missed. For years I advocated measuring SPS excitation function looking for the “softest point”, and we will have the 40 GeV run soon. Now excitation function of $`v_2`$ became an important issue, with a potential to see “the QGP push” at SPS. Another compelling argument for a scan is hunting for the tricritical point of QCD: finding it would be a major breakthrough, going to textbooks etc. * – The nature of the dilepton excess for $`M2GeV`$ found by NA50 should be understood. If it is indeed charm enhancement, up to factor 3 for central PbPb, then the $`J/\psi `$ suppression issue is much more serious. If it is QGP radiation, it should be studied more. The number one hadronic measurements which remains to be done is therefore direct observation of charm by D’s. * $`J/\psi `$ $`suppression`$ is high priority, as the only observable in which relatively sharp centrality dependence. By changing the beam (A,collision energy) one should test whether the variation seen is or is not related to fixed energy density. Time-scale of the suppression can be accessed by studying its $`p_t`$ dependence or “ellipticity”. * – CERES: significant improvement of signal/background ratio is expected from its recent upgrade, leading to clear separation of the $`\omega `$ from the $`\rho `$ peak, as well as independent look at the $`\varphi `$ shape. If it works out as expected, new CERES would be an excellent tool to study dramatic in-matter modification (rather than just enhancement or suppression) of vector resonances. It is also significantly statistics-limited experiment, deserving running time as much as possible.
no-problem/9906/gr-qc9906084.html
ar5iv
text
# Untitled Document On the relation between a zero-point-field-induced inertial effect and the Einstein-de Broglie formula Bernard Haisch Solar and Astrophysics Laboratory, Dept. H1-12, Bldg. 252, Lockheed Martin 3251 Hanover Street, Palo Alto, California 94304 haisch@starspot.com Alfonso Rueda Department of Electrical Engineering & Department of Physics, ECS Building California State University, 1250 Bellflower Blvd., Long Beach, California 90840 arueda@csulb.edu (Physics Letters A, Vol. 268, pp. 224–227, 2000) Abstract It has been proposed that the scattering of electromagnetic zero-point radiation by accelerating objects results in a reaction force that may account, at least in part, for inertia . This arises because of asymmetries in the electromagnetic zero-point field (ZPF) or electromagnetic quantum vacuum as perceived from an accelerating reference frame. In such a frame, the Poynting vector and momentum flux of the ZPF become non-zero. If one assumes that scattering of the ZPF radiation takes place at the level of quarks and electrons constituting matter, then it is possible for both Newton’s equation of motion, $`𝐟=m𝐚`$, and its relativistic covariant generalization, $`=d𝒫/d\tau `$, to be obtained as a consequence of the non-zero ZPF momentum flux. We now conjecture that this scattering must take place at the Compton frequency of a particle, and that this interpretation of mass leads directly to the de Broglie relation characterizing the wave nature of that particle in motion, $`\lambda _B=h/p`$. This suggests a perspective on a connection between electrodynamics and the quantum wave nature of matter. Attempts to extend this perspective to other aspects of the vacuum are left for future consideration. Current Address: California Institute for Physics and Astrophysics, 366 Cambridge Ave., Palo Alto, CA 94306 (http://www.calphysics.org) Using the techniques of stochastic electrodynamics we examined the Poynting vector of the electromagnetic ZPF of the quantum vacuum in accelerating reference frames . This led to a surprisingly simple and intuitive relation between what should be the inertial mass, $`m_i`$, of an object of proper volume $`V_0`$ and the energy density of the ZPF instantaneously contained in $`V_0`$. Besides simplicity, this new approach improved over a previous one in that it yielded a covariant generalization. As derived from the force associated with the ZPF momentum flux in transit through the object, $`m_i`$ and $`\rho _{ZP}`$ were found to be related as follows: $$m_i=\frac{V_0}{c^2}\eta (\omega )\rho _{ZP}(\omega )𝑑\omega ,$$ $`(1)`$ where $`\rho _{ZP}`$ is the well known spectral energy density of the ZPF $$\rho _{ZP}(\omega )=\frac{\mathrm{}\omega ^3}{2\pi ^2c^3}.$$ $`(2)`$ Viewed this way, inertial mass, $`m_i`$, appeared to be a peculiar form of coupling parameter between the electromagnetic ZPF and the electromagnetically interacting fundamental particles (quarks and electrons) constituting matter. The key to deriving an equation of motion ($`=d𝒫/d\tau `$ in the relativistic case) from electrodynamics is to assume that some form of scattering of the non-zero (in accelerating frames) ZPF momentum flux takes place. The reaction force resulting from such scattering would appear to be the physical origin of inertia. The parameter $`\eta (\omega )`$ in eqn. (1) parametrizes such a scattering efficiency whose strength and frequency dependence have been unknown. It was proposed by de Broglie that an elementary particle is associated with a localized wave whose frequency is the Compton frequency, yielding the Einstein-de Broglie equation: $$\mathrm{}\omega _C=m_0c^2.$$ $`(3)`$ As summarized by Hunter : “…what we regard as the (inertial) mass of the particle is, according to de Broglie’s proposal, simply the vibrational energy (divided by $`c^2`$) of a localized oscillating field (most likely the electromagnetic field). From this standpoint inertial mass is not an elementary property of a particle, but rather a property derived from the localized oscillation of the (electromagnetic) field. De Broglie described this equivalence between mass and the energy of oscillational motion…as ‘une grande loi de la Nature’ (a great law of nature).” The rest mass $`m_0`$ is simply $`m_i`$ in its rest frame. What de Broglie was proposing is that the left-hand side of eqn. (3) corresponds to physical reality; the right-hand side is in a sense bookkeeping, defining the concept of rest mass. This perspective is consistent with the proposition that inertial mass, $`m_i`$, is also not a fundamental entity, but rather a coupling parameter between electromagnetically interacting particles and the ZPF. De Broglie assumed that his wave at the Compton frequency originates in the particle itself. An alternative interpretation is that a particle “is tuned to a wave originating in the high-frequency modes of the zero-point background field.” The de Broglie oscillation would thus be due to a resonant interaction with the ZPF, presumably the same resonance that is responsible for creating a contribution to inertial mass as in eqn. (1). In other words, the ZPF would be driving this $`\omega _C`$ oscillation. We therefore suggest that an elementary charge driven to oscillate at the Compton frequency, $`\omega _C`$, by the ZPF may be the physical basis of the $`\eta (\omega )`$ scattering parameter in eqn. (1). For the case of the electron, this would imply that $`\eta (\omega )`$ is a sharply-peaked resonance at the frequency, expressed in terms of energy, $`\mathrm{}\omega =512`$ keV. The inertial mass of the electron would physically be the reaction force due to resonance scattering of the ZPF at that frequency. This leads to a surprising corollary. It can be shown that as viewed from a laboratory frame, the standing wave at the Compton frequency in the electron frame transforms into a traveling wave having the de Broglie wavelength, $`\lambda _B=h/p`$, for a moving electron. The wave nature of the moving electron appears to be basically due to Doppler shifts associated with its Einstein-de Broglie resonance frequency. This has been shown in detail in the monograph of de la Peña and Cetto (see also Kracklauer ). Assume an electron is moving with velocity $`v`$ in the $`+x`$-direction. For simplicity consider only the components of the ZPF in the $`\pm x`$ directions. The ZPF-wave responsible for driving the resonant oscillation impinging on the electron from the front will be the ZPF-wave seen in the laboratory frame to have frequency $`\omega _{}=\gamma \omega _C(1v/c)`$, i.e. it is the wave below the Compton frequency in the laboratory that for the electron is Doppler shifted up to the $`\omega _C`$ resonance. Similarly the ZPF-wave responsible for driving the electron resonant oscillation impinging on the electron from the rear will have a laboratory frequency $`\omega _+=\gamma \omega _C(1+v/c)`$ which is Doppler shifted down to $`\omega _C`$ for the electron. The same transformations apply to the wave numbers, $`k_+`$ and $`k_{}`$. The Lorentz invariance of the ZPF spectrum ensures that regardless of the electron’s (unaccelerated) motion the up- and down-shifting of the laboratory-frame ZPF will always yield a standing wave in the electron’s frame. It has been proposed by de la Peña and Cetto and by Kracklauer that in the laboratory frame the superposition of these two waves results in an apparent traveling wave whose wavelength is $$\lambda =\frac{c\lambda _C}{\gamma v},$$ $`(4)`$ which is simply the de Broglie wavelength, $`\lambda _B=h/p`$, for a particle of momentum $`p=m_0\gamma v`$. This is evident from looking at the summation of two oppositely moving wave trains of equal amplitude, $`\varphi _+`$ and $`\varphi _{}`$, in the particle and laboratory frames . In the rest frame of the particle the two wave trains combine to yield a single standing wave. In the laboratory frame we have for the sum, $$\varphi =\varphi _++\varphi _{}=\mathrm{cos}(\omega _+tk_+x+\theta _+)+\mathrm{cos}(\omega _{}t+k_{}x+\theta _{})$$ $`(5)`$ where $$\begin{array}{ccc}\hfill \omega _\pm & =\omega _z\pm \omega _B\hfill & (6a)\hfill \\ \hfill k_\pm & =k_z\pm k_B\hfill & (6b)\hfill \end{array}$$ and $$\begin{array}{ccc}\hfill \omega _z=\gamma \omega _C,& \omega _B=\gamma \beta \omega _C\hfill & (7a)\hfill \\ \hfill k_z=\gamma k_C,& k_B=\gamma \beta k_C.\hfill & (7b)\hfill \end{array}$$ The respective random phases associated with each one of these independent ZPF wavetrains are $`\theta _{+,}`$. After some algebra one obtains that the oppositely moving wavetrains appear in the form $$\varphi =2\mathrm{cos}(\omega _ztk_Bx+\theta _1)\mathrm{cos}(\omega _Btk_zx+\theta _2)$$ $`(8)`$ where $`\theta _{1,2}`$ are again two independent random phases $`\theta _{1,2}=\frac{1}{2}(\theta _+\pm \theta _{})`$. Observe that for fixed $`x`$, the rapidly oscillating “carrier” of frequency $`\omega _z`$ is modulated by the slowly varying envelope function in frequency $`\omega _B`$. And vice versa observe that at a given $`t`$ the “carrier” in space appears to have a relatively large wave number $`k_z`$ which is modulated by the envelope of much smaller wave number $`k_B`$. Hence both timewise at a fixed point in space and spacewise at a given time, there appears a carrier that is modulated by a much broader wave of dimension corresponding to the de Broglie time $`t_B=2\pi /\omega _B`$, or equivalently, the de Broglie wavelength $`\lambda _B=2\pi /k_B`$. This result may be generalized to include ZPF radiation from all other directions, as may be found in the monograph of de la Peña and Cetto . They conclude by stating: “The foregoing discussion assigns a physical meaning to de Broglie’s wave: it is the modulation of the wave formed by the Lorentz-transformed, Doppler-shifted superposition of the whole set of random stationary electromagnetic waves of frequency $`\omega _C`$ with which the electron interacts selectively.” Another way of looking at the spatial modulation is in terms of the wave function. Since $$\frac{\omega _C\gamma v}{c^2}=\frac{m_0\gamma v}{\mathrm{}}=\frac{p}{\mathrm{}}$$ $`(9)`$ this spatial modulation is exactly the $`e^{ipx/\mathrm{}}`$ wave function of a freely moving particle satisfying the Schrödinger equation. The same argument has been made by Hunter . In such a view the quantum wave function of a moving free particle becomes a “beat frequency” produced by the relative motion of the observer with respect to the particle and its oscillating charge. It thus appears that a simple model of a particle as a ZPF-driven oscillating charge with a resonance at its Compton frequency may simultaneously offer insight into the nature of inertial mass, i.e. into rest inertial mass and its relativistic extension, the Einstein-de Broglie formula and into its associated wave function involving the de Broglie wavelength of a moving particle. If the de Broglie oscillation is indeed driven by the ZPF, then it is a form of Schrödinger’s Zitterbewegung. Moreover there is a substantial literature attempting to associate spin with Zitterbewegung tracing back to the work of Schrödinger ; see for example Huang and Barut and Zanghi . In the context of ascribing the Zitterbewegung to the fluctuations produced by the ZPF, it has been proposed that spin may be traced back to the (circular) polarization of the electromagnetic field, i.e. particle spin may derive from the spin of photons in the electromagnetic quantum vacuum . It is well known, in ordinary quantum theory, that the introduction of $`\mathrm{}`$ into the ZPF energy density spectrum $`\rho _{ZP}(\omega )`$ of eqn. (2) is made via the harmonic-oscillators-quantization of the electromagnetic modes and that this introduction of $`\mathrm{}`$ is totally independent from the simultaneous introduction of $`\mathrm{}`$ into the particle spin. The idea expounded herein points however towards a connection between the $`\mathrm{}`$ in $`\rho _{ZP}(\omega )`$ and the $`\mathrm{}`$ in the spin of the electron. In spite of a suggestive preliminary proposal, an exact detailed model of this connection remains to be developed . Finally, although we amply acknowledge that other vacuum fields besides the electromagnetic do contribute to inertia, e.g. see , no attempt has been made within the context of the present work to explore that extension. A standard procedure of conventional quantum physics is to limit the describable elements of reality to be directly measurable, i.e. the so-called physical observables. We apply this philosophy here by pointing out that inertial mass itself does not qualify as an observable. Notions such as acceleration, force, energy and electromagnetic fields constitute proper observables; inertial mass does not. We propose that the inertial mass parameter can be accounted for in terms of the forces and energies associated with the electrodynamics of the ZPF. Acknowledgement This work is supported by NASA research grant NASW-5050. References A. Rueda and B. Haisch, Phys. Lett. A 240, 115 (1998). A. Rueda and B. Haisch, Foundation of Phys. 28, 1057 (1998). B. Haisch, A. Rueda, and H. E. Puthoff, H. E., Phys. Rev. A 48, 678 (1994). G. Hunter, in The Present Status of the Quantum Theory of Light, S. Jeffers et al. (eds.), (Kluwer), pp. 37–44 (1997). L. de la Peña and M. Cetto, The Quantum Dice: An Introduction to Stochastic Electrodynamics, (Kluwer Acad. Publ.), chap. 12 (1996). A. F. Kracklauer, Physics Essays 5, 226 (1992). E. Schrödinger, Sitzungsbericht preuss. Akad. Wiss., Phys. Math. Kl. 24, 418 (1930). K. Huang, Am. J. Physics 20, 479 (1952). A. O. Barut and N. Zanghi, Phys. Rev. Lett., 52, 2009 (1984). A. Rueda, Foundations of Phys. Letts. 6, No. 1, 75 (1993); 6, No. 2, 139 (1993). J.-P. Vigier, Foundations of Phys. 25, No. 10, 1461 (1995).
no-problem/9906/hep-ph9906385.html
ar5iv
text
# Constraints on the R-parity Violating Couplings from 𝐵^±→𝑙^±⁢𝜈 Decays ## acknowledgements This work was supported in part by KOSEF postdoctoral program (S.B) and the KAIST Center for Theoretical Physics and Chemistry (Y.G.K).
no-problem/9906/hep-lat9906033.html
ar5iv
text
# Non-Perturbative Fine-Tuning in Approximately Supersymmetric Models ## Abstract We present two Fermi-Bose models with an approximate supersymmetry and which can be solved numerically with great accuracy using the renormalization group method. The bosonic parts of these models consist in Dyson’s hierarchical model with one and two scalar components respectively. We discuss the question of the perturbative cancellations of divergences and compare with the non-perturbative fine-tunings necessary to keep the renormalized scalar mass small in cut-off units. We show evidence for non-perturbative cancellations of quantum corrections, however, we were not able to achieve exact cancellations without fine-tuning. preprint: U. of Iowa 99-2502 The fact that the bare parameters of a scalar field theory require a fine-tuning in order to keep the renormalized mass small in cut-off units is usually regarded as an argument against fundamental scalars. A possible resolution of this inelegant feature consists in adding degrees of freedom in such a way that the quantum fluctuations cancel, making small scalar masses a more natural outcome. From a non-perturbative renormalization group analysis, a relevant direction is necessary in order to describe massive particles, making the fine-tuning process unavoidable. This situation can be seen very simply in the case of a free scalar theory with a cut-off. In this example, the fact that the bare mass ($`m_B`$) has to be small (in cut-off units) in order to get a small physical mass ($`m_R`$) is obvious since these quantities are identical. On the other hand in an interacting scalar theory, we usually need to take $`m_B^2`$ negative and large in absolute value and also to adjust many digits of this quantity in order to get a small $`m_R`$. The difference between the two situations is that in the second case, there is no bare quantity which controls the size of $`m_R`$. One would like to understand under which circumstances the inclusion of fermions allows us to obtain a small $`m_R`$ whenever we choose a small $`m_B^2`$. There are known four-dimensional examples where one can cancel the perturbative quadratic divergences by imposing simple relations between the Yukawa couplings and the scalar quartic couplings. However, it is not clear that there exists a non-perturbative regularization which fully preserves the perturbative naturalness. In the following, we present two models where there is an equal number of fermions and bosons and which can be solved non-perturbatively with great accuracy. The bosonic part of these model is a Dyson’s hierarchical model. In this model, the renormalization group transformation maps the local measure into another local measure. The price to pay for this simplifying feature is that the kinetic term is not ultra-local. The free action for $`N`$ massless scalar fields $`\varphi _x^{(i)}`$ reads $$S_B^{free}=\frac{1}{2}\underset{x,y,i}{}\varphi _x^{(i)}D_{xy}^2\varphi _y^{(i)},$$ (1) where $`x`$ and $`y`$ run over the sites and $`i`$ from 1 to $`N`$. The explicit form of $`D_{xy}^2`$ is given below in Eq. (9). The action for free massless fermions reads $$S_F^{free}=\underset{x,y,i}{}\overline{\psi }_x^{(i)}D_{xy}\psi _y^{(i)},$$ (2) where the $`\psi _x^{(i)}`$ and $`\overline{\psi }_x^{(i)}`$ are Grassmann numbers integrated with a measure $$\underset{x,i}{}d\psi _x^{(i)}d\overline{\psi }_x^{(i)}.$$ (3) As indicated by the notation, we have $$D_{xy}^2=\underset{z}{}D_{xz}D_{zy}.$$ (4) The free action $`S_B^{free}+S_F^{free}`$ is invariant at first order under the transformation $`\delta \varphi _x^{(i)}=ϵ\overline{\psi }_x^{(i)}+\psi _x^{(i)}\overline{ϵ}`$ (5) $`\delta \psi _x^{(i)}=ϵ{\displaystyle \underset{x}{}}D_{xy}\varphi _y^{(i)}`$ (6) $`\delta \overline{\psi }_x^{(i)}=\overline{ϵ}{\displaystyle \underset{x}{}}D_{xy}\varphi _y^{(i)}.`$ (7) The $`ϵ`$ and $`\overline{ϵ}`$ are Grassmann numbers. Integration by part or Leibnitz’s rule cannot be used for $`D_{xy}`$ and the order $`ϵ\overline{ϵ}`$ variations do not cancel. We now give the explicit form of $`D_{xy}^2`$ at finite volume. For a hierarchical model with $`2^{n_{max}}`$ sites, we label the sites with $`n_{max}`$ indices $`x_{n_{max}},\mathrm{},x_1`$, each index being 0 or 1. In order to understand this notation, one can divide the $`2^{n_{max}}`$ sites into two blocks, each containing $`2^{n_{max}1}`$ sites. If $`x_{n_{max}}=0`$, the site is in the first box, if $`x_{n_{max}}=1`$, the site is in the second box. Repeating this procedure $`n`$ times (for the two boxes, their respective two sub-boxes, etc… ), we obtain an unambiguous labeling for each of the sites. With these notations, $$S_B^{free}=\beta _B\underset{n=1}{\overset{n_{max}}{}}(\frac{c_B}{4})^n\underset{x_{n_{max}},\mathrm{},x_{n+1},i}{}(\underset{x_n,\mathrm{},x_1}{}\varphi _{(x_{n_{max}},\mathrm{}x_1)}^{(i)})^2+\frac{\beta _Bc_B}{2c_B}\underset{x_{n_{max}},\mathrm{},x_{n+1},i}{}(\varphi _{(x_{n_{max}},\mathrm{}x_1)}^{(i)})^2.$$ (9) The index $`n`$ corresponds to the interaction of the total field in blocks of size $`2^n`$. The constant $`c_B=2^{12/D}`$ is a free parameter which controls the decay of the iterations with the size of the boxes and can be adjusted in order to mimic a $`D`$-dimensional model. Similarly the free massless fermionic action reads $`S_F^{free}=\beta _F{\displaystyle \underset{n=1}{\overset{n_{max}}{}}}({\displaystyle \frac{c_F}{4}})^n{\displaystyle \underset{x_{n_{max}},\mathrm{},x_{n+1},i}{}}({\displaystyle \underset{x_n,\mathrm{},x_1}{}}\overline{\psi }_{(x_{n_{max}},\mathrm{}x_1)}^{(i)})({\displaystyle \underset{x_n,\mathrm{},x_1}{}}\psi _{(x_{n_{max}},\mathrm{}x_1)}^{(i)})`$ (10) $`+{\displaystyle \frac{\beta _Fc_F}{2c_F}}{\displaystyle \underset{x_{n_{max}},\mathrm{},x_{n+1},i}{}}\overline{\psi }_{(x_{n_{max}},\mathrm{}x_1)}^{(i)}\psi _{(x_{n_{max}},\mathrm{}x_1)}^{(i)},`$ (11) with $`c_F=2^{11/D}.`$ Using the techniques explained in , one can show that the fermionic operator is the square root of the bosonic operator (see Eq. (4)) provided that $$\frac{\beta _Fc_F}{2c_F}=(\frac{\beta _Bc_B}{2c_B})^{\frac{1}{2}}$$ (12) We now introduce local interactions. The Grassmann nature of the fermionic fields restricts severely the type of interactions allowed. For instance, for one flavor ($`N=1`$), the most general local measure is $$𝒲(\varphi ,\psi ,\overline{\psi })=W(\varphi )+\psi \overline{\psi }A(\varphi )$$ (13) For convenience, we will always reabsorb the second term of Eqs. (9) and (11) which are local, in the local measure. In the following calculations, $`W(\varphi )`$ will take the Landau-Ginzburg (LG) form: $$W(\varphi )e^{((\frac{\beta _Bc_B}{2c_B})+\frac{1}{2}m_B^2\varphi ^2+\lambda _B\varphi ^4)}.$$ (14) If the two functions $`W`$ and $`A`$ are proportional, the fermionic degrees of freedom decouple. The renormalization group transformation takes the form $`W`$ $``$ $`2AW`$ (15) $`A`$ $``$ $`2\beta _FAW+({\displaystyle \frac{4}{c_F}})WW`$ (16) where the $``$ operation means a convolution, a multiplication by an exponential and a rescaling of the new field. More precisely $$AB(\varphi )e^{\frac{\beta _B}{2}(\varphi ^2)}𝑑\varphi ^{}A(\frac{(\varphi 2c_B^{\frac{1}{2}}\varphi ^{})}{2})B(\frac{(\varphi 2c_B^{\frac{1}{2}}+\varphi ^{})}{2}),$$ (17) The introduction of a Yukawa coupling can be achieved by allowing a linear term in $`A(\varphi )`$. Such a term breaks explicitly the $`Z_2`$ symmetry of the LG measure. Such a model is characterized by a sudden change from the symmetric phase behavior to the broken phase behavior followed by unexpectedly long low-temperature “shoulders” (see for an explanation of this terminology). A richer behavior is observed in the case of the two flavors ($`i=1,2`$) models. In the following we have restricted our investigation to the type of bilinear coupling appearing in the Wess-Zumino model, namely $`𝒲(\varphi ^{(i)},\psi ^{(i)},\overline{\psi }^{(i)})=W(\varphi ^{(i)})+A(\varphi ^{(i)})(\overline{\psi }^{(1)}\psi ^{(1)}+\overline{\psi }^{(2)}\psi ^{(2)})+`$ (18) $`B(\varphi ^{(i)})\psi ^{(1)}\psi ^{(2)}B^{}(\varphi ^{(i)})\overline{\psi }^{(1)}\overline{\psi }^{(2)}+T(\varphi ^{(i)})\overline{\psi }^{(1)}\psi ^{(1)}\overline{\psi }^{(2)}\psi ^{(2)}.`$ (19) For convenience, we again absorb the local parts of Eqs. (9) and (11). This is not the most general measure, however it closes under the renormalization group transformation which takes the form $`W`$ $``$ $`(WT+AA+BB^{})W^{}`$ (20) $`A`$ $``$ $`\beta _FW^{}+{\displaystyle \frac{4}{c_F}}AT`$ (21) $`B`$ $``$ $`{\displaystyle \frac{4}{c_F}}BT`$ (22) $`T`$ $``$ $`{\displaystyle \frac{8}{c_F^2}}TT+\beta _F{\displaystyle \frac{8}{c_F}}AT+(\beta _F)^2W^{}.`$ (23) In addition we will impose that the function $`B`$ have the following form: $$B(\varphi ^{(i)})=(\varphi ^{(1)}+i\varphi ^{(2)})P((\varphi ^{(1)})^2+(\varphi ^{(2)})^2),$$ (25) while $`W`$, $`A`$ and $`T`$ are $`O(2)`$-invariant. The model is then invariant under the R-symmetry $`(\varphi ^{(1)}+i\varphi ^{(2)})`$ $``$ $`e^{i\theta }(\varphi ^{(1)}+i\varphi ^{(2)})`$ (26) $`\psi ^{(j)}`$ $``$ $`e^{i\frac{\theta }{2}}\psi ^{(j)}`$ (27) $`\overline{\psi }^{(j)}`$ $``$ $`e^{i\frac{\theta }{2}}\overline{\psi }^{(j)}.`$ (28) We now present three numerical calculations performed with the second model. In all cases we will set $`D=4`$ in $`c_B`$ and $`c_F`$. In the following, we have chosen the value of $`\beta _B`$ and $`\beta _F`$ in such a way that $$\frac{\beta _Fc_F}{2c_F}=(\frac{\beta _Bc_B}{2c_B})^{\frac{1}{2}}=1,$$ (29) in order to make the perturbative expansion more similar to usual Feynman diagram’s calculations. First, we consider the case where the fermions decouple from the bosons. $`W`$ takes a LG form $$W(\varphi )e^{(((\frac{\beta _Bc_B}{2c_B})+\frac{1}{2}m_B^2)_i(\varphi ^{(i)})^2+\lambda _B(_i(\varphi ^{(i)})^2)^2)}.$$ (30) The value of $`m_R^2`$, defined as the inverse of the zero-momentum two-point function, is shown in Fig. 1. as a function of $`m_B^2`$. These quantities are expressed in cut-off units. For reference we have also displayed the one-loop perturbative result and the trivial gaussian result. One sees that the scalar self-interaction moves $`m_R^2`$ up and $`m_R^20.2`$ when $`m_B^2`$ goes to zero. The one-loop result is quite good when $`m_R^2`$ is large enough but deteriorates when this quantity becomes smaller. In the second calculation, we consider a bosonic model with a bare mass $`m_B`$ and $`\lambda _B=0`$ coupled to a fermion with the following couplings: $`A`$ $`=`$ $`(1m_B)W`$ (31) $`P`$ $`=`$ $`g_yW`$ (32) $`T`$ $`=`$ $`((1m_B)^2+g_y^2((\varphi ^{(1)})^2+(\varphi ^{(2)})^2))W.`$ (33) The results are shown in Fig. 2 for $`g_y=\sqrt{0.08}0.28`$. One sees that the Yukawa coupling moves $`m_R^2`$ down. For $`m_B^20.094`$, $`m_R`$ becomes 0 and for smaller of $`m_B^2`$, we enter the broken symmetry phase. We have then repeated the second calculation with $`\lambda _B=0.01`$ instead of 0. In perturbation theory, the one-loop quadratic divergence cancel when $`m_B=0`$ and $$8\lambda _B=g_y^2,$$ (34) which justifies our choice of coupling constant. The results are shown in Fig. 3. One sees that the Yukawa coupling in part cancels the effects of the scalar self-interaction, however, the cancellation is not as good as in the one-loop formula where $`m_R`$ goes to zero when $`m_B^2`$ goes to zero. Instead, we found numerically that $`m_R^20.044`$ when $`m_B^2`$ goes to zero. A summary of the three numerical results is shown in Fig. 4. It is possible to fine-tune $`g_y`$ in order to get $`m_R=0`$. An example is shown in Fig. 5 for $`\lambda _B=0.01`$ and $`m_B^2=0.01`$. We see that there exist a critical value of $`g_y`$ which is approximately 0.46 and where $`m_R`$ becomes 0. For larger values of $`g_y`$, we enter the symmetry broken phase. An essentially similar figure is obtained for $`m_B^2=0`$. In both cases, the exact critical value of $`g_y`$ is about 50 percent larger than the perturbative one. In conclusion, we have shown that the idea of canceling the quantum correction inspired by perturbation theory have qualitatively a non-perturbative counterpart. However, we have not found a way to make this cancellation very accurate or exact without fine-tuning. We thank the Institut de Physique Theorique of Louvain-la-Neuve, the CERN theory division and the Aspen Center for Physics where part of this work was completed and B. Oktay for valuable conversations. This research was supported in part by the Department of Energy under Contract No. FG02-91ER40664.
no-problem/9906/hep-ex9906010.html
ar5iv
text
# Parton energy loss limits and shadowing in Drell-Yan dimuon production ## Abstract A precise measurement of the ratios of the Drell-Yan cross section per nucleon for an 800 GeV/$`c`$ proton beam incident on Be, Fe and W targets is reported. The behavior of the Drell-Yan ratios at small target parton momentum fraction is well described by an existing fit to the shadowing observed in deep-inelastic scattering. The cross section ratios as a function of the incident parton momentum fraction set tight limits on the energy loss of quarks passing through a cold nucleus. The Drell-Yan process, where a beam quark (antiquark) fuses with a target antiquark (quark) producing a muon pair, can be used to study the interactions of fast partons penetrating through cold nuclei. Only initial state interactions are important in Drell-Yan since the dimuon in the final state does not interact strongly with the partons in the medium. This makes Drell-Yan scattering an ideal tool to study energy loss of fast quarks in nuclear matter by comparing the observed yields from a range of nuclear targets. The dynamics of fast parton energy loss in nuclear matter is the subject of considerable theoretical interest and has significant implications for the physics of relativisitic heavy ion collisions. Drell-Yan scattering is closely related to deep-inelastic scattering (DIS) of leptons, but unlike DIS it can be used specifically to probe antiquark contributions in target parton distributions. When DIS on nuclei occurs at $`x<0.08`$, where $`x`$ is the parton momentum fraction, the cross section per nucleon decreases with increasing nuclear number $`A`$ due to shadowing . Shadowing should also occur in Drell-Yan dimuon production at small $`x_2`$, the momentum fraction of the target parton, and theoretical calculations indicate that shadowing in the two reactions has a common origin . This should be particularly apparent at $`x<0.06`$ where DIS on nuclei, like Drell-Yan, is dominated by scattering off sea quarks. Fermilab Experiment 866 (E866) measured the nuclear dependence of Drell-Yan dimuon production by 800 GeV/$`c`$ protons on Be, Fe and W targets at larger values of $`x_1`$, the momentum fraction of the beam parton, larger values of $`x_F`$ ($`x_1x_2`$), and smaller values of $`x_2`$ than reached by the previous experiment, Fermilab E772 . The extended kinematic coverage of E866 significantly increases its sensitivity to parton energy loss and shadowing. This Letter reports the results. The experiment used the same 3-dipole magnet spectrometer that was described in . An 800 GeV/$`c`$ extracted proton beam averaging $`3\times 10^{11}`$ protons per 20 s spill bombarded one of three solid targets or an empty target frame. The Be, Fe and W targets were 9.4% – 19% of an interaction length thick. Their relative thicknesses were chosen carefully to match the background rates present in the spectrometer. The targets were located far upstream of the main dipole magnet to optimize the acceptance at low $`x_2`$ and large $`x_F`$. During a cycle, 2 beam spills were taken on each target and 1 spill was collected on an empty location. After passing through a target, the remaining beam was intercepted by a copper beam dump, which was followed by a thick absorber that removed hadrons produced in the target and the dump. This ensured that only muons traversed the spectrometer’s detectors, which consisted of four tracking stations and a momentum analyzing magnet. The trigger required a pair of triple hodoscope coincidences having the topology of a muon pair from the target. Typically 1400 triggers per second were recorded with an average live time of 93%. Over 130,000 muon pairs with dimuon mass in the range $`4.0<M<8.4`$ GeV/$`c^2`$ survived the data cuts. Random pairs were subtracted from the data using simulated random dimuons constructed by mixing single muon tracks that were obtained simultaneously from prescaled single muon triggers. The normalization factor for the random correction was evaluated by comparing to the observed rate of same-charge muon pairs in the data. The random rate was 9% of the real events, and it occurred primarily at low dimuon mass and high $`p_T`$. A residual 3% background from the empty target was also subtracted. The solid targets intercepted the beam at nearly the same $`z`$ location, making differences in dimuon acceptance negligible. The overall systematic normalization uncertainty in the cross section ratios reported here is 1%. The Drell-Yan events obtained by E866 extend over the ranges $`0.01<x_2<0.12`$ and $`0.21<x_1<0.95`$, with $`x_2`$ = 0.038 and $`x_1`$ = 0.46. They also cover the range $`0.13<x_F<0.93`$ and provide good $`p_T`$ coverage to 4 GeV/$`c`$. Ratios of the cross section per nucleon for Fe to Be and W to Be versus dimuon mass, $`x_2`$, $`x_F`$ and $`x_1`$ are shown in Fig. 1, along with similar results from E772 for Fe to C and W to C. The figure shows very good agreement between the experiments for the cross section ratios versus $`x_2`$. The agreement versus other variables is also quite good, given the much smaller $`x_2`$ and, hence, increased shadowing of the present data. Note that the $`A`$ dependence observed here and in E772 implies the change in the cross section ratios in Fig. 1 due to the choice of Be versus C is small compared to the effect of the difference in $`x_2`$. The reduction in the cross section per nucleon on the heavy targets, characteristic of shadowing, is clearly apparent at small $`x_2`$. A similar reduction, part of which could be related to incident parton energy loss, is apparent at large $`x_F`$ and $`x_1`$. However, it is important to recognize that the spectrometer acceptance coupled to the intrinsic Drell-Yan cross section leads to a strong anti-correlation between $`x_2`$ and $`x_F`$ for the observed events. Therefore, the events that show the cross section reduction at large $`x_F`$ and $`x_1`$ are in general the same events that appear in the shadowing region. In order to identify the contributions from shadowing, Fig. 1 also shows the predicted cross section ratios, integrated over the hidden variables, from leading-order Drell-Yan calculations using the code EKS98 together with the MRST parton distribution functions . Essentially identical results are obtained using CTEQ5M or CTEQ5L parton distributions instead. EKS98 describes the ratios $`f_A(x,Q^2)/f_p(x,Q^2)`$ of the various quark flavors in nucleus $`A`$, compared to those in the proton. It has been tuned to fit the shadowing observed in DIS and E772 while conserving baryon number and momentum. EKS98 provides an unbiased way to separate the effects of shadowing and energy loss in the present data because it uses a single shadowing function to describe the nuclear dependence of all sea quark distributions at its initial scale $`Q_{0}^{}{}_{}{}^{2}`$ = 2.25 GeV<sup>2</sup> and only DIS results are used to constrain that function for $`x0.1`$. The shape and magnitude of the ratios versus $`x_2`$ are well reproduced by the shadowing predictions. Most of the $`A`$ dependence observed in the ratios versus mass, $`x_F`$ and $`x_1`$ can also be explained by shadowing at small $`x_2`$. As a further test of the shadowing parametrization, the events in Fig. 1 have been separated into two sets at the median $`x_2`$ value. The cross section ratios differ by up to 6% when the events at low and high $`x_2`$ – but at the same mass, $`x_F`$ or $`x_1`$ – are observed separately. EKS98 also describes these differences well. This is the first experimental demonstration that the shadowing observed in Drell-Yan and DIS is quantitatively similar. Figure 2 shows the measured ratios of the Drell-Yan cross section per nucleon as a function of $`p_T`$. Unlike the longitudinal variables that have strongly correlated acceptances, the $`p_T`$ acceptance in E866 is nearly independent of the other kinematic variables, so the shadowing calculations predict essentially constant cross section ratios versus $`p_T`$, as shown by the smooth curves in the figure. The data demonstrate a clear $`p_T`$ dependence which must have an independent origin. The slight reduction in the cross section per nucleon observed for heavy nuclei at small $`p_T`$, coupled with the increase in the cross section per nucleon at large $`p_T`$, is characteristic of multiple scattering of the incident partons as they traverse the nucleus. The $`x_1`$ dependence of the cross section ratios provides the best direct measure of the energy loss of the incident quarks in the nuclear medium. Table I gives the ratios of the measured cross section per nucleon as a function of $`x_1`$, and the mean values of $`x_1`$, $`x_2`$ and dimuon mass for each bin. However, as shown above, shadowing at small $`x_2`$ explains a substantial fraction of the apparent variation in the cross section ratios versus $`x_1`$. This must be removed before one can isolate a nuclear dependence due to energy loss. Figure 3 shows the same cross section ratios versus $`x_1`$ as given in Table I, but corrected for shadowing by weighting each event with the calculated ratio of the Drell-Yan cross sections per nucleon for deuterium and nucleus $`A`$ at the same $`(x_1,x_2,Q^2)`$, using EKS98 and MRST. Several groups have studied energy loss of partons in nuclei. Their results can be expressed in terms of the average change in the incident parton momentum fraction, $`\mathrm{\Delta }x_1`$, as a function of target nucleus. Gavin and Milana analyzed the E772 Drell-Yan data for energy loss based on the parametrization $$\mathrm{\Delta }x_1=\kappa _1x_1A^{1/3},$$ (1) where the factor $`\kappa _1`$ may have a $`Q^2`$ dependence. They based the form (1) on an analogy to the transverse spin asymmetry in direct photon production. From a comparison to the $`x_F`$ dependence seen by E772 and neglecting shadowing, they concluded that the fractional energy loss of quarks passing through nuclei is $``$ 0.4%/fm. Recently, other groups have assumed equivalent formulations with even more rapid energy loss . Brodsky and Hoyer argued that the energy loss found by Gavin and Milana was too large since the time scale for QCD bremsstrahlung was too short to allow for multiple contributions to the energy loss. Brodsky and Hoyer used an analogy to the photon bremsstrahlung process to obtain a form for gluon radiation leading to an initial parton energy loss $$\mathrm{\Delta }x_1\frac{\kappa _2}{s}A^{1/3}.$$ (2) They found an upper limit for the gluon radiation from the uncertainty relation. They also noted that elastic scattering should make a similar contribution to the energy loss. Overall, they concluded that energetic partons should lose $``$ 0.5 GeV/fm in nuclei. The formulation developed by Brodsky and Hoyer was extended by Baier et al. They found that the energy loss of sufficiently energetic partons depends on a characteristic length and the broadening of the squared transverse momentum of the parton. For finite nuclei, both factors vary as $`A^{1/3}`$, so Baier et al. predict $$\mathrm{\Delta }x_1\frac{\kappa _3}{s}A^{2/3}.$$ (3) Baier et al. predict that energy loss may be different in hot and cold nuclear matter and that there could a large coherent effect in relativistic heavy ion collisions. In contrast, tight limits have been placed on energy loss in S+S and Pb+Pb collisions , but the parton momenta relative to the hadronic medium were very much smaller in those cases than in the present experiment, making a direct comparison difficult. Given these energy loss expressions, it is possible to obtain empirical values for the $`\kappa `$’s by performing simultaneous fits to the Fe/Be and W/Be Drell-Yan cross section ratios versus $`x_1`$ in Fig. 3. We assume that $`\kappa _1`$ is $`Q^2`$-independent when using form (1) and that our incident quarks are sufficiently energetic when using form (3). Curves corresponding to the fits are included in Fig. 3. When assuming the form (1), we find $`\kappa _1=0.0004\pm 0.0009`$. This implies that the observed fractional energy loss of the incident quarks is $`<`$ 0.14%/fm. For the energy loss forms (2) and (3), the best fits imply essentially zero energy loss. We find the $`1\sigma `$ upper limits to be $`\kappa _2<0.75`$ GeV<sup>2</sup> and $`\kappa _3<0.10`$ GeV<sup>2</sup>. The $`\kappa _2`$ limit indicates that the incident quarks lose energy at a constant rate of $`<`$ 0.44 GeV/fm. The $`\kappa _3`$ limit implies that the observed energy loss of the incident quarks within the model of Baier et al. is $`\mathrm{\Delta }E<0.046`$ GeV/fm<sup>2</sup> $`\times `$ $`L^2`$, where $`L`$ is the quark propagation length through the nucleus. This is very close to the lower value given by Baier et al. for cold nuclear matter . In all three cases, the quoted errors include both statistics and the overall normalization uncertainty, with the latter dominating. One can also obtain an indirect estimate of the energy loss due to gluon radiation in the model of Baier et al. from the broadening of the $`p_{T}^{}{}_{}{}^{2}`$ as the incident quark passes through the nucleus, as shown in Fig. 2. However, such an analysis involves two significant difficulties. While the Drell-Yan cross section is very small for $`p_T`$ beyond 4 GeV/$`c`$, the change in $`p_{T}^{}{}_{}{}^{2}`$ from nucleus to nucleus, being the small difference of large numbers, is quite sensitive to the yield at very large $`p_T`$ where our acceptance is poor and the random background becomes significant. Meanwhile, the $`p_T`$ dependence of the ratio of the Drell-Yan cross sections per nucleon on hydrogen and deuterium shows possible evidence for a change in the reaction mechanism at $`p_T3`$ GeV/$`c`$ , complicating interpretation of the large $`p_T`$ ratio data in Fig. 2. Further analysis of the $`p_T`$ dependence of the Drell-Yan cross section will be presented in a future publication. In summary, this Letter reports a measurement of the ratios of the Drell-Yan cross section per nucleon for Fe to Be and W to Be. Nuclear shadowing is found to be important in the small $`x_2`$ domain. For sufficiently small $`x_2`$, the shadowing observed in Drell-Yan has been demonstrated to be quantitatively similar to that in DIS. Subsequently, a correction for shadowing has been applied. The cross section ratios versus $`x_1`$ provide direct upper limits on the energy loss of the incoming parton as it traverses a cold nucleus that are tighter than previous constraints. Shadowing and initial state energy loss are processes that occur in both Drell-Yan production and $`J/\psi `$ formation. Hence, these results should also further the understanding of $`J/\psi `$ production, which is required if it is to be used as a signal for the quark-gluon plasma in relativistic heavy ion collisions. We would like to acknowledge R. Vogt for many useful discussions. We thank the Fermilab Particle Physics, Beams and Computing Divisions for their assistance in performing this experiment. This work was supported in part by the U.S. Department of Energy.
no-problem/9906/hep-th9906052.html
ar5iv
text
# String field theory Hamiltonians from Yang-Mills theories ## Abstract Marchesini showed that the Fokker–Planck Hamiltonian for Yang-Mills theories is the loop operator. Jevicki and Rodrigues showed that the Fokker–Planck Hamiltonian of some matrix models coïncides with temporal gauge non-critical string field theory Hamiltonians constructed by Ishibashi and Kawai (and their collaborators). Thus the loop operator for Yang–Mills theory is the temporal gauge Hamiltonian for a noncritical string field theory, in accord with Polyakov’s conjecture. The consistency condition of the string interpretation is the zigzag symmetry emphasized by Polyakov. Several aspects of the noncritical string theory are considered, relating the string field theory Hamiltonian to the worldsheet description. Mandelstam realized the importance of gauge invariant loop observables in Yang–Mills theory. See also . The fact that a string interpretation for Yang–Mills theories is natural is particularly transparent in the loop equation derived by Guerra and collaborators, and rediscovered in . This Schwinger–Dyson equation governs the dynamics of gauge invariant Wilson loops in Yang–Mills theory. There are simple geometric interpretations for the various terms that appear in terms of string propagation and interactions, with the string joining interaction suppressed by a factor of $`1/N^2`$ relative to the string splitting interaction in a normalization natural for the large $`N`$ limit. The loop equation governing Wilson loop expectation values in gauge theories can be written as the expectation value of the action of an operator, the loop operator, acting on Wilson lines. It is a fundamental observation due to Marchesini that the loop operator coïncides with the Fokker–Planck Hamiltonian that appears in the stochastic quantization of the gauge theory. As such, the loop operator plays a central rôle in gauge theories but its significance in, for example, a string field theory equivalent to a Yang–Mills theory has not been elucidated. The contribution of this paper is to point out that the loop operator is precisely the temporal gauge string field theory Hamiltonian. I show that the consistency condition for this identification is Polyakov’s zig–zag symmetry. I explain how the peculiar asymmetry between joining and splitting vertices can be accounted for in a worldsheet description, and I explain how recent conjectures on Yang–Mills theories and their dimensional reductions are related to the explicit identification given here. The explicit connection given in this paper between string theories and gauge theories should be compared to the efforts that have gone into attempted constructions of worldsheet descriptions of first–quantized string propagation in supergravity backgrounds believed to be dual to gauge theories. The conceptual simplicity of the string field theory Hamiltonian is striking because in no other instance in string theory is a string field theory simpler or easier to construct than the first–quantized string theory. The identification of this Hamiltonian can be used to deduce properties of a worldsheet first–quantized description, as I shall show below. It is plausible that the first–quantized description will capture only part of the physics incorporated in the exact string field theory Hamiltonian, just as an expansion in Feynman diagrams misses non–perturbative physics. Concretely, the string field theory that arises from the Yang–Mills theory is a noncritical string theory. The spacetime dimension in which the string propagate is therefore one more than the sum of the number of adjoint scalars and the number of gauge potential components. Thus, e.g. for the $`d=4`$ maximally supersymmetric Yang–Mills theory, the spacetime dimension of the noncritical string theory is 11, which differs from the naïve interpretation of the Maldacena conjecture. However the present construction precisely agrees with the conjectured QCD string in Polyakov’s approach which has been argued on physical grounds to be a noncritical string living in five dimensions. Since the infinite $`N`$ Maldacena conjecture is supported by several precise comparisons with IIB supergravity in ten dimensions, it is important to explain how a noncritical string theory can maintain this agreement. I explain this below by recalling the difference between noncritical string measures and critical string measures following . Furthermore, comparing to the calculation of Dorey, Hollowood, Khoze, Mattis and Vandoren, it is interesting to observe that at finite $`N`$ the instanton calculation does not localize on $`S^5`$ but rather seems to involve an 11–dimensional space. (The connection to the noncritical string given in a previous version of this paper was not mentioned in .) The loop operator is easiest understood in the lattice theory since a precise interpretation requires a cutoff—this cutoff is not necessarily the same as the cutoff used in defining the quantum gauge theory. Define the loop function for a U$`(N)`$ gauge theory on a $`d`$–dimensional lattice $$W(C)\frac{1}{N}\text{Tr}U(C)$$ (1) where $`C`$ is a closed contour on the lattice, and $`U(C)`$ is the path ordered product of link variables along this contour $`U_{\mu _1}(x_1)U_{\mu _2}(x_2)\mathrm{}U_{\mu _n}(x_n).`$ Define an electric field operator appropriate for a $`d+1`$–dimensional theory: $$E_\mu :[E_\mu ^a(x),U_\nu (y)]=T^aU_\nu (y)\delta _{\mu \nu }\delta _{x,y}$$ (2) where $`\mu ,\nu =1,\mathrm{}d,`$ and $`a=1,\mathrm{}N^2,`$ with $`\text{Tr}T^aT^b=N/2.`$ For any lattice action, $`S(U),`$ the stochastic Fokker–Planck Hamiltonian is $$H\frac{2}{N}\underset{x,\mu ,a}{}\mathrm{e}^{S(U)}E_\mu ^a(x)\mathrm{e}^{S(U)}E_\mu ^a(x);$$ (3) while $`H`$ is not Hermitian, by a similarity transformation we can make it Hermitian. There are simple geometric interpretations for the terms that appear in $`H,`$ which are clearer when we write $$H\frac{2}{N}\underset{x,\mu ,a}{}E_\mu ^a(x)E_\mu ^a(x)+[E_\mu ^a(x),S(U)]E_\mu ^a(x).$$ (4) The first term, which is independent of the lattice action, includes terms that correspond to the splitting and joining of strings, with the latter term suppressed relative to the former by a factor of $`N^2.`$ The equilibrium condition for the stochastic field theory correlation function turns out to be the loop equation: $$<HW(C)>=0.$$ (5) Independently, Jevicki and Sakita showed that the large $`N`$ saddle point equation of Yang–Mills theory formulated as a collective field theory is equivalent to the loop equation, as well. The asymmetry in the powers of $`N`$ for joining and splitting interactions is a clue for the string field theory interpretation since in temporal gauge one does find such an asymmetry between string splitting and string joining vertices. This reassignment of powers of $`N`$ is consistent with the usual topological loop counting of string diagrams. The $`S`$–dependent term has operators corresponding to the motion in the space of loops, and tadpole operators that annihilate loops. Thinking of $`S`$ as defining the background configuration of the space of loops, the propagation and annihilation of loops is background dependent, but the interactions of loops are background independent. Notice that the ‘t Hooft coupling $`g_{\mathrm{YM}}^2N`$ appears only in this term. This is essentially another realization of an idea of Horowitz, Lykken, Rohm and Strominger regarding string field theory. In the large $`N`$ limit there are no terms that can join strings. Thus in the large $`N`$ limit, $`H`$ acts as a derivation on products $`_iW(C_i).`$ In a beautiful set of papers, Ishibashi, Kawai, and collaborators have constructed non-critical string field theories in temporal gauge. As an example, consider their string field theory Hamiltonian for the $`c=0`$ model: $$H_{\mathrm{IK}}=_0^{\mathrm{}}\text{d}l_1\text{d}l_2\mathrm{\Psi }_{l_1}^{}\mathrm{\Psi }_{l_2}^{}\mathrm{\Psi }_{l_1+l_2}(l_1+l_2)+_0^{\mathrm{}}\text{d}l\rho (l)\mathrm{\Psi }_l+O(g^2)$$ (6) where $`\mathrm{\Psi }_l`$ annihilates a string of length $`l,`$ and $`[\mathrm{\Psi }_l,\mathrm{\Psi }_l^{}^{}]=\delta (ll^{}).`$ The term of $`O(g^2)`$ is $$g^2_0^{\mathrm{}}\text{d}l_1\text{d}l_2\mathrm{\Psi }_{l_1+l_2}^{}\mathrm{\Psi }_{l_2}\mathrm{\Psi }_{l_1}l_1l_2$$ (7) and describes the merging of strings. $`\mathrm{\Psi }`$ annihilates unmarked loops and $`\mathrm{\Psi }^{}`$ creates marked loops. The string joining and string splitting terms have different powers of $`g,`$ just as there are different powers of $`N^1`$ in the loop joining and splitting terms in the loop operator. This Hamiltonian was first written down for the $`c=1`$ model by Das and Jevicki and it was pointed out subsequently by Moore, Seiberg and Staudacher that it applied as well to $`c<1`$ models with changes in the tadpole term. A connected amplitude with $`b`$ boundaries and $`h`$ handles comes with a factor $`g^{2h2+2b},`$ which is not the usual topological combination. (I have corrected a small error in here.) This can be traced to the fact that the disk amplitude is non-zero and has a factor $`g^0`$ in this formalism. Normalizing connected amplitudes with respect to the conventionally normalized disk amplitudes gives the usual Euler characteristic power $`g^{(22hb)}.`$ In the temporal gauge interpretation, the tadpole term is $`O(1),`$ just as for large $`N`$ Yang–Mills theory the normalized expectation value is $`O(1).`$ One could rescale $`\mathrm{\Psi }`$ and $`\mathrm{\Psi }^{}`$ while preserving the commutation relation to make the joining and the splitting terms of the same order in $`g,`$ but this would change the normalization of the tadpole term as well, making the expectation value of a Wilson loop $`O(N),`$ appropriate for a string disk amplitude. This is not, however, natural from the point of view of the gauge theory, so I will give below another geometric interpretation of the powers of $`g.`$ Jevicki and Rodrigues showed that the Hamiltonian $`H_{\mathrm{IK}}`$ (including the $`O(g^2)`$ term) arises naturally as the double–scaling limit of the $`k=2`$ matrix model Fokker–Planck Hamiltonian. Thus one has a direct map from matrix models to noncritical string field theories in the temporal gauge, via the Fokker–Planck Hamiltonian. Crucial to the string interpretation of the Fokker–Planck Hamiltonian is a consistency check, equivalent to diffeomorphism invariance on the string worldsheet. In the case of the loop operator, it is a generalization of the derivation given in (Sect. 3) that the consistency of the string interpretation is equivalent to the zigzag symmetry of Wilson loops, particularly emphasized by Polyakov. The key point to note here is that two Wilson loops $`L_i,i=1,2`$ (which are the strings in this formalism) may join either by an infinitesimal loop attaching to $`L_1`$ leading to contact with $`L_2`$ or vice versa. For the string interpretation, diffeomorphism invariance on the worldsheet implies that the difference between these two amplitudes should vanish. Subtracting the amplitudes for the two processes, we encounter amplitudes involving insertions of infinitesimal back–tracking loops, which are trivial if and only if the zig–zag symmetry holds. In connection with the zig–zag symmetry, a recent investigation of the loop equation in the AdeS/CFT correspondence has found that the agreement between gauge theory expectations and critical string expectations is regularization dependent. Recalling the analysis of Polyakov and Klebanov, Kogan and Polyakov suggests then that the zig–zag symmetry requires the non–critical Liouville dimension to be taken into account. Thus we reach the main conclusion of this paper: The loop operator of Yang–Mills theory is the temporal gauge string field theory Hamiltonian of a noncritical string theory, provided that the zig–zag symmetry is maintained. This observation suggests that the appropriate place to look for a worldsheet description of the Yang–Mills string is temporal gauge. This gauge is quite different from conformal gauge so much of the usual intuition for first–quantized string theory requires revision. For example, in light–cone string field theory, for example, string splitting and joining interaction vertices are given by mirror image worldsheets, with curvature singularities precisely at the joining/splitting point. These curvature singularities are integrable and precisely lead to the expected power of $`g`$ for both string joining and splitting vertices, since the spacetime dilaton $`\mathrm{\Phi }`$ couples to $`{}_{}{}^{(2)}R.`$ The incoming and outgoing strings lie on curves with vanishing geodesic curvature in light–cone string field theory. In temporal gauge, as mentioned above, the splitting and joining interactions come with different powers of $`g.`$ Since the Euler characteristic of the pants diagram is $`\chi _E=1,`$ how can this be? A simple resolution of this puzzle is to note that only the worldsheet bulk curvature couples to the dilaton, whereas the Euler characteristic is a sum of two terms $${}_{}{}^{(2)}R+\underset{\mathrm{boundaries}}{}_{C_i}\kappa =2\pi \chi _E$$ (8) according to Gauss and Bonnet, with $`\kappa `$ the geodesic curvature of the boundary components. There is a time asymmetry in the definition of temporal gauge, so we can sensibly distinguish between incoming and outgoing strings. Let $`\kappa =2\pi K_\pm `$ for ‘standard’ outgoing/incoming loops respectively, then for the joining vertex we have $$\frac{1}{2\pi }{}_{}{}^{(2)}R=2=\chi _E(\text{pants})K_+2K_{}$$ (9) and for the splitting vertex we have $$\frac{1}{2\pi }{}_{}{}^{(2)}R=0=\chi _E(\text{pants})2K_+K_{}.$$ (10) We find then $$K_+=K_{}=1.$$ (11) This is precisely as it must be since in the standard gluing of worldsheet diagrams in string field theory, the geodesic curvatures of boundaries being identified must cancel for constructing smooth surfaces. As an independent check this predicts that the tadpole diagram (a disk) will have one incoming string giving a contribution $`K_{}=+1`$ to the Euler characteristic and therefore have vanishing integrated bulk curvature, $$\chi _E(\text{disk})K_{}=+11=0=\frac{1}{2\pi }{}_{}{}^{(2)}R$$ (12) implying that the tadpole will be $`O(1)`$, which is exactly correct. As another consistency check, the Euler characteristic of the cylinder is zero,thus incoming and outgoing boundaries must have boundary geodesic curvature that differs only in sign since the propagation term in the loop operator has no factors of $`N`$ and therefore corresponds to a vanishing bulk contribution. The fact that there is nonvanishing geodesic curvature associated with the boundaries in this gauge is related to the fact that the Wilson loops are normalized with factors of $`1/N.`$ The discussion trivially extends to dimensionally reduced and/or supersymmetric gauge theories, with $`A_\mu (x(s))\mathrm{d}x^\mu /\mathrm{d}s`$ replaced by $`\mathrm{\Phi }_i(x(s))_nv^i\delta (ss_n)`$ in the path–ordered exponential for dimensionally reduced directions. This is of interest in light of the conjectures of Banks, Fischler, Shenker and Susskind, and Maldacena, though it needs to emphasized again that these conjectures are associated with critical string theories. In Polyakov’s attempts at constructing string theories equivalent to Yang–Mills theory, on the other hand, the strings that appear are non–critical strings. As such, the identification which has been proposed in this paper is exactly in line with his work, especially noting the zig–zag symmetry’s rôle in a string interpretation. The string field theory that I have identified in this paper is gauge–fixed. This is related to the fact that we have worked entirely in terms of gauge–invariant operations on gauge–invariant Wilson loops. Recollect that residual spacetime gauge transformations in critical string theory appear as conformal transformations on the worldsheet in conformal gauge. Since I have given here a construction of a non–critical string theory from Yang–Mills theory that is different from the standard conjectures, I must explain how the numerous agreements supporting the standard conjectures that appear at string tree–level come about. This is easily done: What is the difference between a critical string amplitude and a noncritical string amplitude? The functional measure for a critical string amplitude has an integration over the conformal mode and a division by the volume of the group of conformal transformations, vol(Conf). These cancel for correlation functions of on–shell vertex operators, and do not cancel if we consider off–shell amplitudes. Even for off–shell amplitudes, on the sphere, there are no moduli and vol(Conf) is a amplitude independent constant. On higher genus surfaces, vol(Conf) depends on the moduli and cannot be factored out. For noncritical strings, there is still an integration over the conformal mode, but no corresponding division by vol(Conf). Thus, if we are interested in tree amplitudes, we can extract critical string tree amplitudes from a subset of the noncritical string tree amplitudes. (Recall that string tree amplitudes do not suffice to reconstruct string loop amplitudes.) This subset corresponds to on–shell string states that have no dependence on the conformal mode. From the Fokker–Planck perspective, it is natural to focus on correlations $$\underset{D\mathrm{}}{lim}0|\mathrm{exp}(DH)W(C_i)|0$$ (13) where $`D`$ is the Fokker–Planck time (the Liouville dimension coördinate), here identified with the worldsheet time coördinate, of Wilson loops $`C_i`$ that satisfy the equilibrium condition $$[H,W(C_i)]=0\text{for each }i.$$ (14) It seems clear that correlations of observables satisfying the equilibrium condition will correspond to amplitudes with truncated external leg propagators, natural for on–shell vertex operators in critical string theory. However, the precise relation between these two sets of objects remains to be computed. Lastly, I mention that supporting evidence for my identification comes from a very recent preprint by Lidsey. In this work it is shown that the critical IIB supergravity backgrounds can be embedded in eleven–dimensional Ricci–flat spaces, including backgrounds with non–trivial Ramond–Ramond fields. This is precisely in accord with the above–mentioned relation between critical strings and non–critical strings at tree level, see especially Klebanov, Kogan and Polyakov. I am grateful to I. Klebanov, G. Lifschytz, J. Polchinski and A. Polyakov for helpful conversations and A. Jevicki for a useful communication. This work was supported in part by NSF grant PHY-9802484.
no-problem/9906/hep-th9906090.html
ar5iv
text
# YITP-99-37hep-th/9906090 June 1999 Branes in type 0/type II duality ## 1 Introduction During the last year, type 0 string theories, which are superstring theories without spacetime supersymmetry, have attracted much attention. They allow us the possibility to analyze non-supersymmetric QCD by means of the brane techniques developed for supersymmetric theories. However, the properties of type 0 theories are not as well known as other supersymmetric string theories, because BPS conditions, which provide a great deal of information concerning supersymmetric theories, are not available. Recently, Bergman and Gaberdiel found a duality between type 0 and type II theories. With this duality, it may become possible to obtain new properties of type 0 theories from knowledge of type II theories. Therefore, it is important to investigate this duality. The purpose of this paper is to find the duality relations between brane configurations in type 0 theories and those in type II theories. ## 2 Type 0/type II duality As is mentioned in , type II theory compactified on $`𝐒^1`$ with monodromy $`(1)^{F_s}`$ is T-dual to type 0 theory on $`𝐒^1`$ with monodromy $`(1)^{F_R}`$, where $`F_s`$ and $`F_R`$ are operators that count spacetime fermions and right-moving worldsheet fermions, respectively.<sup>1</sup><sup>1</sup>1I would like to thank S. Sugimoto for clearing up my misunderstanding of the definition of these operators. In Ref., these compactifications are represented as compactifications on $`𝐒^1/𝐙_2`$. $`𝐙_2`$ is generated by an operator $`S(1)^{F_s}`$ or $`S(1)^{F_R}`$, where $`S`$ is a half shift of $`𝐒^1`$. These two expressions are the same, except that the definitions of the radius of $`𝐒^1`$ are different by a factor $`2`$. (Precisely, duals of type IIA and type IIB are type 0B and type 0A, respectively. We omit the letters A and B in what follows.) Instead of explicitly showing the equivalence of these two theories in terms of the world sheet CFT, we only give the spectra of these theories in Table 1 and Table 2. In these tables, $`n`$ and $`p^9`$ represent the wrapping number and the Kaluza-Klein momentum along the compactified direction. Indeed, according to the results given in these tables, we find that these two spectra are identical if we swap the wrapping number and the Kaluza-Klein momentum. In particular, a type II (type 0) closed string with wrapping number $`1`$ corresponds to a type 0 (type II) Kaluza-Klein mode with momentum $`p=1/(2R)`$. This implies the relations $$2\pi R^{(II)}T^{(II)}=\frac{1}{2R^{(0)}},2\pi R^{(0)}T^{(0)}=\frac{1}{2R^{(II)}},$$ (1) where $`T^{(II)}`$ ($`T^{(0)}`$) and $`R^{(II)}`$ ($`R^{(0)}`$) are the string tension and the compactification radius in type II (type 0) theory. Using these two equations, we obtain $`T^{(II)}=T^{(0)}`$ and $$R^{(II)}R^{(0)}=\frac{l_s^2}{2},$$ (2) where $`l_s`$ is the string length scale defined with the common string tension by $`T^{(II)}=T^{(0)}=1/(2\pi l_s^2)`$. The relation (2) is different from that in usual type II/type II T-duality by the factor $`1/2`$ on the right-hand side. This is the same as in the case of heterotic/heterotic T-duality where, for example, an $`E_8\times E_8`$ heterotic string with wrapping number $`1`$ is dual to a Kaluza-Klein mode with momentum $`p=1/(2R)`$ in the $`(\mathrm{𝟏𝟔},\mathrm{𝟏𝟔})`$ representation of unbroken $`SO(16)\times SO(16)SO(32)`$. Concerning the string coupling constants $`g^{(II)}`$ and $`g^{(0)}`$, we can obtain the following relation by requiring the nine-dimensional Newton’s constants of the two theories to agree: $$\frac{R^{(II)}}{(g^{(II)})^2}=\frac{R^{(0)}}{(g^{(0)})^2}.$$ (3) ## 3 Dualities of D-branes Similarly to the type II/type II T-duality, under the type 0/type II T-duality, wrapped D-branes and unwrapped D-branes are transformed into unwrapped D-branes and wrapped D-branes, respectively. However, the situation is more complicated and interesting. It is known that the tensions of type 0 D-branes are $`1/\sqrt{2}`$ times those of type II D-branes. Namely, with the string coupling constant and the string length scale, the tensions of D-branes are represented as $$T_{Dp}^{(II)}=\frac{1}{(2\pi )^pl_s^{p+1}g^{(II)}},T_{Dp}^{(0)}=\frac{1}{\sqrt{2}}\frac{1}{(2\pi )^pl_s^{p+1}g^{(0)}}.$$ (4) Using (2), (3) and (4), we can obtain $$T_{Dp}^{(II)}=2\times 2\pi R^{(0)}T_{D(p+1)}^{(0)}.$$ (5) This equation implies that an unwrapped type II D-brane is transformed into a wrapped D-brane with wrapping number $`2`$ by the duality. At first sight, this seems strange because it seems that we can decompose the type 0 D-brane into two wrapped D-branes with wrapping number $`1`$. This is, however, impossible on the type II side. This is interpreted as follows. In the type 0 theory, by the monodromy $`()^{F_R}`$ around $`𝐒^1`$, the two R-R fields $`C`$ and $`\overline{C}`$ are transformed into $`C`$ and $`\overline{C}`$, respectively. (The fields $`C`$ and $`\overline{C}`$ are massless fields in the (R$`\pm `$,R$`+`$) and (R$``$,R$``$) sectors, respectively.) In terms of the D-brane charge, this implies that when an electric (magnetic) D-brane goes around $`𝐒^1`$, it is changed into a magnetic (electric) D-brane. (We call D-branes ‘electric’ when the signs of two R-R charges are the same and ‘magnetic’ when they are opposite, although, except for D3-branes, these are not electric nor magnetic in the usual meanings.) Therefore, if the wrapping number of a D-brane is $`1`$, it cannot be closed. To connect two endpoints of a wrapped D-brane, the wrapping number should be even. This is the reason that a type 0 D-brane with wrapping number $`2`$, which is dual to the single type II D-brane, cannot be decomposed (Figure 1). On a stack of $`N`$ type II D-branes, there exist $`U(N)`$ gauge bosons and adjoint fermions, and they make a vector multiplet of ten-dimensional $`𝒩=1`$ supersymmetry. Following the argument above, the type 0 dual of this brane configuration is a stack of $`N`$ electric and $`N`$ magnetic type 0 D-branes. The field theory on the type 0 brane configuration is known to be $`U(N)\times U(N)`$ Yang-Mills theory with fermions belonging to the bi-fundamental representation. However, now, because the electric D-branes and the magnetic D-branes are connected, the gauge group is broken to the diagonal $`U(N)`$. As a result, all massless fields on D-branes belong to the adjoint representation of this $`U(N)`$, and the massless spectrum agrees with that on type II D-branes. Next, let us consider the duality between wrapped type II D-branes and unwrapped type 0 D-branes. Using (2), (3) and (4) we obtain $$2\pi R^{(II)}T_{D(p+1)}^{(II)}=T_{Dp}^{(0)}.$$ (6) This equation implies that the dual of a wrapped type II D($`p+1`$)-brane is an unwrapped type 0 D$`p`$-brane. Is it electric or magnetic? We should remember the fact that because of the existence of the monodromy $`()^{F_R}`$, we cannot globally define ‘electric’ and ‘magnetic’ D-branes. Therefore, we can freely call a certain D-brane electric or magnetic. However, it makes sense to say that two D-branes are of the same type or not. If two type 0 D-branes are of different types, where does the difference come from on the type II side? To obtain an answer to this question, let us consider an electric type 0 unwrapped D-brane. It is dual to a type II wrapped D-brane. On the type 0 side, we can change the D-brane to a magnetic one by moving it around $`𝐒^1`$. On the type II side, this corresponds to a change of the Wilson line $`g=\mathrm{exp}iA_9𝑑x^9`$. In the case of type II/type II duality, to move an unwrapped D-brane round $`𝐒^1`$ corresponds to a change of the Wilson line $`ge^{2\pi i}g`$. Now, however, the size of $`𝐒^1`$ of type 0 theory is $`R^{(0)}=1/(2R^{(II)})`$, unlike $`1/R^{(II)}`$ in type II/type II duality, and this implies that the change of the Wilson line should also be halved. Namely, the change should be $`ge^{i\pi }g`$. Therefore, we can conclude that electric and magnetic unwrapped type 0 D-branes correspond to wrapped type II D-branes with Wilson lines $`g`$ and $`g`$, respectively (Fig.2). We can show that the massless spectra on these type 0 and type II D-branes agree. Let us consider a stack of $`N_1`$ electric and $`N_2`$ magnetic unwrapped type 0 D-branes. The massless fields on this stack are $`U(N_1)\times U(N_2)`$ gauge fields and fermions in the $`(𝐍_1,\overline{𝐍}_2)`$ and $`(\overline{𝐍}_1,𝐍_2)`$ representations. The dual brane configuration is a stack of $`N_1+N_2`$ type II D-branes with the Wilson line $`diag(g\mathrm{𝟏}_{N_1},g\mathrm{𝟏}_{N_2})`$, where $`g`$ is a phase factor which does not affect the spectrum. With the Wilson line, the gauge group $`U(N_1+N_2)`$ is broken to $`U(N_1)\times U(N_2)`$. Fermions in the adjoint representation are decomposed into $`(\mathrm{𝐚𝐝𝐣}_1,\mathrm{𝟏})`$, $`(\mathrm{𝟏},\mathrm{𝐚𝐝𝐣}_2)`$, $`(𝐍_1,\overline{𝐍}_2)`$ and $`(\overline{𝐍}_1,𝐍_2)`$. In addition to the Wilson line acting on these four representations as $`+1`$, $`+1`$, $`1`$ and $`1`$, we should take account of the monodromy $`(1)^{F_s}`$, which reverses the sign of wave functions of all the fermion fields. As a result, fermions in $`(𝐍_1,\overline{𝐍}_2)`$ and $`(\overline{𝐍}_1,𝐍_2)`$ remain massless. This is consistent with the type 0 spectrum. ## 4 Dualities of NS5-branes As D-branes, NS5-branes play important roles in brane constructions of Yang-Mills theories, and they are important to establish the duality relations of NS5-branes. In this section, we give the duals of type 0 and type II NS5-branes. Regardless of the type of string theory in question, we can obtain the tensions of NS5-branes by using the Dirac quantization condition as follows: $$T_{NS5}^{(0)}=\frac{1}{(2\pi )^5l_s^6(g^{(0)})^2},T_{NS5}^{(II)}=\frac{1}{(2\pi )^5l_s^6(g^{(II)})^2}.$$ (7) From (3) and (7), the relation $$2\pi R^{(0)}T_{NS5}^{(0)}=2\pi R^{(II)}T_{NS5}^{(II)},$$ (8) is obtained. This relation strongly suggests that the dual of a wrapped type II NS5-brane is a wrapped type 0 NS5-brane. In type II/type II T-duality, an unwrapped NS5-brane is transformed into a Kaluza-Klein monopole with NUT charge $`1`$. On the other hand, in the case of type 0/type II duality, the dual of an unwrapped NS5-brane should be a Kaluza-Klein monopole with NUT charge $`2`$. The reason for this is as follows. Because the momentum of Kaluza Klein modes along the compactified direction is quantized as $`n/(2R)`$, the minimum electric charge coupled with the $`U(1)`$ gauge field $`g_{\mu 9}`$ is $`1/2`$. Therefore, due to the Dirac quantization condition, the minimum NUT charge is $`2`$. Kaluza-Klein monopoles with NUT charge $`2`$ can be made by dividing the Taub-NUT manifold by $`𝐙_2`$, generated by a half shift of the $`𝐒^1`$ cycle. Because there exists the monodromy $`(1)^{F_s}`$ ($`(1)^{F_R}`$) in type II (type 0) theory, the shift should be accompanied by the operator $`(1)^{F_s}`$ ($`(1)^{F_R}`$). The central region of this Kaluza-Klein monopole is $`𝐑^4/𝐙_2`$. $`𝐙_2`$ is generated by $`𝒫(1)^{F_s}`$ (type II) or $`𝒫(1)^{F_R}`$ (type 0), where $`𝒫`$ is the parity operator which changes the coordinates $`x^\mu `$ on the $`𝐑^4`$ to $`x^\mu `$. Therefore, the origin of the manifold is the $`A_1`$ singularity. Next, let us consider massless spectra on NS5-branes. On every NS5-brane, there exist four massless scalar fields representing the position of the brane. In the dual picture, they appear as zero modes of NS-NS fields on the Kaluza-Klein monopole. In addition to these, on type II NS5-branes, a one-form $`U(1)`$ gauge field (type IIB) or self-dual two-form and zero-form gauge fields (type IIA) exist.<sup>2</sup><sup>2</sup>2We refer to scalar fields which do not correspond to the fluctuation of branes as ‘zero-form gauge fields’ because they are magnetic duals of four-form gauge fields on five-branes. (Here, we focus only on bosonic fields.) These gauge fields are represented as zero modes of R-R fields on the Kaluza-Klein monopoles. Indeed, there is one zero mode of a self-dual two-form field on the Taub-NUT manifold. Therefore, in type 0A theory, a zero mode of an R-R three-form field $`C_{\mu \nu i}`$ (The indices $`\mu `$ and $`\nu `$ label the directions contained in the Taub-NUT manifold, and $`i`$ labels the flat directions parallel to the Kaluza-Klein monopole.) gives a $`U(1)`$ gauge field, while a zero mode of the other R-R three-form field $`\overline{C}_{\mu \nu i}`$ is projected out by $`𝐙_2`$. In the same way, on the Kaluza-Klein monopole in type 0B theory, there are zero modes of the four-form field $`C_{\mu \nu ij}^+`$ and the two-form field $`C_{\mu \nu }`$, and they correspond to the self-dual two-form field and zero-form gauge field on type IIA NS5-branes. Zero modes of $`\overline{C}_{\mu \nu ij}^{}`$ and $`\overline{C}_{\mu \nu }`$ are projected out by $`𝐙_2`$ again. Conversely, by analyzing the type II Kaluza-Klein monopoles, we can find gauge fields on type 0 NS5-branes. On the type IIA Kaluza-Klein monopole, there is a zero mode of the R-R three-form field $`C_{\mu \nu i}`$, and on the type IIB Kaluza-Klein monopole, zero modes of $`C_{\mu \nu ij}^+`$ and $`C_{\mu \nu }`$ exist. Because the monodromy $`(1)^{F_s}`$ acts on bosonic fields trivially, none of these zero modes are projected out. Furthermore, to obtain complete spectra, we should consider the twisted sector at the orbifold singularities. From the (R$``$,R$``$) sectors of twisted closed strings at the singularities, we obtain a one-form gauge field (type IIA) or a zero-form gauge field and an anti-self-dual two-form field (type IIB). Therefore, we conclude that gauge fields on type 0 NS5-branes are doubled and non-chiral, like the R-R fields in the bulk. That is, in addition to four scalar fields corresponding to fluctuations of branes, there exist two zero-form gauge fields and an unconstrained two-form field on a type 0A NS5-brane and two one-form fields on a type 0B NS5-brane.<sup>3</sup><sup>3</sup>3In the previous version of this paper, the argument regarding the twisted sector was missing. The massless spectra on type 0 NS5-branes were first obtained in Ref. using type 0A/type 0B duality. These fields are necessary for the electric and magnetic D-branes to be attached independently on NS5-branes. In the supersymmetric $`𝐑^\mathrm{𝟒}/𝐙_2`$ orbifold case, four massless scalar fields appear in the (NS$`+`$,NS$`+`$) sector of twisted closed strings. They belong to the $`(\mathrm{𝟏},\mathrm{𝟑})+(\mathrm{𝟏},\mathrm{𝟏})`$ of the $`SO(4)=SU(2)_L\times SU(2)_R`$ rotation group and correspond to the (geometric and non-geometric) blow-up modes of the singularity. In the present case, from the twisted (NS$``$,NS$``$) sector, we have four massless scalar fields belonging to the $`(\mathrm{𝟑},\mathrm{𝟏})+(\mathrm{𝟏},\mathrm{𝟏})`$ representation. Because they have chirality opposite to that of those in the supersymmetric case, they correspond to ‘non-supersymmetric blow up modes’; they express a blow up respecting hyper-Kähler structure that differs from that of the Kaluza-Klein monopole before the blow up. If these modes are switched on, all the supersymmetries are broken. Due to this property, they can acquire mass, and, perhaps, they are removed from the massless spectrum. After the blow up, the manifold becomes smooth and simply connected. Therefore the wrapping number of a string is not conserved and, at first sight, it seems that the GSO projection, which depends on the wrapping number, is ill-defined. However, this is not a problem, as explained below. Because the worldsheet fermions $`\psi ^\mu `$ (We use NS-R formalism.) take values in the tangent bundle of the target space at $`X^\mu `$, the boundary conditions of $`\psi ^\mu `$ depend on the target space holonomy along closed strings. Using the complex coordinates $`\mathrm{\Psi }^1=\psi ^6+i\psi ^7`$, $`\mathrm{\Psi }^2=\psi ^8+i\psi ^9`$ and their hermitian conjugates $`(\mathrm{\Psi }^i)^{}`$ ($`i=1,2`$), the action of $`SU(2)_L\times SU(2)_R`$ rotation is expressed as $$Ug_LUg_R,U\left(\begin{array}{cc}\mathrm{\Psi }^1& (\mathrm{\Psi }^2)^{}\\ \mathrm{\Psi }^2& (\mathrm{\Psi }^1)^{}\end{array}\right),g_LSU(2)_L,g_RSU(2)_R.$$ (9) The Kaluza-Klein monopole with non-supersymmetric blow up is approximately divided into two parts: the central region, which is almost $`A_1`$ ALE, and the outer region, which is only weakly affected by the blow up modes. These two regions possess different hyper-Kähler structures, and holonomies of cycles in the outer and central regions are elements of $`SU(2)_L`$ and $`SU(2)_R`$, respectively. Let us consider a process in which a wrapped string in the outer region moves adiabatically into the central region and the wrapping becomes loose. In the initial and final string configurations, the holonomy along the string is $`\mathrm{𝟏}_2`$ (the rank two unit matrix). On a string passing across the border between the two regions, where the target space is approximately $`𝐑^4/𝐙_2`$, the holonomy is $`\mathrm{𝟏}_2`$. Connecting these with passes in $`SU(2)_L`$ and $`SU(2)_R`$, we obtain the change of the holonomy during the adiabatic process as follows: $$\begin{array}{ccccccccc}\text{the outer region}& & & & \text{the border}& & & & \text{the central region}\\ \mathrm{𝟏}_2& & g_LSU(2)_L& & \mathrm{𝟏}_2& & g_RSU(2)_R& & \mathrm{𝟏}_2\end{array}$$ (10) More explicitly, for example, we can choose the path of the string so that $`g_L=\mathrm{exp}(i\theta _1\sigma _z)`$ and $`g_R=\mathrm{exp}(i\theta _2\sigma _z)`$, where $`\theta _1`$ ($`\theta _2`$) is a parameter in the outer (central) region changing from $`0`$ to $`\pi `$ (from $`\pi `$ to $`0`$). In this case, the phases $`\omega _i`$ in the boundary conditions of the fermions $`\mathrm{\Psi }^i(\sigma +2\pi )=\mathrm{exp}(i\omega _i)\mathrm{\Psi }^i(\sigma )`$ are change as $$\begin{array}{ccccccccccc}& & \text{the outer region}& & & & \text{the border}& & & & \text{the central region}\\ \omega _1& :& 0& & \theta _1& & \pi & & \theta _2& & 0\\ \omega _2& :& 0& & \theta _1& & \pi & & \theta _22\pi & & 2\pi .\end{array}$$ (11) (Here, we are considering the NS-NS sector. For the R-R sector, we should use the boundary conditions $`\mathrm{\Psi }^i(\sigma +2\pi )=+\mathrm{exp}(i\omega _i)\mathrm{\Psi }^i(\sigma )`$.) Therefore, this process causes a shift of the oscillators of $`\mathrm{\Psi }^2`$ ($`\mathrm{\Psi }_{n+1/2}^2\mathrm{\Psi }_{n1/2}^2`$ for the left-moving fermions and $`\stackrel{~}{\mathrm{\Psi }}_{n1/2}^2\stackrel{~}{\mathrm{\Psi }}_{n+1/2}^2`$ for the right-moving fermions) while the oscillators of $`\mathrm{\Psi }^1`$ do not change ($`\mathrm{\Psi }_{n+1/2}^1\mathrm{\Psi }_{n+1/2}^1`$ and $`\stackrel{~}{\mathrm{\Psi }}_{n+1/2}^1\stackrel{~}{\mathrm{\Psi }}_{n+1/2}^1`$). This shift implies that even if we started from a string in the ground state $`|0|0`$, we have an excited state $`\mathrm{\Psi }_{1/2}^2|0(\stackrel{~}{\mathrm{\Psi }}_{1/2}^2)^{}|0`$ after the adiabatic process. These two states have different eigenvalues of operators $`(1)^{F_R}`$ and $`(1)^{F_L}`$, where $`F_L`$ is operator that counts left-moving worldsheet fermions. If the initial state is (not) projected away by the GSO projection, the final state also should (not) be projected away. Therefore, we should impose different GSO projection on the initial and final states. Finally, I would like to note that the scenario described above does not work for type 0 Kaluza-Klein monopoles, which are dual to type II NS5-branes, because the $`()^{F_R}`$ twist cannot be explained by holonomies of the target space. This is consistent with the fact that no bosonic fields appear in the twisted sector of type 0 strings. ## 5 Conclusions We have derived the following duality relations between type II and type 0 brane configurations. * The dual of an unwrapped type II D$`p`$-brane is a stack consisting of a wrapped type 0 electric D($`p+1`$) brane and a wrapped type 0 magnetic D($`p+1`$) brane. * The dual of a wrapped type II D($`p+1`$)-brane is an unwrapped type 0 D-brane. The charge of the type 0 D-brane depends on the Wilson line on the type II D-brane. * The dual of a type 0 (type II) wrapped NS5-brane is a type II (type 0) wrapped NS5-brane. * The dual of a type 0 (type II) unwrapped NS5-brane is a Kaluza-Klein monopole with NUT charge $`2`$ ($`A_1`$-singularity) in type II (type 0) theory. At the singularity in type II theory, there exist non-supersymmetric blow up modes, which probably acquire mass and are removed from the massless spectrum. ## Acknowledgements This work is supported in part by a Grant-in-Aid for Scientific Research from the Ministry of Education, Science, Sports and Culture (#9110).
no-problem/9906/quant-ph9906061.html
ar5iv
text
# Causation and Physics Abstract Philosophical analyses of causation take many forms but one major difficulty they all aim to address is that of the spatio-temporal continuity between causes and their effects. Bertrand Russell in 1913 brought the problem to its most transparent form and made it a case against the notion of causation in physics. In this essay, I focus on this subject of causal continuity and its related issues in classical and quantum physics. I. Introduction II. Causal Connotations and David Hume’s Theory of Causation III. Russell’s Objection to Hume’s Temporal Contiguity Thesis IV. Causal Continuity and Recent Physicalist Accounts of Causation V. Concluding Remarks Acknowledgments This essay is inspired by the kind invitation of the organizers, especially Professor Francesco de Martini, to speak at the III Adriatico Research Conference on Quantum Interferometry. I am indebted to my thesis advisor, Professor Nancy Cartwright for introducing me to this research and her constant inspiration. I am also thankful for the useful comments and encouragement received from Professors Miles Blencowe, Carl Hoefer, Drs. Joseph Berkovitz, Etienne Hofstetter and Makoto Itoh. Professor Yanhua Shih has suggested a number of important modifications for which I am very grateful, and for an important point brought to my attention by Professor Asher Peres during the discussion session that has opened new scope for my research, I extend to him my warm gratitude. I. Introduction Causation is an active area of philosophical research and it is one notion that is deeply entwined with the foundational aspects of both classical and quantum physics. One need not look far but only at the contributions to this volume that causation and its related concepts abound in physics today. When asked, “What is Causation? What do we mean when we think that one object is the cause of another or one event causes another to happen?”, no doubt different connotations would immediately come to mind. And indeed we ought to ask “What are the connotations of causation?” These basic connotations, being largely empirical in character in the sense that they come from our experience and interactions with the physical world, generally receive different treatments by physicists and philosophers. The distinction is a matter of the difference in practice. Physicists accept these basic intuitions as facts about causation and physical theories are constructed to conform to these “conditions of causality”, which are not to be violated. A good example would be that of the “past-future” directed Minkowski light-cone structure defined in terms of “cause-effect” relations<sup>1</sup><sup>1</sup>1See for example, Taylor and Wheeler (1966), Spacetime Physics, p.39.. Philosophers, on the other hand, approach the subject from a different angle; they conduct conceptual analyses of these causal connotations and see whether they do make good logical sense or are infected with grave inconsistencies. With the advent of relativity theories, physics has added important items to the stock of causal facts. Perhaps the most significant one is that special relativity places a limit on the velocity of propagation of causal influences. For the serious philosophical minds, these results cannot afford to go unnoticed and it would indeed be of considerable interest to investigate the extent of the possible interplay between the findings from the respective disciplines. With this motivation in mind, the plan of this essay is as follows. In part II, I shall consider a number of basic causal connotations and the various aspects that may be deduced from these considerations. We are then be placed in a position which leads naturally to a presentation of the main ideas of David Hume’s theory of causation. Hume’s theory is the start of the empiricist philosophical analysis that aims to capture our causal intuitions. The Humean account and its more modern variants collectively still represent the predominant philosophical view on causation. However, in a cleverly written paper in 1913, Bertrand Russell was able to show that the Humean view is not entirely free from inconsistencies given an important assumption on the nature of time. Russell’s argument will be examined carefully in Part III. Part IV focuses on the issue of causal continuity and the recent physicalist approaches to causation which attempt to resolve some of the more pressing difficulties associated with this issue. This will then be followed by a brief conclusion of the main points discussed and some speculative remarks in Part V. To provide an overview of causation is well beyond the scope of this work but it is however my modest aim to bring into focus, in the following pages, some major philosophical worries on the subject which may be fairly regarded as one of the underlying puzzles encountered in the foundations of both classical and quantum physics. II. Causal Connotations and David Hume’s Theory of Causation When one event (or something<sup>2</sup><sup>2</sup>2What kind of entities does the causal relation relate is an important aspect of philosophical analyses of causation. Some argue that the “relata” should be events, while others insist that they may be facts, processes or states of affairs. However, as both Hume and Russell take events as the proper causal relata, we may therefore consider events in the present work.) is regarded as the cause of another, we have “postulated”<sup>3</sup><sup>3</sup>3Notice I have deliberately used the word “postulated” because it is only correct to remain philosophically neutral and avoid making undue assumptions from the outset. the existence of a special relation or connection between the two events. How special is it? We may want to emphasize the importance of such a connection by the expression that the event we have chosen to call the cause and the one that is called the effect are necessarily connected to each other so that had the cause not happened, the effect would not have happened either. Put slightly differently, this counterfactual mode of representation of the special connection refers to an element of necessity in the sense that given the cause, the effect “must” follow and any other situations just simply cannot and would not happen. How then are we to discover this necessary connection, whatever it may be? One useful place to look is to start from our observations of how causes and effects behave generally. An obvious observation is that “causes precede their effects”; namely, causes occur earlier in time than their effects. One realises of course that not every pair of events happening at the same two respective instants of time are to be thought of as causes and effects. A concrete everyday example illustrates (Fig.1). Suppose we have two people, Angelo and Bianca, standing side-by-side in a room and Angelo who is nearer to the light switch turns it on and at the same moment in time, Bianca starts to sing. Even though both Angelo’s action and Bianca’s singing are both events happening prior to the lamp being lit up, we would deem it appropriate to attribute the cause of the lamp lighting to the switching action provided by Angelo but not to Bianca’s singing. Why? It is in part because a continuous physical connection is envisaged between the light switch and the lamp and there is in general no obvious and direct correlation between the processes of singing and the lamp lighting. So temporal succession between two events alone is not a sufficient condition for causation. To place the matter in a more scientific perspective, consider the Minkowski lightcone in Fig.2. Events $`E_j`$ and $`E_k`$<sup>4</sup><sup>4</sup>4These are treated as simultaneous events - lying on the same hyperplane - but the argument would remain valid even if they do not. both lie in the future of the event $`E_i`$ and hence are causally connectible to $`E_i`$ since signals sent from $`E_j`$ can reach either $`E_j`$ or $`E_k`$. However, neither $`E_j`$ nor $`E_j`$ is necessarily connected to $`E_i`$ for there need not be an existing connection after all. Whether there is in fact an actual connection depends upon the existence of actual physical processes linking $`E_i`$ to $`E_j`$ and/or $`E_k`$. Therefore, temporal succession and the spatio-temporal continuity between the cause and effect provided by physical processes together seem necessary of causation. Granted the physical connection between the light switch and the lamp, is this connection really necessary, in the sense that contrary situations are precluded from occurring? Unfortunately, the answer is in the negative. Although it takes an awfully short time for the electrical signal to travel from the light switch to the lamp, it is however always conceivable that an accident like a power cut may occur within this short time interval, as a result of which the lamp would fail to light up. It is therefore not necessarily the case that whenever the switch is turned on, the lamp must light up; this is only true if no other factors are to interfere. Let us imagine instead the scenario where Bianca is the only person in the room and there is no light switch attached to the wall. We observe that when Bianca starts to sing the light comes on a split second later. On one mere instance of this observation, it would be reasonable to put it down as a case of sheer coincidence because we do not usually conceive of a possible (physical) connection between these two events. However, if such an observation is repeated many times and in each and every time, the same sequence obtains so that whenever Bianca starts to sing, the light comes on, then we would conclude that the occurrences of both events in close temporal succession are too regular to be discounted as pure chancy coincidences. And so from repeated observations of the regular succession of the two events, we find it proper and indeed justified to “infer” a special connection (Fig.3) between this pair of events and set out to search for the “hidden” mechanism that could have been responsible for giving rise to such a correlation. The question remains: is such a connection we have so inferred upon repeated observations of regularities a necessary one? Although experience teaches us that frequent correlations are usually prima facie good indicators of causation, it is however well-known that correlations do not have to imply causation. Logic does not prohibit the apparent correlations we see as arising from pure chance. One may well imagine the world to be a chancy enterprise in such a way that “it so happens” that whenever $`E_1`$ occurs, $`E_2`$ follows later. Still, one may argue that quite unbeknown to us, there could exist some kind of a voice-recognition device which provides the physical connection between Bianca’s singing and the lighting up of the lamp. But what gives us the impression and prompts us to look for this “unknown device” in the first place? Nothing other than the constant conjunction of the two events of the singing and the lamp being lit - given the same cause, the same effect follows. Once again, such a physical connection is by no means necessary as for instance, a power cut may occur and tamper with the normal functioning of the device, thus rendering the succession of the two events unattainable. Temporal succession of two events is not sufficient for causation. A case of causation is deemed to obtain when temporal succession is supplemented by the presence of a spatially and temporally continuous physical connection which provides the link between the two events in question. Similarly, the situation where the same cause has found to be followed by the same effect on a great number of occasions also gives us the feel of a causal correlation and induces us to look for an underlying connection. It is in this way that it is often thought that spatio-temporal continuous physical processes are essential for making sure that “the same cause is to be followed by the same effect”. Unforeseen circumstances may happen during the spatial and temporal course of this continuous physical process that frustrate the connection between the two events. Due to the absence of necessity, in the sense that unforeseen circumstances can always occur, it follows that there can never be any guarantee without fail that the same cause is to be followed by the same effect. This clearly indicates an incompatibility between these two causal connotations and that in reality, “the same cause is always followed by the same effect” is not warranted by spatio-temporal continuity. This is quite contrary to our ordinary way of thinking. Spatially and temporally continuous physical processes, though they are by no means necessary connections (in a logical sense), do nevertheless impress upon us the idea of the existence of a causal relation. Such physical connections between two events may either be (i) direct or, (ii) indirect. In the latter case, we seek a third event $`C^{}`$ that existed in the overlap region of the respective past lightcones of the two events such that both events are direct consequences of $`C^{}`$. In this model, both events are effects of the common cause $`C^{}`$ (Fig.4). This illustrates nicely the appeal of hidden variable programs in the efforts to provide an explanation of the Einstein-Podolsky-Rosen (EPR) paradox. Special relativity rules out a direct physical connection between the two space-like separated measurement events. However, the existence of the remarkable correlations of the results of measurements calls urgently for an explanation. It is thus natural to look for possible common causes (hidden variables) in the past which may have given rise to these correlations. But, while such a procedure may satisfy one’s intellectual urge, logic permits a world in which these correlations are all there are; the empire of chance rules in such a manner that the correlations always obtain even without any underlying spatio-temporal continuous connection, be it direct or indirect. In fact, the significant achievement of Bell’s 1964 inequalities rests on their success in dispelling local hidden variable models in favour of quantum mechanics, as has since then been so forcibly confirmed by many sophisticated experiments. In relation to the spatio-temporal continuity between causes and effects, another common connotation we have for causation is to suppose that a certain cause event has occurred but nevertheless the expected effect has somehow failed to materialize. Then we usually explain this by saying that other events must have got in-between them which inhibited the occurrence of the effect. In other words, the spatio-temporal continuity between the supposed cause and the supposed effect is disturbed. In order to diminish the opportunity for other events to get “in-between” and be mischievous, we want to make both the spatial and temporal intervals between the cause and the effect as short as possible. The shorter these intervals are, generally the smaller the probability for other factors to interfere. So, causes and effects are expected to be spatially and temporally close to each other, that is, they should exhibit a degree of spatio-temporal “nearness”. Remote factors are not immediate causes and it is in this way that causal continuity is to be so intimately bound up with the notion of locality. For events happening at vastly separated spacetime locations to be causally connected, we look for events to provide the intermediate links between these two spacetime locales; hence the idea of a “causal chain” to ensure causal continuity across spacetime regions. The foregoing discussions now lead appropriately to the introduction of the major ideas of David Hume’s theory of causation, which exerted tremendous influence over the logical positivists and their contemporaries such as Bertrand Russell. Hume maintains that there are two basic elements to human understanding that form the pillars to his philosophical system: impressions and ideas. Impressions correspond to all “lively signals” we receive from the physical world through our senses, like perceptions, sensations, feelings etc. Ideas, on the other hand, consist in the formation of a conception of impressions. The general principle he adopts for his philosophical analyses is that “all ideas originate from the association and combination of the different impressions”. That is to say, a certain idea we may have for something has to come from our experience through our senses. Armed with this principle, he then asks from which impressions do we form the idea of cause-and-effect as some sort of a necessary connection? He is able to identify three such impressions from our empirical experience behaving like causes and effects. These are: “priority in time” of the cause , “constant conjunction” between the cause and effect and “contiguity” in space and time between the causes and effects. And these should be of some familiarity since they refer to none other than the three causal connotations we have considered in the above: causes precede their effects, given the same cause, the same effect follows and continuity respectively. But as we have already discussed, from these three properties and these three alone, one can never deduce the element of necessity. Hume argued that even though there may actually exist connections in the world which are necessary in some sense, beyond this the only real idea we can have of this connection is of the three properties above. Since these properties are not sufficient to entail necessity, philosophical prudence must now compel us to take a skeptical view of the idea of necessary connection between causes and effects. Granted that our experience is incapable of furnishing us with the idea of a necessary causal connection, how are we able to associate the three impressions of causes and effects to arrive at the idea of a necessary connection between two events? Hume’s answer consists in the fact that after many instances of observing the behaviours of constant conjunction, priority in time and contiguity in time and space between the two events without exception, the mind has in the course grown accustomed to expect that a special connection does indeed exist between the two. This feeling of expectation then gives us the impression from which our idea of connection is copied. The idea is thus not from our experiences of the external world but comes rather from our own response to it. In a sense then the causal relation as a necessary connection is an idea “imposed” by the mind upon unfailing, successive observations of these regular behaviours of causes and effects. The three impressions of priority in time, constant conjunction and contiguity in time and space can never provide us with the idea of a necessary connection. It must again be emphasized that it is never Hume’s intention to deny the existence of necessary connections in nature. Rather, that if the three impressions are all we have by way of the evidence for causal necessity, and since this evidence alone is not adequate for us to serve to reveal to us such an element of necessity, it would be more reasonable not to impose their existence on nature, leaving this instead as an open question. And Hume concludes<sup>5</sup><sup>5</sup>5David Hume, (1888), A Treatise of Human Nature, second edition (1978), with revised text and notes by P.H. Nidditch, Clarendon Press, Oxford. (THN), “As to what may be said, that the operations of nature are independent of our thought and reasoning, I allow it; and accordingly have observed, that objects bear to each other the relations of contiguity and succession; that like objects may be observed in several instances to have like relations; and that all this is independent of, and antecedent to the operations of the understanding. But if we go any farther, and ascribe a power or necessary connection to these objects; this is what we can never observe in them, but must draw the idea from what we feel internally in contemplating them.” (THN, p.168-9) III. Russell’s Objection to Hume’s Temporal Contiguity Thesis In 1912, Russell made his presidential address<sup>6</sup><sup>6</sup>6Delivered on 4 November 1912. The ensuing essay was published in the Proceedings of the Aristotelian Society, 13 (1912-13) and reprinted in Russell, B. (1917), Mysticism and Logic, p.180-208, George Allen and Urwin. to the Aristotelian Society the occasion to cast doubt on the tenability of the Humean account of causation and to argue against the notion of cause in physics. We have already taken pains to stress the inherent incompatibilities among the three causal impressions of priority in time, contiguity in space and time and constant conjunction. In particular, it has been indicated that given the absence of the ingredient of necessity, spatio-temporal continuity is not really capable of ensuring the constant conjunction of the causes and effects. The main reason for this is that even if there is a continuous spatio-temporal physical process connecting the cause and the effect, anything can still happen during the time interval when the causal influence is transmitted down the connection and this results in an uncertainty in the production of the effect. The light switch and the lamp in the last section form a good example. This is why events which are too removed from each other in both the spatial and temporal dimensions are not considered as reliable causes and effects. An immediate solution would be to require that both the spatial and temporal distances between the two events be decreased to such an extent that they stand “adjacent” (or contiguous) to each other so that we may have the assurance that other factors cannot impose themselves and thwart the occurrence of the effect. But what exactly does one mean by two events being “adjacent” to each other when embedded in a background of spacetime continuum? The notion of spatial contiguity between two events is easily satisfied and in the limit it is met by the case where two events can occupy the same location when happening at different times. The notion of temporal contiguity is however more problematic since given that two events occur at the same spatial location with one after another, how are we to ensure that they are temporally contiguous to each other? So this problem reduces to one which concerns temporal contiguity and this is indeed the important issue addressed by Russell in his paper. Russell’s argument begins with a statement of the temporal contiguity thesis (TC). The properties of priority in time and temporal contiguity between cause and effect can be summarised as follows: TC: “Whenever the first event (cause) ceases to exist, the second comes into existence immediately after.” To place TC in the correct perspective, Russell makes the major assumption that time is to be modeled as a mathematical continuum (MC) and is therefore considered as a dense series. A dense series has the distinctive feature that the notion of a “next point” does not make sense because between any two points there always exists others, no matter how close these two points are to each other. It is instructive to contrast the idea of a dense series such as the real number line with the discrete series of positive integers where the notion of consecutive (or “next”) members does take on a well-posed meaning. Having specified how the temporal continuum is to be represented, we now consider two point events $`c`$ and $`e`$ occurring at two respective instants of time $`t_1`$ and $`t_2`$ ($`t_1<t_2`$). Because time is a dense series, it follows that between any two instants (points) of time, there are always other instants (points) no matter how short we make the interval $`t_2t_1`$. That is, there is always a temporal gap between $`c`$ and $`e`$ and hence $`c`$ cannot be contiguous in time to $`e`$. Furthermore, this temporal gap provides ample opportunities for other events to creep in between $`c`$ and $`e`$ and to interfere. While these other factors may not prove harmful to the production of $`e`$ at $`t_2`$, they may however also behave otherwise and hinder the occurrence of $`e`$ (Fig.5). Under these circumstances, one cannot be certain that the same cause is always followed by the same effect since there can always be the chance of $`e`$ not occurring whenever there is to be this temporal gap between the two events. In order to be rid of unsolicited factors, one must devise a means to ensure that the temporal gap is filled. An obvious way is to suppose the cause event as having a temporal dimension (Fig.6). The cause is now a static, unchanging event<sup>7</sup><sup>7</sup>7If a non-static, changing event such as one composed of a causal chain of discrete events as in Fig.5, then the problem of temporal gaps existing in-between these events within the causal chain remains., occupying the half-open interval, and is imagined to sit there from time $`t_1`$ to $`t_2`$, filling the temporal gap and all of a sudden, turns into $`e`$ at $`t_2`$. However, Russell objects strongly to such kinds of events: he argues that it is not at all logical why, being unchanging and sitting there complacently, $`c`$ has to turn into $`e`$ at $`t_2`$ and not at any other moments, say earlier at $`t_0`$ or later at $`t_3`$? And so static, unchanging events are dismissed outright by Russell as an impossibility. Since these static, unchanging events which seem to be the only means by which the temporal gap can be occupied are not plausible, we must therefore draw the conclusion that there always exists a temporal gap between $`c`$ and $`e`$ so that $`c`$ cannot be contiguous to $`e`$. Our intuition about the temporal continuity of causes and effects comes under threat given the assumption of physical time as a mathematical continuum and constant conjunction cannot be guaranteed under such a circumstance. Russell has succeeded in showing that there exist tremendous tensions between our usual connotations of the causal relation. VI. Causal Continuity and Recent Physicalist Accounts of Causation Despite the difficulty brought to light by Russell’s critique of the Humean temporal contiguity thesis, one is, of course, allowed to argue that the major issue is really the definition of events as points occurring at discrete temporal instants within the temporal continuum. There is simply no place for the notion of discreteness with a temporal continuum. So, a more amicable approach would be to “superpose” a continuum of events - a continuous rope of events - upon this temporal continuum in the sense that we consider all the events that happen locally within this time interval. Here we focus more closely on the aspect of causal continuity by first considering this example from Elizabeth Anscombe<sup>8</sup><sup>8</sup>8Anscombe, E (1974), ”Times, Beginnings and Causes” in the Proceedings of the British Academy. Reprinted in G.E.M. Anscombe, Metaphysics and the Philosophy of Mind, Collected Philosophical Papers Volume II, p.148-162. (1974, p.150): “Find an object here and ask how it comes to be there?” (Fig.7) A causal explanation, says Anscombe, would be “it went along some path from A to B”. The locution “along some path” in fact entails more than the case where the object just turned up at location B after having been at A previously. It requires the object to occupy also all the intermediate positions between A and B. To satisfy constant conjunction, it is sufficient for the object to turn up at location B after having been at A some moments earlier and without having to assume the intermediate positions between the two locations. But this would not be deemed to be an adequate causal explanation. And to “explain causally”, a path has to be imposed to provide the connection between the two events of the object being at the two respective spacetime locations. For the purpose of explanation, it is therefore proper to consider spatio-temporal continuous connections when thinking about causation. In physics, the notion of continuity is usually either represented by spatially and temporally continuous paths or by trajectories in phase space. These spacetime paths and phase space trajectories are in turn the solutions of differential and integral equations. Recalling the fact that while these differential and integral equations guarantee continuity, they however lack the crucial causal aspect of an explicit temporal order for cause and effect. The pressing question which must now occupy us is how we may introduce a temporal order into a theory of causation which takes seriously the view that spatio- temporal continuous physical processes, as represented by the equations of physics, provide for us the appropriate causal connections. We find a clue in the following consideration of the Minkowski lightcone (Fig.8), Does it make sense to assign a temporal direction from the past to future as directing from the lower-half of the lightcone towards the upper-half and not vice versa? In other words, does it make any sense to provide the worldline as going through the worldpoints A , B with an arrow and to assert that B is before P and A after P? Einstein asks in 1949<sup>9</sup><sup>9</sup>9Einstein A (1949), “Reply to Criticisms” in Albert Einstein: Philosopher-Scientist, p.687-688, The Library of Living Philosophers Volume VII, Edited by P.A. Schilpp, Open Court., “Is what remains of temporal connection between world-points in the theory of relativity an asymmetrical relation, or would one be just as much justified, from the physical point of view, to indicate the arrow in the opposite direction and to assert that A is before P, B after P?” Physics has a ready reply to this. The “temporal arrow” is secured by the observations of irreversible processes in nature despite the much advertised time- symmetrical character of the laws of physics. These processes are believed to be ultimately related to the growth of entropy in the universe. The “past-future” direction of the lightcone may equally well be defined in terms of either the direction of the cause-effect relation or that of irreversible processes. This suggests that the causal direction may indeed be identified with the direction of irreversible processes. Or put differently, a cause can be considered as an event which introduces a “change” which is irreversible. The idea of a cause as an irreversible change forms the backbone of the physicalist theories of causation. Although there are variations amongst the theories of physical causation that have been put forward, they nevertheless share one basic underlying idea: causal continuity is guaranteed by the transmission of causal influences (objective physical quantities) along continuous space-time paths governed by physical laws. The “objective physical quantities” being transferred refer usually to either momentum or energy in the class of approaches subsumed under the title of “transference theories of causation”. A slightly more sophisticated version is the so-called “process-theory” of Wesley Salmon<sup>10</sup><sup>10</sup>10Salmon, W C (1984), Scientific Explanation and the Causal Structure of the World, Princeton University Press.; there it refers to the transmission of marks with the marks being changes that have resulted from irreversible physical interactions. Interactions are responsible for bringing about or producing the irreversible changes. To cast this concept in a better context, consider the simple case of one mass in motion colliding with another which is initially at rest, subsequently setting the second into motion (Fig.9). The same state of affairs can be described by two different causal stories. In the rest frame of m the moving mass M travelling with velocity v appears to be the earlier event - the cause which is responsible for the change of states of both masses. On the other hand, in the rest frame of M, the earlier event of m moving with velocity -v is now regarded as the cause giving rise to the subsequent change of motion of both masses. Hence, we find ourselves confronted by two different causal stories whose accuracies depend on the frame of reference in which the same state of affairs is viewed. The objective matter-of-fact is however that for all inertial frames of reference, the “collision” between M and m produces the subsequent “changes” of motion of each of the masses. It is important to realise that the collision occurs in all frames of reference and after which is to be followed by changes in motions of these masses. It is indeed by this very means that an objective temporal order may indeed be established. Because of this “causal interaction” between the masses M and m, their respective energies and momenta are correlatively modified accordingly. Both masses, having interacted, will carry the causally modified dynamical properties via their continuous spatio-temporal trajectories and may participate in further interactions (Fig.10). This picture of causation, central to the physicalist approaches, must be modified when it is carried over to the quantum regime. There, the idea of a physical system following strict continuous trajectories has long evaporated and in its place stands instead a series of discrete points corresponding to specific measurement events performed on the system. The inherent probabilistic nature of the quantum world involving interference does not sanction any definite interpolation between these points. Measurement is after all a kind of interaction between the system and the measuring apparatus that brings about an irreversible change to both. Even though one may now find the notion of continuity dubious in this domain, the concept of causal interaction survives seemingly unscathed in the face of probabilistic indeterminism. In-between measurements, the system is described by the continuous evolution of the wavefunction as governed by the Schrödinger equation. However, this continuous evolution refers only to a distribution of the different probabilities of obtaining various outcomes of the measurements of a certain dynamical observable. It does not represent an evolution of the successive dynamical values as “possessed” by the system in time as in the classical case. Quantum mechanically, the very act of measurements brings about irreversible changes of the state of a physical system and so interactions can be thought of as the cause as a result of which the probability distributions of the values of dynamical observables are altered. This is similar to the picture suggested independently by Rudolf Haag<sup>11</sup><sup>11</sup>11See, for example, Haag, R (1990), “Fundamental Irreversibility and the Concept of Events”, Commun. Math. Phys., 132, p.245-251. in a series of papers which take the view that in the regime of low density, quantum field theory describes a world where events are the collision processes between particles and the particles themselves provide the causal tie, i.e., the causal connection carrying the modified dynamical structure. V. Concluding Remarks Scientists seek causes for the purpose of providing scientific explanations. It is often regarded that effects follow necessarily from causes. Indeed, it is the major contribution of Hume’s theory of causation to show that this is where we err and the inference from causes to effects is not deductive but rather inductive in nature. This immediately calls into question the idea of a necessary connection which is thought to be the vital element of causal relations. Russell went further and showed that there clearly exist inconsistencies when we consider the causal relation as one between events happening at discrete temporal intervals superposed upon the temporal continuum. Advocates of the class of the so-called physicalist approaches to causation provide a promising framework for a theory of physical causation in the face of Hume’s and Russell’s problems. The problematic “temporal gap” in Russell’s analysis is closed by the consideration of a continuum of events so that the temporal continuum can be matched by the spatio-temporal continuous character of physical processes. Causes, in these accounts, are interactions which bring about irreversible changes from which a causal order can be defined. This picture works well in the domain of classical physics and requires appropriate modifications when applied to the quantum regime. Perhaps I should now leave the reader with the following thought. The most startling case for causation is that of spontaneous decay where an atom sits there for a while and then undergoes decay. These are conceived to be “uncaused” events by most philosophers because the time when the atom is to decay cannot be known with exactitude. Might the nucleus resemble Russell’s “static, unchanging” event<sup>12</sup><sup>12</sup>12It has been kindly pointed out to me by Professor Morton Rubin that the atom is not really “unchanging” in Russell’s sense: previous interactions with the vacuum must be present in order to ”prepare” the nucleus in an unstable state in the first place. However, I would argue that this does not weaken the thrust of Russell’s argument, namely that even given that we have a changing event, how might one explain how the atom comes to decide at which exact moment in time it should undergo decay?, sitting there for a period of time and suddenly undergoing decay? Recalling that the major assumption that Russell makes is that time is to be viewed as a mathematical continuum, it would be a most interesting investigation to see whether adopting instead a picture of time as discrete units will shed different light on this problem. Once again, this illustrates how deeply the notion of causation is connected with the very nature of space and time.
no-problem/9906/astro-ph9906342.html
ar5iv
text
# RELATIVISTIC THERMAL BREMSSTRAHLUNG GAUNT FACTOR FOR THE INTRACLUSTER PLASMA. II. ANALYTIC FITTING FORMULAE ## 1 INTRODUCTION Three of the present authors (S.N., N.I., and Y.K.) have recently carried out accurate calculations on the relativistic thermal bremsstrahlung Gaunt factor for the intracluster plasma (Nozawa, Itoh, & Kohyama 1998). Their calculation is based on the method of Itoh and his collaborators (Itoh, Nakagawa, & Kohyama 1985; Nakagawa, Kohyama, & Itoh 1987; Itoh, Kojo, & Nakagawa 1990; Itoh et al. 1991,1997). In calculating the relativistic thermal bremsstrahlung Gaunt factor for the high-temperature, low-density plasma, Nozawa, Itoh, & Kohyama (1998) have made use of the Bethe-Heitler cross section (Bethe & Heitler 1934) corrected by the Elwert factor (Elwert 1939). They have also calculated the Gaunt factor using the Coulomb-distorted wave functions for nonrelativistic electrons following the method of Karzas & Latter (1961). They have presented the results of the numerical calculations in the form of extensive numerical tables. The grid of their numerical table is made fine enough so that smooth interpolations can be carried out for the applications of the results to the analyses of various astrophysical observational data. Other references on the calculation of thermal bremsstrahlung Gaunt factor include Culhane (1969), Culhane & Acton (1970), Raymond & Smith (1977), Gronenschild & Mewe (1978), Mewe, Lemen, & van den Oord (1986), and Carson (1988). In this paper we will present accurate analytic fitting formulae which summarize the results of the extensive numerical tables presented by Nozawa, Itoh, & Kohyama (1998). Analytic fitting formulae can be readily implemented in the computer programs for the analyses of the observational data. Therefore, we recommend the present analytic fitting formulae be used widely by the observers for the analyses of their observational data. The present paper is organized as follows. We will give the analytic fitting formulae for the relativistic Gaunt factor in $`\mathrm{\S }`$2. We will give the analytic formula for the nonrelativistic exact Gaunt factor in $`\mathrm{\S }`$3. Concluding remarks will be given in $`\mathrm{\S }`$4. ## 2 ANALYTIC FITTING FORMULAE FOR THE RELATIVISTIC GAUNT FACTOR The thermal bremsstrahlung emissivity is expressed in terms of the thermally averaged relativistic Gaunt factor $`g_{Z_j}`$ (Nozawa, Itoh, & Kohyama 1998) by $`<W(\omega )>d\omega `$ $`=`$ $`1.426\times 10^{27}g_{Z_j}(T,u)n_en_jZ_j^2T^{1/2}`$ (1) $`\times `$ $`e^udu\mathrm{ergs}\mathrm{s}^1\mathrm{cm}^3,`$ $`u`$ $``$ $`{\displaystyle \frac{\mathrm{}\omega }{k_BT}},`$ (2) where $`\omega `$ is the angular frequency of the emitted photon, $`T`$ is the temperature of the electrons (in kelvins), $`n_e`$ is the number density of the electrons (in cm<sup>-3</sup>), and $`n_j`$ is the number density of the ions with the charge $`Z_j`$ (in cm<sup>-3</sup>). In Nozawa, Itoh, & Kohyama (1998), the nonrelativistic Gaunt factor $`g_{NR}`$ has been also calculated. This Gaunt factor is exact in the low-temperature limit. At intermediate temperatures, it has been confirmed by Nozawa, Itoh, & Kohyama (1998) that the relativistic Gaunt factor which has been calculated with the use of the Bethe-Heitler cross section corrected by the Elwert factor shows excellent agreement with the nonrelativistic exact Gaunt factor. At higher temperatures, the nonrelativistic Gaunt factor deviates from the relativistic Gaunt factor because of the insufficiency of the nonrelativistic approximation. These two Gaunt factors have been tabulated in Nozawa, Itoh, & Kohyama (1998). We have thoroughly examined the numerical results presented in the tables in Nozawa, Itoh, & Kohyama (1998). We have confirmed that the data are almost perfect. Nevertheless, we have found 11 cases where the numerical computation has not converged sufficiently. All are for the nonrelativistic exact Gaunt factor at low temperatures. Therefore, they will not be relevant to the intracluster plasma. However, for the sake of completeness, we will replace some of the tables in Nozawa, Itoh, & Kohyama (1998) with the revised ones. In particular, we replace the tables on pages 541, 551, 552 of Nozawa, Itoh, & Kohyama (1998) with TABLES 1-3. In TABLES 1-3, $`\gamma ^2`$ is defined by $`\gamma ^2`$ $`=`$ $`{\displaystyle \frac{Z_j^2\mathrm{R}_\mathrm{y}}{k_BT}}=Z_j^2{\displaystyle \frac{1.579\times 10^5\mathrm{K}}{T}}.`$ (3) At sufficiently high temperatures, we adopt the relativistic Gaunt factor. At sufficiently low temperatures, we adopt the nonrelativistic exact Gaunt factor. At intermediate temperatures, these two Gaunt factors coincide with each other for small values of $`Z_j`$. For larger values of $`Z_j`$, the two Gaunt factors show small discrepancies even at intermediate temperatures. Therefore, we generally interpolate between the two Gaunt factors smoothly at intermediate temperatures. To be more precise, we find the point at which the discrepancy between the two Gaunt factors (for fixed values of $`Z_j`$ and u) is the smallest as a function of the temperature. Then we interpolate between the two Gaunt factors smoothly using a sine function. The temperature range for the interpolation is $`\mathrm{\Delta }`$ $`\mathrm{log}T`$ = $`\pm `$ 0.1 to $`\pm `$ 0.5 with respect to the central temperature at which the discrepancy is the smallest depending on the minimum value of the discrepancy. We give analytic fitting formulae for $`Z_j=128`$ . The range of the fitting is $`6.0\mathrm{log}T8.5`$, $`4.0\mathrm{log}u1.0`$. We express the Gaunt factor by $`g_{Z_j}`$ $`=`$ $`{\displaystyle \underset{i,j=0}{\overset{10}{}}}a_{ij}t^iU^j,`$ (4) $`t`$ $``$ $`{\displaystyle \frac{1}{1.25}}[\mathrm{log}T7.25],`$ (5) $`U`$ $``$ $`{\displaystyle \frac{1}{2.5}}[\mathrm{log}u+1.5].`$ (6) The coefficients $`a_{ij}`$ for $`Z_j=128`$ are presented in TABLE 4. The accuracy of the fitting is generally better than 0.1%. ## 3 ANALYTIC FITTING FORMULA FOR THE NONRELATIVISTIC EXACT GAUNT FACTOR The thermal bremsstrahlung emissivity in the nonrelativistic limit is expressed in terms of the nonrelativistic exact Gaunt factor $`g_{\mathrm{NR}}`$ (Nozawa, Itoh, & Kohyama 1998) by $`<W(\omega )>_{\mathrm{NR}}d\omega `$ $`=`$ $`1.426\times 10^{27}g_{\mathrm{NR}}(\gamma ^2,u)n_en_jZ_j^2T^{1/2}`$ (7) $`\times `$ $`e^udu\mathrm{ergs}\mathrm{s}^1\mathrm{cm}^3,`$ $`u`$ $``$ $`{\displaystyle \frac{\mathrm{}\omega }{k_BT}},`$ (8) $`\gamma ^2`$ $``$ $`{\displaystyle \frac{Z_{j}^{}{}_{}{}^{2}\mathrm{Ry}}{k_BT}}=Z_{j}^{}{}_{}{}^{2}{\displaystyle \frac{1.579\times 10^5\mathrm{K}}{T}},`$ (9) where $`\omega `$ is the angular frequency of the emitted photon, $`T`$ is the temperature of the electrons (in kelvins), $`n_e`$ is the number density of the electrons (in cm<sup>-3</sup>), and $`n_j`$ is the number density of the ions with the charge $`Z_j`$ (in cm<sup>-3</sup>). It should be noted that the thermal bremsstrahlung emissivity in the nonrelativistic limit is a function of $`\gamma ^2`$ and $`u`$ only. It does not depend on $`Z_j`$ and $`T`$ separately, but on the ratio $`Z_{j}^{}{}_{}{}^{2}/T`$. This is a remarkable fact for nonrelativistic electrons. In Figure 1 we show the nonrelativistic Gaunt factor as a function of $`u`$ for various values of $`\gamma ^2`$. In Figure 2 we show the nonrelativistic Gaunt factor as a function of $`\gamma ^2`$ for various values of $`u`$. We give an analytic fitting formula for the nonrelativistic exact Gaunt factor. The range of the fitting is $`3.0\mathrm{log}\gamma ^22.0`$, $`4.0\mathrm{log}u1.0`$. We express the Gaunt factor by $`g_{\mathrm{NR}}`$ $`=`$ $`{\displaystyle \underset{i,j=0}{\overset{10}{}}}b_{ij}\mathrm{\Gamma }^iU^j,`$ (10) $`\mathrm{\Gamma }`$ $``$ $`{\displaystyle \frac{1}{2.5}}[\mathrm{log}\gamma ^2+0.5],`$ (11) $`U`$ $``$ $`{\displaystyle \frac{1}{2.5}}[\mathrm{log}u+1.5].`$ (12) The coefficients $`b_{ij}`$ are presented in TABLE 5. The accuracy of the fitting is generally better than 0.1%. ## 4 CONCLUDING REMARKS We have presented accurate analytic fitting formulae for the relativistic Gaunt factor as well as for the nonrelativistic exact Gaunt factor for thermal bremsstrahlung. The analytic fitting formulae have been constructed to reproduce the numerical results of the calculation reported in Nozawa, Itoh, & Kohyama (1998). The accuracy of the fitting is generally better than 0.1%. The present fitting formulae can be used widely for the analysis of thermal bremsstrahlung radiation. We thank Professor Y. Oyanagi for allowing us to use the least square fitting program SALS. We also wish to thank our referee for useful comments. We are indebted to the editors Dr. Helmut A. Abt and Dr. F. W. Stecker for their valuable advices on the form of the paper. This work is financially supported in part by the Grant-in-Aid of Japanese Ministry of Education, Science, Sports, and Culture under the contract #10640289. Figure Legends | FIG.1. | Nonrelativistic exact Gaunt factor as a function of $`u`$ for various values of $`\gamma ^2`$. | | --- | --- | | FIG.2. | Nonrelativistic exact Gaunt factor as a function of $`\gamma ^2`$ for various values of $`u`$. |
no-problem/9906/astro-ph9906246.html
ar5iv
text
# Combining the baryon budget with CMBR measurements ## 1 introduction Many recent measurements seem to indicate that our universe is not a critical density, matter dominated Friedmann universe. Rather, the recent measurements of type Ia supernovae \[Perlmutter et al. 1998, Perlmutter et al. 1997, Garnavich et al. 1998, Riess et al. 1998, Schmidt et al. 1998\] strongly suggest that, although the universe has a flat geometry, the matter density is low \[Perlmutter et al. 1998, Riess et al. 1998\]. The energy density is instead dominated by either a cosmological constant \[Carroll, Press and Turner 1992\], or by a similar type of energy with negative pressure, such as quintessence \[Wang et al. 1999, Zlatev, Wang and Steinhardt 1999, Caldwell, Dave and Steinhardt 1998\]. Numerous investigations have shown that the fluctuation spectrum of the cosmic microwave background radiation (CMBR) provides a potentially very powerful tool for determining cosmological parameters \[Jungman et al. 1996, Bond, Efstathiou and Tegmark 1997, Zaldarriaga, Spergel and Seljak 1997, Hu, Eisenstein and Tegmark 1999\]. Thus, one could hope that accurate measurements of these fluctuations could resolve the issue of whether or not the universe is dominated by vacuum energy. Based on present observations \[Lineweaver and Barbosa 1998, Oliveira-Costa et al. 1999, Coble et al. 1999\] there already exist a number of estimates of the cosmological parameters \[Efstathiou et al. 1999, Webster et al. 1998, Hancock et al. 1998, Lineweaver and Barbosa 1998, Bond and Jaffe 1998, Bernardis et al.1997, Tegmark 1998, Lineweaver 1998, White 1998\]. There is, however, a severe problem in that a change in one parameter can often be mimicked by suitable changes in a combination of other parameters \[Hu, Eisenstein and Tegmark 1999, Eisenstein, Hu and Tegmark 1998a, Eisenstein, Hu and Tegmark 1998b\]. This is for instance very problematic when trying to simultaneously determine the cosmological matter density, $`\mathrm{\Omega }_m`$, and the Hubble parameter, $`H_0`$. It has therefore been suggested that CMBR measurements should be combined with other constraints, coming for instance from large scale structure surveys \[Eisenstein, Hu and Tegmark 1998a, Eisenstein, Hu and Tegmark 1998b, Webster et al. 1998\] or supernova type Ia measurements \[Tegmark 1998, White 1998, Lineweaver 1998, Eisenstein, Hu and Tegmark 1998a, Eisenstein, Hu and Tegmark 1998b, Efstathiou et al. 1999\]. In this paper we explore another possibility, namely how the cosmic baryon abundance can be used together with CMBR to provide tight constraints on $`\mathrm{\Omega }_m`$ and $`H_0`$. Galaxy clusters contain large amounts of hot, X-ray emitting gas from which the baryon mass to total mass ratio can be measured \[White and Fabian 1995, David, Jones and Forman 1995, Evrard, Metzler and Navarro 1996, Evrard 1997\]. It turns out that measurements of this baryon cluster fraction can break the parameter degeneracy inherent in the CMBR measurements, and although most of the present CMBR data is of relatively low accuracy \[Lineweaver and Barbosa 1998, Oliveira-Costa et al. 1999, Coble et al. 1999\], it is still sufficiently accurate to provide good constraints on the relevant cosmological models. We have used the presently available data, together with data from Big Bang nucleosynthesis and measurements of the cluster baryon fraction, and find that the observations are indeed strongly incompatible with a critical density matter dominated model. Our results are easily compared with for instance the new supernova type Ia measurements \[Perlmutter et al. 1998\], and are found to be completely compatible. There still remains a quite large uncertainty on the determination of $`\mathrm{\Omega }_m`$ and $`H_0`$ because of the low accuracy of the CMBR data. However, in the very near future, CMBR data of very high quality should become available from several different experiments. There is the balloon borne experiment BOOMERANG \[Hanany 1997\] which has already been flown. Also, there are two new satellite experiments, MAP and PLANCK <sup>1</sup><sup>1</sup>1For information on these missions see the Internet pages for MAP (http://map.gsfc.nasa.gov) and PLANCK (http://astro.estec.esa.nl/Planck/). which will measure the fluctuation spectrum very accurately on sub-degree scales. This should provide data which is accurate enough to diminish the uncertainties by an order of magnitude. ## 2 breaking degeneracy As mentioned above, parameter extraction from the CMBR data suffers from some very large parameter degeneracies \[Eisenstein, Hu and Tegmark 1998a, Eisenstein, Hu and Tegmark 1998b\]. For instance it is not possible to constrain $`\mathrm{\Omega }_m`$ and $`H_0`$ separately, effectively only the combination $`\mathrm{\Omega }_mh^2`$ \[Eisenstein, Hu and Tegmark 1998a, Eisenstein, Hu and Tegmark 1998b, Hu, Eisenstein and Tegmark 1999\], where $$h=\frac{H_0}{100\mathrm{k}\mathrm{m}\mathrm{s}^1\mathrm{Mpc}^1},$$ (1) can be measured accurately. However, as shown previously by several authors, it is possible to break this degeneracy by combining CMBR data with other, complementary, measurements. It was shown by Eisenstein et al. \[Eisenstein, Hu and Tegmark 1998a, Eisenstein, Hu and Tegmark 1998b\] that large scale structure surveys like the Sloan Digital Sky Survey (SDSS) <sup>2</sup><sup>2</sup>2See for instance http://www.astro.princeton.edu/BBOOK/ can probe the combination $`\mathrm{\Omega }_mh^{1/3}`$. This means that the error ellipses from such surveys are almost orthogonal to those of CMBR. The joint likelihood function should show a spike-like structure at the crossing of the two ellipses instead of the very elongated structure of either of the two individual measurements \[Eisenstein, Hu and Tegmark 1998a, Eisenstein, Hu and Tegmark 1998b\]. In fact, Webster et al. \[Webster et al. 1998\] have already performed a joint analysis of the present CMBR data together with data from the IRAS 1.2 Jy galaxy survey \[Fisher et al. 1995\]. They find that the CMBR likelihood contours are indeed narrowed significantly in the $`(\mathrm{\Omega }_m,H_0)`$ plane. The above way of breaking degeneracy will work when combining CMBR with any type of measurement that does not depend on $`\mathrm{\Omega }_m`$ and $`H_0`$ as $`\mathrm{\Omega }_mh^2`$. One such possibility is to use the cluster baryon fraction. It has long been known that measurements of the baryon cluster fraction favour a low density universe because the measured fraction is so high that it cannot support $`\mathrm{\Omega }_m=1`$ without violating BBN constraints \[Bludman 1998, Steigman, Hata and Felten 1999, White et al. 1993, Evrard 1997\]. A standard assumption in this game is to assume that the cluster baryon fraction is the same as the universal fraction, an assumption usually referred to as the fair sample hypothesis. Numerical simulation seem to justify this assumption. In fact recent simulations indicate that the cluster baryon fraction is slightly lower that that of the universe as a whole (see \[Steigman, Hata and Felten 1999\] for a discussion). One problem is that observed clusters have diverging baryon fractions, depending on their total mass. It is argued by Evrard et al. \[Evrard, Metzler and Navarro 1996\] that this is most likely due to errors in the estimate of the total cluster mass, and that it can be corrected by use of statistical methods. In this paper we shall use the results obtained by Evrard \[Evrard 1997\] for the universal cluster baryon fraction. The method for estimating $`\mathrm{\Omega }_m`$ and $`H_0`$ then works as follows: Big Bang nucleosynthesis (BBN) can be used to measure the baryon-to-photon ratio \[Kolb and Turner 1990\], $`\eta `$, which is related to $`\mathrm{\Omega }_b`$ by $$\mathrm{\Omega }_bh^2=3.66\times 10^7\eta .$$ (2) Thus, nucleosynthesis only measures $`\mathrm{\Omega }_bh^2`$, not $`\mathrm{\Omega }_b`$. On the other hand, the cluster baryon fraction, $`f_B`$, is really measured as \[Bludman 1998, Evrard 1997\] $$f_Bh^{3/2}=\frac{\mathrm{\Omega }_b}{\mathrm{\Omega }_m}h^{3/2}.$$ (3) By combining BBN measurements with the cluster baryon fraction we can therefore constrain the combination $$\mathrm{\Omega }_mh^{1/2}.$$ (4) This combination of $`\mathrm{\Omega }_m`$ and $`h`$ is sufficiently different from $`\mathrm{\Omega }_mh^2`$ that when combining them it becomes possible to constrain both parameters well. ## 3 Measurements ### 3.1 CMBR measurements In general, the fluctuations in the CMBR is measured in terms of spherical harmonics $$T(\theta ,\varphi )=\underset{lm}{}a_{lm}Y_{lm}(\theta ,\varphi ),$$ (5) where the coefficients are related to the power-spectrum $`C_l`$ coefficients by $$C_l|a_{lm}|^2_m.$$ (6) At present there is a host of different CMBR experiments, ranging from the largest scales (COBE), down to very small scales. The data that we use are based on the compilation by Lineweaver and Barbosa \[Lineweaver and Barbosa 1998\], but with the addition of the new Python V \[Coble et al. 1999\] results, and the results from the QMAP experiment \[Oliveira-Costa et al. 1999\]. ### 3.2 BBN measurements The past few years have seen a very large fluctuation in the estimated baryon-to-photon ratio (see for instance \[Steigman 1998\] for a review). It has long been known that the primordial value of deuterium provides a very sensitive probe of $`\eta `$ \[Kolb and Turner 1990\], but the problem has been to measure the primordial deuterium abundance. In the local interstellar medium, the abundance is quite well determined \[Hata et al. 1995\], $$D/H=1.6\times 10^5,$$ (7) but this can only really be used to provide a strict lower limit to the primordial abundance since deuterium is only destroyed, not produced, in astrophysical environments. Measurements of deuterium in quasar absorption systems at high redshift have provided a completely new way of measuring the primordial abundance because such systems are chemically unevolved and should therefore contain deuterium abundances close to the primordial \[Steigman 1998\]. There have been two conflicting estimates of the deuterium abundance in these systems, one which is much higher than the local value \[Songaila et al. 1994, Carswell et al. 1994, Rugers and Hogan 1996, Webb et al. 1997, Tytler et al. 1998\], $$D/H2\times 10^4,$$ (8) and one which is only a factor of two higher than the local value \[Burles and Tytler 1998a, Burles and Tytler 1998b\], $$D/H3\times 10^5.$$ (9) There is growing evidence that the low deuterium value is the correct one, and as observational values we shall take the so-called Low-deuterium/High-helium data set, which is given by \[Burles and Tytler 1998a, Burles and Tytler 1998b, Izotov and Thuan 1998\] $`Y_P`$ $`=`$ $`0.245\pm 0.002`$ (10) $`D/H`$ $`=`$ $`(3.4\pm 0.3)\times 10^5.`$ (11) This set of data is completely compatible with standard Big Bang nucleosynthesis for a baryon-to-photon ratio of $$\eta =(5.1\pm 0.3)\times 10^{10},$$ (12) or, in terms of baryon density $`\mathrm{\Omega }_bh^2=0.019\pm 0.001`$. ### 3.3 Baryon cluster fraction As mentioned above, Evrard \[Evrard 1997\] has calculated the universal cluster baryon fraction based on a large number of clusters. His result is $$f_B=(0.060\pm 0.003)h^{3/2}.$$ (13) The $`1\sigma `$ uncertainty is perhaps somewhat underestimated in the measurement, and we shall follow Hata et al. \[Steigman, Hata and Felten 1999\] in assuming taking the $`1\sigma `$ uncertainty to be twice as large, i.e. 0.006. The above value is derived from the gas fraction alone. The fraction in hot gas relative to collapsed baryonic objects has been estimated by White et al. \[White et al. 1993\] to be $$\frac{M_{\mathrm{gas}}}{M_{\mathrm{gal}}}=5.5h^{3/2}$$ (14) which is large enough to be insignificant. ## 4 Likelihood analysis In order to estimate the underlying cosmological parameters, we have calculated a large number of synthetic models which are then compared with the different data sets. Our synthetic models range through a large parameter space and have all been calculated using the publicly available CMBFAST code \[Seljak and Zaldarriaga 1996\]. We restrict the discussion to strictly flat models, i.e. models with $`\mathrm{\Omega }_m+\mathrm{\Omega }_\mathrm{\Lambda }=1`$. As free parameters we choose the normalisation, $`Q`$, the matter density, $`\mathrm{\Omega }_m`$, the baryon density, $`\mathrm{\Omega }_b`$, the Hubble parameter, $`h`$, and the spectral index, $`n`$. All the above parameters are allowed to vary with the restrictions described in table I. Altogether, 65000 independent CMBR spectra have been calculated, and, apart from the fact that we do not investigate open models, this analysis is comparable to that of Lineweaver \[Lineweaver 1998\]. In a new analysis, Tegmark \[Tegmark 1998\] has calculated the likelihood function in a larger, 9-dimensional parameter space. The extra parameters are: the optical depth to reionisation, $`\tau `$, and the amplitude and spectral index of tensor fluctuations. Here, it was shown that relaxing the assumption about zero curvature significantly loosens the constraint on the combination $`\mathrm{\Omega }_mh^2`$. In fact Tegmark finds that there is no upper bound on $`\mathrm{\Omega }_mh^2`$, even at the 68% confidence limit, contrary to our findings below. The reason for this is that the constraint relies to a certain extent on the amplitude differences in the power spectrum at different $`l`$. The information stored in this difference is sensitive to changing the spatial curvature, and therefore relaxing the flatness assumption leads to a much looser upper bound on $`\mathrm{\Omega }_mh^2`$. This being said, there are good reasons for neglecting spatial curvature. Flatness is a generic prediction of almost all inflationary models, and furthermore the new data on Supernovae of type Ia clearly favour a spatially flat universe \[Perlmutter et al. 1998\]. From the same prejudice that has led us to consider only flat models, we have neglected possible tensor fluctuations. Almost all inflationary models predict only very small tensor fluctuations, so this is likely to be a good assumption. We have also neglected reionisation in our calculations. There is no justification for this since we know for a fact that the universe was reionised at fairly high redshift. As shown by Tegmark \[Tegmark 1998\], adding reionisation to the model parameters broadens the likelihood function, but does not move the maxima. Therefore we stress that our likelihood estimates should be considered optimistic and that they will broaden if reionisation is taken into account, but that they should still home in on the correct central values. In order to compare with the different data sets we then perform a $`\chi ^2`$ analysis. In general, $`\chi ^2`$ is given as $$\chi ^2=\underset{i=1}{\overset{N}{}}\frac{1}{\sigma _i^2}(F_{i,\mathrm{obs}}F_{i,\mathrm{theory}})^2,$$ (15) where $`F`$ for CMBR data is equal to $`\mathrm{\Delta }T`$ for different values of $`l`$. For BBN data, $`F`$ is equal to either the helium or the deuterium abundance. $`\sigma _i`$ is the uncertainty in the given data point. Here it has been assumed that the experimental errors are of Gaussian nature, which in reality they are not. Therefore the confidence limits that we obtain are not strict in a mathematical sense, but should rather be seen as indicative. If the measurements are furthermore assumed to be independent then the likelihood function is then given by $`Le^{\chi ^2/2}`$, and $$L_{\mathrm{joint}}=L_{\mathrm{CMBR}}L_{\mathrm{BBN}}L_{\mathrm{clusters}}.$$ (16) In Fig. 1 we show the likelihood contours taking into account different measurements. The likelihood contours have been based on $`\mathrm{\Delta }\chi ^2`$, such that the 68%, 95%, and 99% confidence limits correspond to $`\mathrm{\Delta }\chi ^2=2.29`$, 6.18 and 9.21 respectively. The top panel shows the likelihood contours for CMBR data alone. As expected it shows a very elongated structure, roughly on the $`\mathrm{\Omega }_mh^2`$ line. The middle panel shows how the likelihood contours are affected if one combines CMBR with the BBN estimate of $`\mathrm{\Omega }_bh^2`$. Clearly, there is only a quite small effect on the determination of $`\mathrm{\Omega }_m`$ and $`H_0`$, because $`\mathrm{\Omega }_b`$ is not very degenerate with either of these two parameters. The real improvement only comes when one takes into account the cluster baryon fraction measurements. Doing this narrows the likelihood contours to a more vertical and much tighter structure. Intriguingly, the combined likelihood function from this method favours a medium density universe with $`\mathrm{\Omega }_m`$ around 0.5 and a very low value of the Hubble parameter. This is strongly incompatible with the standard $`\mathrm{\Omega }_m=1`$ CDM model, unless the Hubble parameter is much lower than 0.3. However, as mentioned above our likelihood contours are somewhat optimistic, especially the upper bound on the Hubble parameter is sensitive to adding extra parameters to the analysis. It is also of interest to compare our results with the recent results from measurements of supernova type Ia’s. The best available data come from the supernova cosmology project, who have measured 42 high redshift supernovae \[Perlmutter et al. 1998\]. From these they obtain stringent limits on $`\mathrm{\Omega }_m`$ in a flat universe, but no constraint on the Hubble parameter. Their best fit value for $`\mathrm{\Omega }_m`$ in a flat universe is \[Perlmutter et al. 1998\] $$\mathrm{\Omega }_m=0.28_{0.08}^{+0.09}(\mathrm{statistical})_{0.04}^{+0.05}(\mathrm{systematic}).$$ (17) These data are compatible with our results, even at the $`1\sigma `$ level, but only for relatively low values of the Hubble parameter. Joint maximum likelihood for our analysis and the supernova data are shown in Fig. 2. The likelihood function has a maximum at $$(\mathrm{\Omega }_m,h)=(0.45,0.39),$$ (18) but with a fairly large uncertainty on the Hubble parameter. At the 95% confidence level the Hubble parameter is only constrained to be less than 0.64 and, again, we stress that adding more parameters to the analysis could weaken this bound further. ## 5 Conclusions We have calculated the possible constraints on $`\mathrm{\Omega }_m`$ and $`H_0`$ from combining CMBR data with data on the cosmic baryon abundance. It was found that these measurements are inconsistent with the standard flat CDM model, even for very low values of the Hubble parameter. Our favoured region of parameter space is consistent with recent measurements from type Ia supernovae at the $`1\sigma `$ level \[Perlmutter et al. 1998\]. The joint likelihood function suggests that $`\mathrm{\Omega }_m`$ is close to 0.4 and that a rather low value of the Hubble parameter is favoured. It is interesting to compare our results with those of Webster et al. \[Webster et al. 1998\], who did a joint analysis of the CMBR data and the IRAS 1.2 Jy data \[Fisher et al. 1995\] on large scale structure. They found a best fit for the joint likelihood of \[Webster et al. 1998\] $$(\mathrm{\Omega }_m,h)_{\mathrm{CMBR}+\mathrm{IRAS}}=(0.39_{0.10}^{+0.14},0.53_{0.14}^{+0.05}),$$ (19) which is completely compatible with our estimate of $$(\mathrm{\Omega }_m,h)_{\mathrm{present}\mathrm{analysis}}=(0.45_{0.07}^{+0.07},0.39_{0.13}^{+0.14}).$$ (20) In conclusion what we have shown in the present paper is that if the cosmic baryons are used in unison with CMBR data to constrain the cosmological parameters $`\mathrm{\Omega }_m`$ and $`h`$, the results are very close to other recent measurements using different methods. This suggests that there is a convergence towards the true, underlying, cosmological model, which is apparently a flat, low density universe. The fact that so many completely independent methods all yield roughly consistent estimates make it increasingly unlikely that the data are contaminated by some large, unknown, systematic error. Of course when the new high precision data become available from the BOOMERANG, MAP and PLANCK experiments, these uncertainties should be resolved.
no-problem/9906/hep-ph9906259.html
ar5iv
text
# Classical Sphaleron Rate on Fine Lattices ## 1 Introduction Baryon number is not a conserved quantity in the standard model. Rather, because of the anomaly, its violation is related to the electromagnetic field strength of the SU(2) weak group , $$_\mu J_B^\mu =N_G\frac{g^2}{32\pi ^2}ϵ_{\mu \nu \alpha \beta }\mathrm{Tr}F^{\mu \nu }F^{\alpha \beta }=N_G\frac{g^2}{8\pi ^2}E_i^aB_i^a,$$ (1) where $`N_G=3`$ is the number of generations.<sup>3</sup><sup>3</sup>3There is also a contribution from the hypercharge fields, but it will not be relevant here because the topological structure of the abelian vacuum does not permit a permanent baryon number change. The right-hand side of this equation is not surprisingly a total derivative, with an associated charge called the Chern-Simons number, $`{\displaystyle \frac{N_B}{N_G}}={\displaystyle \frac{1}{N_G}}{\displaystyle d^3xJ_B^0}`$ $`=`$ $`N_{\mathrm{CS}}(\mathrm{integer})+{\displaystyle \frac{g^2}{32\pi ^2}}{\displaystyle d^3xϵ_{ijk}\left(F_{ij}^aA_k^a\frac{g}{3}f_{abc}A_i^aA_j^bA_k^c\right)},`$ (2) $`N_{\mathrm{CS}}(t_1)N_{\mathrm{CS}}(t_2)`$ $``$ $`{\displaystyle _{t_2}^{t_1}}𝑑t{\displaystyle d^3x\frac{g^2}{8\pi ^2}E_i^aB_i^a}.`$ (3) Chern-Simons number $`N_{\mathrm{CS}}`$ has topological meaning; its change in a vacuum to vacuum process is the (integer) second Chern class of the gauge connection. Note that, according to the first equation, the total baryon number $`N_B`$ need only be an integer in vacuum, when $`N_{\mathrm{CS}}`$ is an integer. The baryon number in vacuum fixes the constant of integration in the definition of $`N_{\mathrm{CS}}`$. In vacuum, baryon number can be violated by a vacuum fluctuation large enough to have a nonzero integer Chern-Simons number. The efficiency of baryon number violation by this mechanism is totally negligible ; but at a sufficiently high temperature baryon number violation can proceed by thermal excitations which change $`N_{\mathrm{CS}}`$, and the rate for such a process is not necessarily very small . This can have very interesting cosmological significance, since it complicates GUT baryogenesis mechanisms and opens the possibility of baryogenesis from electroweak physics alone. This motivates a more careful investigation of baryon number violation in the standard model at high temperatures. The baryon number violation rate relevant in cosmological settings can be related by a fluctuation dissipation relation to the diffusion constant for Chern-Simons number, $$\mathrm{\Gamma }\underset{V\mathrm{}}{lim}\underset{t\mathrm{}}{lim}\frac{(N_{\mathrm{CS}}(t)N_{\mathrm{CS}}(0))^2}{Vt},$$ (4) where the angular brackets $``$ represent a trace over the thermal density matrix. This quantity is called the sphaleron rate for historical reasons. There has been some controversy not only to its size in the symmetric electroweak phase (which closely resembles pure Yang-Mills theory, an approximation we will make from here on) but even to its parametric dependence. On purely dimensional grounds, at high temperature it must scale as $`T^4`$, but the dependence on the coupling constant has been more controversial. Since the natural nonperturbative length scale of hot Yang-Mills theory at weak coupling is $`1/(g^2T)`$ it was long believed that $`\mathrm{\Gamma }\alpha _w^4T^4`$. In this case, $`\mathrm{\Gamma }`$ could equal its value in classical Yang-Mills theory, as originally suggested by Grigoriev and Rubakov . More recently Arnold, Son, and Yaffe (ASY) have argued that while the natural length scale for hot, weakly coupled Yang-Mills theory is $`1/(g^2T)`$, the natural time scale is different; because of interactions between the nonperturbative infrared excitations and essentially perturbative but very numerous UV excitations, the time evolution of IR Yang-Mills fields should be overdamped and the natural time scale for their evolution should be $`1/(g^4T)`$, or, restoring $`\mathrm{}`$, $`1/(\mathrm{}g^4T)`$. The appearance of $`\mathrm{}`$ in this expression precludes any simple correspondence between the classical theory and the quantum theory. They then argue that, since only the nonperturbative IR fields can contribute to the diffusion of $`N_{\mathrm{CS}}`$, the correct parametric behavior for $`\mathrm{\Gamma }`$ is $`\mathrm{\Gamma }\alpha _w^5T^4`$. The analogous expression in the classical theory has some UV regulator serving the role of $`\mathrm{}`$ in the quantum theory; for instance, for Yang-Mills theory on a lattice under the standard lattice action, the UV regulator scale is the inverse of the lattice spacing $`a`$, so the natural time scale should be of form $`1/(g^4aT^2)`$, leading to $`\mathrm{\Gamma }a\alpha ^5T^5`$ . Subsequently, Bödeker showed that the coefficient of the $`\alpha ^5`$ law should contain a further logarithmic dependence on $`g^2`$ , or, on the lattice, on $`g^2aT`$. The arguments of ASY are still considered somewhat controversial. While numerical simulations of classical Yang-Mills theory, supplemented with added degrees of freedom intended to serve the role of the “hard” quantum UV degrees of freedom, clearly support their arguments , results for $`\mathrm{\Gamma }`$ in pure classical Yang-Mills theory on the lattice have never convincingly displayed linear scaling in lattice spacing. Furthermore, the more recent work of Ambjørn and Krasnitz finds two other results which are problematic to interpret if Arnold, Son, and Yaffe are correct; they find overly strong lattice spacing dependence for $`\mathrm{\Gamma }`$ in small volumes, and they find unexpectedly rapid falloff for unequal time, Coulomb gauge fixed correlators . We will not address the question of unequal time, Coulomb gauge correlators here, except to note that we believe such correlators should show a strong volume dependence even at fixed $`k`$. Since the physical volume was varying along with the lattice spacing in Ambjørn and Krasnitz’ results, it is difficult to disentangle these two dependencies. We leave settling this problem to future work, but it is our general belief that such correlators will not prove very useful probes of infrared dynamics. This work is intended to answer the other two questions about the applicability of the ASY picture to classical Yang-Mills theory. First, we present results for $`\mathrm{\Gamma }`$ on a much wider range of lattice spacings, including much smaller spacings $`a`$, than has previously been done. Although the results at fairly large $`a`$ show weak lattice spacing dependence, as $`a`$ becomes smaller a strong $`a`$ dependence sets in, which is incompatible with $`\alpha ^4`$ scaling, but is in accordance with expectations if Arnold, Son, and Yaffe are correct. We also present a re-analysis of the behavior of $`\mathrm{\Gamma }`$ in a small, fixed volume. When care is taken to ensure that the physical volume really remains fixed as the lattice spacing is varied, we find $`\mathrm{\Gamma }`$ to depend on $`a`$ slightly more weakly than in large volumes. This is also expected in the ASY picture. ## 2 Expected scaling behaviors To explain the different proposed scaling behaviors for $`\mathrm{\Gamma }`$, we will briefly review the thermodynamics of classical Yang-Mills theory. The partition function for classical Yang-Mills theory is equivalent to the path integral for three dimensional, quantum Yang-Mills theory with an added adjoint scalar, as originally shown by Ambjørn and Krasnitz . The classical Yang-Mills partition function, absorbing $`g`$ into the definition of the connection so covariant derivatives are $`D_i=_i+iA_i`$, is $`Z`$ $`=`$ $`{\displaystyle 𝒟A_i^a𝒟E_i^a\delta ((D_iE_i)^a)\mathrm{exp}(H/T)},`$ (5) $`H`$ $`=`$ $`{\displaystyle d^3x\left(\frac{1}{4g^2}F_{ij}^aF_{ij}^a+\frac{1}{2}E_i^aE_i^a\right)}.`$ (6) The delta function enforces Gauss’ law. It is convenient to rewrite it by introducing a Lagrange multiplier, $$\delta ((D_iE_i)^a)=𝒟A_0^a\mathrm{exp}\left(\frac{i}{g}A_0^a(D_iE_i)^a\right),$$ (7) where $`A_0`$ has the same normalization as $`A_i`$. The integral over $`E`$ is then Gaussian. Performing it yields the partition function $`Z`$ $`=`$ $`{\displaystyle 𝒟A_i𝒟A_0\mathrm{exp}(H/T)},`$ (8) $`H/T`$ $`=`$ $`{\displaystyle d^3x\left(\frac{1}{4g^2T}F_{ij}^aF_{ij}^a+\frac{1}{2g^2T}(D_iA_0)^a(D_iA_0)^a\right)}.`$ (9) Physically the $`A_0`$ field corresponds to the time component of the connection, but if we choose to interpret the result as a path integral for a 3-D quantum field theory then the $`A_0`$ field just corresponds to some massless adjoint scalar. The form for the partition function coincides, except for the absence of a mass term for the $`A_0`$ field, with the path integral for the full quantum theory in the dimensional reduction approximation , and so it quite accurately reproduces the thermodynamics of infrared Yang-Mills fields. This motivates the belief that the dynamics of the full Yang-Mills theory will also coincide with the dynamics of the classical theory. Looking carefully at Eq. (9), we observe that $`g^2`$ and $`T`$ only enter in the combination $`g^2T`$. Now $`\mathrm{\Gamma }`$ has engineering dimension 4, so since $`T`$ is the only dimensionful quantity in hot Yang-Mills theory at weak coupling, we must have $`\mathrm{\Gamma }T^4`$. The “naive” scaling argument for $`\mathrm{\Gamma }`$ corresponds to requiring that each $`T`$ carry a factor of $`g^2`$, as is motivated by the form of the partition function. The problem with this argument, and with the argument that the dynamics of the classical and quantum theories should correspond, is that it neglects the effects of UV divergences. For thermodynamic quantities the UV divergences of the classical theory arise from a finite number of graphs and can be absorbed by counterterms. However, from the way that the $`A_0`$ field arose above, the classical theory is obliged to have a zero bare mass squared for this field. Hence the physical mass squared will approximately equal the linear one loop divergent contribution, $`m_D^2g^2T\mathrm{\Lambda }`$, with $`\mathrm{\Lambda }`$ some UV cutoff. In particular, on the lattice and using the conventional (Wilson) action, $`1/a`$ serves the role of $`\mathrm{\Lambda }`$, and <sup>4</sup><sup>4</sup>4The result in the reference differs by a factor of 5/4 because it includes a contribution from a Higgs field, which is absent here since we treat pure Yang-Mills theory. $$m_D^2=\frac{\mathrm{\Sigma }g^2T}{\pi a},\mathrm{\Sigma }=3.175911536\mathrm{}.$$ (10) The thermodynamics of the gauge fields $`A_i`$ do not care about this divergence, since their thermodynamics have a finite limit as $`m_D^2\mathrm{}`$. Therefore the argument that the only natural length scale for the $`A_i`$ fields is $`1/g^2T`$, remains valid in the presence of UV divergences. However, since the dynamics of the $`A_i`$ and $`A_0`$ fields are intertwined, it is not at all clear that unequal time phenomena involving the $`A_i`$ fields should be UV regulation insensitive. Arnold, Son, and Yaffe examined the propagator of the full quantum theory at soft momentum, including the UV influences at leading order in the coupling by including the hard thermal loop (HTL) self-energy contribution . For spatial momenta $`p`$ in the parametric regime $`pgT`$ they conclude that the evolution of the transverse (magnetic) degrees of freedom responsible for baryon number violation is overdamped, of form $$\frac{dA(p)}{dt}=\frac{4p^3}{\pi m_D^2}A(p)+(\mathrm{noise}).$$ (11) The perturbative treatment which gives this result holds provided that $`pg^2T`$, but it breaks down when $`p`$ is of order $`g^2T`$, the scale which we are interested in. Ignoring this difficulty and applying it for $`pg^2T`$, we find that the time scale for significant change in $`A`$ is $`m_D^2/g^6T^3`$, which in the continuum theory is $`1/(g^4T)`$ and on the lattice is $`1/(g^4aT^2)`$. Of course it is problematic to apply Eq. (11) beyond its range of applicability. However, for $`g`$ sufficiently small, we may apply it at the scale $`pg^{2ϵ}T`$ to find that the natural time scale here is $`(1/t)g^{43ϵ}T`$. Since the time scale for field evolution at $`pg^2T`$ cannot be faster than that at a larger value of $`p`$, Eq. (11) does place a bound on the natural time scale for fields with $`pg^2T`$; the time scale cannot differ from the estimate $`1/(g^4T)`$ by a nonzero power of $`g`$. Thus, Eq. (11) is enough to ensure the ASY result up to corrections weaker than any power of $`g`$. A much more careful study of the HTL effective theory by Bödeker finds that $`\mathrm{log}(1/g)`$ corrections do occur; however, they prove to be numerically small . Note however that the ASY argument is a parametric treatment which relies on a large separation between the $`gT`$ and $`g^2T`$ scales. Numerical results, for instance for the subleading contributions to the Debye screening length , suggest that $`g^2`$ (or, on the lattice, $`g^2aT`$), may need to be fairly small before such a separation of scales really exists. Therefore, we might expect that the ASY scaling behavior only sets in at reasonably small lattice spacings $`a`$. This motivates the study of the sphaleron rate on very fine lattices, which we take up in the next section. ## 3 Results: large lattices, fine spacings To address the question: “How does $`\mathrm{\Gamma }`$ depend on lattice spacing $`a`$ when $`a`$ is small?” we track $`N_{\mathrm{CS}}`$ on the lattice. We use the conventional Kogut-Susskind lattice action , which is the Minkowski time analog of the Wilson action for Euclidean lattice Yang-Mills theory. We use the same discrete implementation for the time evolution as in Ambjørn et. al. and almost all subsequent work. Besides the size of the lattice there is one variable, $`\beta _\mathrm{L}`$, which is a reciprocal temperature in lattice units. At tree level it is $`\beta _\mathrm{L}=4/(g^2aT)`$, but this relation receives radiative corrections because the UV lattice and continuum fields behave differently. These are treated, for the case $`m_D^21/a^2`$, in and extended to larger $`m_D`$ in Appendix B of . We use the expression from there for the one loop improved relation between $`\beta _\mathrm{L}`$ and lattice spacing $`a`$, $`\beta _\mathrm{L}`$ $`=`$ $`{\displaystyle \frac{4}{g^2aT}}+\left({\displaystyle \frac{1}{3}}+{\displaystyle \frac{37\xi }{6\pi }}\right)\left[\left({\displaystyle \frac{4}{3}}+{\displaystyle \frac{2a^2m_D^2}{3}}+{\displaystyle \frac{a^4m_D^4}{18}}\right){\displaystyle \frac{\xi (am_D)}{4\pi }}\left({\displaystyle \frac{1}{3}}+{\displaystyle \frac{a^2m_D^2}{18}}\right){\displaystyle \frac{\mathrm{\Sigma }(am_D)}{4\pi }}\right]`$ $`\beta _\mathrm{L}`$ $``$ $`{\displaystyle \frac{4}{g^2aT}}+0.6\mathrm{or}a{\displaystyle \frac{4}{(\beta _\mathrm{L}0.6)g^2T}},`$ (12) where $`\xi =0.152859\mathrm{}`$ and the functions $`\xi (x)`$ and $`\mathrm{\Sigma }(x)`$ are defined in ; the approximation that the correction term is $`0.6`$ is good to $`10\%`$ for all $`a`$ considered here but we will use the full expression. In this section the difference between the naive and improved match will not be important, but in the next section it is essential. Henceforth, when we refer to $`\beta _\mathrm{L}`$ we will mean the “unimproved” quantity appearing in Eq. (3); while the quantity $`\beta `$ will mean $`\beta 4/(g^2aT)`$ using the improved relation for $`a`$. The measurement of Chern-Simons number deserves some comment. Early work on Chern-Simons number diffusion used “naive” definitions of $`N_{\mathrm{CS}}`$ in which the right hand side of Eq. (3) is implemented as the integral of a local operator over the unmodified lattice fields. This approach is not topological because such a local operator on the lattice is never a total derivative, as it should be for $`N_{\mathrm{CS}}`$ to have topological meaning. Because of this, such a definition of $`N_{\mathrm{CS}}`$ shows diffusive behavior even when there is no true diffusion of baryon number occurring . The response to true topology change can also get renormalized by UV fluctuations. The latter problem gets less severe as $`a`$ is reduced, but we find that the “spurious” diffusion per physical 4-volume grows as $`a^1`$ and is therefore disastrous. Therefore we should use a topological definition of $`N_{\mathrm{CS}}`$. Technically topology is not well defined for lattice fields, but it can be well defined on a restricted class of lattice fields which are sufficiently “smooth” . In our context topology is well defined for suitably small lattice spacing; in practice there is no problem if $`\beta 6`$ $`(a<2/3g^2T)`$, which will be the case for almost all lattices we consider. For the current application, two topological means have been developed; the “slave field” method , similar to the method of Woit ; and “calibrated cooling” , an improvement on the field smearing proposal of Ambjørn and Krasnitz . The “slave field” method is numerically efficient but noisy, which means in practice that a longer numerical evolution is needed to get good statistics. Therefore we will use “calibrated cooling”. The philosophy of the method is that the topological content of the connection cannot be modified by small local changes; therefore we may “smear” the connection, removing UV noise which is responsible for the poor performance of the “naive” measurement method. After the smearing we measure $`N_{\mathrm{CS}}`$ by integrating an $`O(a^2)`$ improved local operator for $`E_i^aB_i^a`$, and we cure the slight residual error in the algorithm by “calibrating” it with occasional coolings all the way to vacuum; the full details can be found in . The problem with the above approach is its numerical cost; measuring $`N_{\mathrm{CS}}`$ as a function of time is much more expensive than generating the Hamiltonian trajectory. This is only really a problem on very fine lattices, as numerical cost rises as $`a^4`$. However, in this case the topology changing configurations we are after are many, many lattice spacings across and are highly over-resolved; no topological information is lost by “blocking” the lattice and then measuring $`N_{\mathrm{CS}}`$. That is, we construct a “blocked” lattice with $`B`$ times the original lattice spacing by making the connection between neighboring points on the blocked lattice equal to the product of links between the same points on the unblocked lattice; the details are in . We do this first, before applying any smearing. This introduces some white noise into $`N_{\mathrm{CS}}`$ but does not change the diffusion at all, which means that results for $`\mathrm{\Gamma }`$ will be unchanged. Figure 1 shows a test of blocking in which we track $`N_{\mathrm{CS}}`$ for the same Hamiltonian trajectory, with and without blocking. The difference is small and purely spectrally white. On the other hand, the difference in numerical effort is enormous; without blocking, measuring $`N_{\mathrm{CS}}`$ took 5 times as much CPU time as updating the fields, while with blocking it took around $`10\%`$ as much. We will block for all data with $`a1/3g^2T`$ in this work, which means that the CPU time taken to measure $`N_{\mathrm{CS}}`$ is negligible for all the most numerically intensive cases. We should also be sure to use a large enough volume to eliminate finite volume systematics (effectively, to perform the $`V\mathrm{}`$ limit in Eq. 4). To do so we measure $`\mathrm{\Gamma }`$ as a function of $`L=V^{1/3}`$ at a fixed lattice spacing, $`a=1/2g^2T`$. The result is shown in Figure 2. The dependence on $`L`$ is in general accordance with the results of , except at small volumes where ours go to zero and theirs do not because of the spurious UV contributions in their definition of $`N_{\mathrm{CS}}`$. We will use $`L=10/g^2T`$ ($`N=2.5\beta `$) for all large volume results. Our results at large volumes are presented in Table 2 and plotted in Figure 3. These constitute the main result of this paper. While $`\mathrm{\Gamma }`$ is not a strong function of lattice spacing around $`a0.5/g^2T`$ ($`\beta 8`$), it then turns over and falls roughly linearly in $`a`$ at finer lattices. This indicates that the lattice spacing needed before the ASY scaling behavior sets in is around $`\beta =12`$ ($`a=1/3g^2T`$). If we insist on believing that the small $`a`$ scaling behavior is of form $`c_1+c_2g^2aT`$ with $`c_2`$ representing a correction to scaling, then a fit to the points with $`a1/3g^2T`$ ($`\beta 12`$) gives $`c_1=0.257\pm .044`$ and $`c_2=3.29\pm .18`$ ($`\xi ^2/\nu =5.4/3`$). The “correction” to scaling only becomes subdominant below $`a=1/12.8g^2T`$ ($`\beta >51`$). The original motivation for believing in a finite small $`a`$ limit for $`\mathrm{\Gamma }`$ is the belief that the UV lattice behavior is not important to the infrared dynamics. Such an enormous correction to scaling contradicts this belief. This makes it very difficult to reconcile our data with a finite small $`a`$ limit for $`\mathrm{\Gamma }`$. On the other hand, if ASY are correct, it makes more sense to plot our results as $`\mathrm{\Gamma }/(\alpha ^5aT^5)`$, as we do in Figure 4. The extrapolation to small $`a`$ looks much better behaved here. If we fit $`\mathrm{\Gamma }`$ to the form $`\mathrm{\Gamma }=\alpha ^5aT^5(c_1+c_2g^2aT)`$ we get $`c_1=74.5\pm 3.4`$ and $`c_2=73\pm 12`$ ($`\xi ^2/\nu =1.6/3`$), which means the correction to scaling comes of order 1 at $`a=g^2T`$ ($`\beta =4`$). In fact, as Bödeker has shown, we should not expect a finite intercept in this figure, there should be a weak logarithmic divergence as $`a0`$ . In the continuum its amplitude is $$\mathrm{log}\mathrm{part}\mathrm{of}\mathrm{\Gamma }=(10.7\pm .7)\frac{g^2T^2}{m_D^2}\mathrm{log}\left(\frac{m_D}{g^2T}\right)\alpha ^5T^4.$$ (13) The appearance of $`1/m_D^2`$ is changed somewhat on the lattice; $`m_D^2`$ is reduced by a weighted average over $`k`$ of the group velocity under lattice dispersion relations, which is about $`0.68`$ . Combining this with the expression for $`m_D^2`$ on the lattice, Eq. (10), gives $$\mathrm{log}\mathrm{part}\mathrm{of}\mathrm{\Gamma }=\left[7.8\pm .5\right]\mathrm{log}\left(\frac{1}{g^2aT}\right)\alpha ^5aT^5.$$ (14) Strictly speaking such logarithmic behavior only pertains in the regime where $`\mathrm{log}(1/g^2aT)1`$, which is probably not satisfied at any conceivable lattice spacing. Nevertheless we perform the fit to see what happens. The fit in Fig. 4 which curves assumes $`\mathrm{\Gamma }=\mathrm{log}\mathrm{part}\mathrm{of}\mathrm{\Gamma }+\alpha ^5aT^5(c_1+c_2g^2aT)`$, which still has only two free fitting parameters. The best fit value is $`c_1=54.6\pm 3.4`$, $`c_2=39\pm 12`$ ($`\xi ^2/\nu =2.1/3`$). In this case the leading $`c_1`$ behavior dominates over the scaling correction out to $`a=1.4g^2T`$ ($`\beta =3`$). Note that the data do not yet justify either believing or disbelieving in the log behavior. This is because the coefficient of the log is numerically small. ## 4 Small volumes Ambjørn and Krasnitz report another puzzling problem with the ASY picture . They analyzed the dependence of $`\mathrm{\Gamma }`$ on lattice spacing at a fixed, small volume, small enough that $`\mathrm{\Gamma }`$ is much smaller than its large volume limit. The idea was that, in such a small volume, $`N_{\mathrm{CS}}`$ fluctuates about an integer value with occasional, abrupt changes from integer to integer, which can be identified even with the naive definition of $`N_{\mathrm{CS}}`$. Therefore, $`N_{\mathrm{CS}}`$ can be tracked topologically without needing a true topological definition. Their results, replotted here in Figure 5, are puzzling. Not only is there a strong lattice spacing dependence in evidence; it is too strong. The rate appears to fall off faster than $`\mathrm{\Gamma }a`$, and certainly faster than the rate falls off in large volumes, where the corrections to scaling at finite $`a`$ make the $`a`$ dependence somewhat weaker than linear. To see why this result does not jive with the ASY picture, we will review again the basic ASY argument. Examining the propagator for the gauge field at momentum $`p`$ gives an equation of motion for $`A(p)`$ which, viewed on suitably long time scales, is approximately $$\frac{d^2A(p)}{dt^2}+\frac{\pi m_D^2}{4p}\frac{dA(p)}{dt}=p^2A(p).$$ (15) The damping term, with the single derivative, is more and more important as $`p`$ gets smaller. We expect that, in a constrained volume, the field configurations responsible for changing $`N_{\mathrm{CS}}`$ are spatially smaller than in a large volume. Therefore they are composed of excitations of larger $`p`$, and should be more weakly damped. In particular it should take a larger value of $`m_D^2`$ (smaller $`a`$) for the overdamped regime to apply, and the corrections to linear in $`a`$ scaling should be larger. The problem lies in the data. Ambjørn and Krasnitz worked in a lattice with $`N=\beta _\mathrm{L}`$, which at the unimproved level means $`L=4/g^2aT`$. However, they used the tree relation between the lattice spacing and the reciprocal temperature $`\beta _\mathrm{L}`$. As can be seen from Eq. 3, the true lattice spacing is larger than the tree relation indicates, and the error is worse as the lattice is made coarser. Therefore, their larger $`a`$ lattices possessed more physical volume than their smaller $`a`$ lattices. The effect is fairly small; the difference in linear dimension between their $`\beta _\mathrm{L}=10`$ and $`\beta _\mathrm{L}=20`$ lattices is only about $`3\%`$. However, as Fig. 2 shows, $`\mathrm{\Gamma }`$ is a very strong function of volume in the regime where they were working. Around $`N=\beta `$ ($`L=4/g^2T`$), the data in the figure give roughly $`d(\mathrm{log}\mathrm{\Gamma })/d(\mathrm{log}L)10`$, so a $`3\%`$ change in length could lead to a $`30\%`$ change in $`\mathrm{\Gamma }`$, which is significant. In fact this effect could be as large as or larger than the $`a`$ dependence due to hard thermal loop dynamics. To fix this problem we recompute $`\mathrm{\Gamma }`$ in a fixed, small volume, but using the improved relation between the lattice spacing and the reciprocal temperature. We choose to use a volume equivalent (at the improved level) to the volume of Ambjørn and Krasnitz’ finest volume, $`\beta _\mathrm{L}=20`$ and $`N=20`$, which, using the improved relation, gives $`L=4.114/g^2T`$. The results are presented in Figure 5, and include a re-calculation of the finest lattice point, using a topological definition of $`N_{\mathrm{CS}}`$ (rather than counting integer winding changes by eye from the old “unimproved” $`N_{\mathrm{CS}}`$ definition). The new data show a lattice spacing dependence similar to the large volume case, although to make a good comparison we would need data at smaller $`a`$ (larger $`\beta `$). The behavior of $`\mathrm{\Gamma }`$ in small volumes is in accord with what we expect if the ASY argument is correct. ## 5 Conclusion Chern-Simons number diffusion in pure Yang-Mills theory follows the Arnold-Son-Yaffe scaling behavior, $`\mathrm{\Gamma }a`$; however it takes a fairly fine lattice to demonstrate this in a convincing way. The same behavior, with similar corrections to scaling, is observed in small volumes, but only after taking care to keep the physical volume constant beyond leading (tree) level while the lattice spacing is varied. The original goal of measuring $`\mathrm{\Gamma }`$ on the lattice was to determine the rate at which a baryon number excess, present before the electroweak phase transition, would be erased. We can do this with the results of this paper by using Arnold’s study of the matching between the lattice and continuum theories . He shows that $`\mathrm{\Gamma }`$ on the lattice matches the continuum value when $$(.68\pm .20)m_D^2(\mathrm{latt})=m_D^2(\mathrm{continuum}),$$ (16) where the error is all systematic, but the error estimate is considered generous . In the minimal standard model (MSM), at leading order $`m_D^2=(11/6)g^2T^2`$, which for $`g^2=0.42`$ and using Eq. (10), means the MSM value is obtained on a lattice with $`a=.157/g^2T`$, or $`\beta =25.4`$. Our data at $`\beta =24`$ give $`\mathrm{\Gamma }=0.85\alpha ^4T^4`$, with an error insignificant compared to the $`30\%`$ error estimate in the lattice to continuum matching procedure. since $`\mathrm{\Gamma }a`$ is well satisfied in this regime, our estimate for the Standard Model value of $`\mathrm{\Gamma }`$ is $$\mathrm{\Gamma }(\mathrm{MSM})=(0.82\pm .24)\alpha ^4T^4\mathrm{or}\mathrm{\Gamma }=(45\pm 13)\left(\frac{g^2T^2}{m_D^2}\right)\alpha ^5T^4,$$ (17) where the latter form shows the correct parametric dependence on the Debye mass. This result is in good agreement with results obtained when classical Yang-Mills theory is supplemented with particle degrees of freedom to induce the hard thermal loop effects. Therefore the diffusion constant for $`N_{\mathrm{CS}}`$ in classical Yang-Mills theory is consistent with both analytic expectations and numerical results obtained by explicit inclusion of hard thermal loops.
no-problem/9906/cond-mat9906003.html
ar5iv
text
# Two-dimensional behavior of the sublattice magnetization in three dimensional Ising antiferromagnets ## 1 Introduction Anisotropic three-dimensional antiferromagnets (AFs) are still an interesting subject of theoretical investigations. There exist many investigations e.g. on their magnetic phase diagram within the framework of Ising or anisotropic Heisenberg models.$`^{}`$ It is found, that layered Ising AFs with two competing interaction parameters (Fig. 1) may exhibit two rather different phase diagrams (Fig. 2) depending only on two parameters, the ratio $`r=zJ/z^{}J^{}`$ and $`z^{}`$, where $`J`$ and $`J^{}`$ are the coupling constants of the intra-layer and of the inter-layer exchange, respectively; $`z`$ and $`z^{}`$ are the coordination numbers of the couplings. The Ising Hamiltonian is of the form: $$=J\underset{<i,j>}{}S_iS_jJ^{}\underset{<i,j>}{}S_iS_jH\underset{i}{}S_i,$$ (1) where $`H`$ is the applied magnetic field acting on all spins $`S_i`$, with $`S_i=\pm 1`$. The two kinds of phase diagrams are continuously transformed into one another by changing the crucial parameters $`r`$ and $`z^{}`$. For large values of $`|r|`$, i.e. $`|r|>0.6`$, and small values for $`z^{}`$, i.e. $`z^{}<10`$, the FeCl<sub>2</sub>-like phase diagram is found (Fig. 2 (b)), whereas for low values of $`|r|`$ and large values for $`z^{}`$ the case (b) in Fig. 2 appears. This second case is characterized by three interesting features. On the one hand, a possible decoupling of the tricritical point (TCP) into a critical endpoint (CEP) and a bicritical endpoint (BCE) is encountered. From previous investigations it follows, that this decoupling is only observed in mean field calculations,$`^{}`$ while in Monte Carlo simulations only one multicritical point is found. On the other hand, the second-order phase line has a balloon-like shape and extends even above the limiting field value $`H_{c0}`$ (spin-flip field at $`T=0`$). This means, that for some fixed field values $`H>H_{c0}`$ it is possible to cross this phase line twice with increasing temperature $`T`$. Furthermore, above and below the critical line anomaly lines are found, where the magnetization exhibits an additional inflection point and the specific heat shows an additional broad maximum. ## 2 Comparison between theory and experiment Fig. 3 (a) shows the $`H`$-$`T`$ phase diagram of a 3-dimensional layered Ising AF obtained from Monte-Carlo simulations on a hexagonal lattice with periodic boundary conditions. All quantities like $`T`$, $`H`$, $`J`$ and $`J^{}`$ are considered to be dimensionless. For the simulation the Metropolis algorithm was used with $`k_B=1`$. The parameters $`z,z^{},J,J^{}`$ are chosen such as to reproduce case (b) of Fig. 2. One observes only one multicritical point (MCP), where a first-order phase line ($`T<T_{MCP}`$) and a second order phase line, $`H_c(T)`$ meet, and where the two anomaly lines, $`H_{}`$ and $`H_+`$, originate. Fig. 3 (b) shows one magnetization curve, referring to the phase diagram of Fig. 3 (a), as a function of temperature for $`H=9.5`$ and its derivative, $`dM/dT`$. The magnetization is defined to be $`M=_iS_i/N`$, where $`N`$ is the number of lattice sites. $`M`$ vs $`T`$ clearly shows an anomalous curvature for $`T<T_c`$, which manifests itself in the derivative as an additional broad maximum. The phase diagram shown in Fig. 3 (a) resembles that of FeBr<sub>2</sub>, an insulating uniaxial antiferromagnet with $`T_N=14.2`$$`^{}`$ (Fig. 4). Both in Fig. 3 and Fig. 4 lines of non-critical fluctuations (or anomaly lines) do appear. In this article the attention is focused on these non-critical fluctuations at $`T_{}=T(H_{})`$. Although this phenomenon is well investigated both in experiments on FeBr<sub>2</sub> and in theory$`^{}`$ no clear explanation for the occurrence of these fluctuations yet exists. Especially one wonders, how it is possible that two completely different types of phase diagrams are found by varying only two parameters, $`r`$ and $`z^{}`$. ## 3 Non-critical fluctuations In order to gain insight into the origin of the non-critical fluctuations at $`H_{}(T)`$, we performed systematic Monte Carlo simulations of the sublattice magnetizations with the same parameters as in Fig. 3, $`zJ=4.2`$ and $`z^{}J^{}=10.0`$. The exchange constants and especially the number of coupled neighbours are comparable to those in FeBr<sub>2</sub>. However, here we used only two exchange couplings. The different intra-planar exchange parameters found in FeBr<sub>2</sub> were absorbed in one ferromagnetic effective intra-planar coupling constant. Fig. 5 shows the temperature dependences of the sublattice magnetizations, $`M_A`$ and $`M_B`$ (see Fig. 1), for $`H=0`$, $`4`$, $`8`$, $`9.5`$ and $`9.95`$, respectively. While $`M_A`$ (parallel) and $`M_B`$ (antiparallel to $`𝐇`$) are symmetric for zero field, $`M_A(T)=M_B(T)`$, they become more and more inequivalent with increasing field. For fields coming close to the spin-flip one, $`H_{c0}=10`$, anomalous bumps appear at $`TT_{}`$ in $`M_B(T)`$, whereas $`M_A(T)`$ is virtually constant up to that temperature. Obviously all of the non-critical fluctuations observed at $`TT_{}`$ happen to occur merely on the B-sublattice. In other words, these fluctuations are essentially constrained to 2-dimensional (2d) layers separated by magnetically saturated up-spin layers of the A-sublattice. It is, hence, tempting to compare the B-sublattice with an ensemble of 2d ferromagnets (FMs) with the same intra-layer parameters, $`zJ>0`$, but subjected to a field $`H_{eff}=HH_{c0}`$, where $`H`$ is the field applied to the corresponding 3d AF. Fig. 6 shows the sublattice magnetization curves for $`H=9.5`$ and $`9.95`$ as before and the magnetization curves of a 2d FM with $`zJ=4.2`$ in fields $`H_{eff}=0.5`$ and $`0.05`$, respectively. One observes, that the magnetization curve of the 2d FM fits well with the spin-down sublattice magnetization $`M_B`$ up to the inflection point at $`T_{}`$ (arrows) . It seems, that the magnetization of the spin-down sublattice behaves like a 2d FM in an effective mean field, which is the sum of the external applied field $`H`$ and the field, produced by the fully magnetized spin-up sublattice (A-sublattice), $`H_{eff}=H+H_A`$. Since $`M_A`$ is completely magnetized, $`H_A=z^{}J^{}=H_{c0}`$, one has $`H_{eff}=HH_{c0}`$. If the effective field $`H_{eff}`$ becomes zero ($`H_{eff}=0H=H_{c0}`$) we have the case of a 2d FM in zero field, which undergoes a phase transition at $`T_c^{2d}`$ and becomes paramagnetic for $`T>T_c^{2d}`$ (see Fig. 6). Although this case cannot be obtained here, because of the spin-flip occurring in the range $`H_{MCP}<H<H_{c0}`$, the anomaly temperatures $`T_{}`$ can be associated with the points of inflection of the $`M(H_{eff})`$ vs $`T`$ curves of the 2d Ising FM (arrows in Fig. 6). Therefore one can conclude, that the anomaly of $`M_B`$ singnifies the thermal destruction of 2d ferromagnetic order on the quasi-decoupled B-sublattice layers, which precedes the global 3d phase transition of both sublattices. In other words, the anomaly we find in the magnetization is due to the finite temperature range lying between $`T_c^{2d}`$ and $`T_c^{3d}`$. If this splitting is reduced by increasing the intra-planar ferromagnetic interaction and thus increasing the 2d transition temperature, the anomaly decreases and vice versa. ## 4 Conclusion The occurrence of anomalies, which are observed in certain 3d Ising AFs is due to the separation of the smeared 2d phase transition on one sublattice from the 3d global phase transition of both sublattices. This is typical of antiferromagnets, whose inter-planar antiferromagnetic coupling is strong compared with the intra-planar exchange ($`|r|=|zJ/z^{}J^{}|<0.6`$). A high value of the inter-planar exchange gives rise to a high value of the spin-flip field $`H_{c0}=z^{}J^{}`$. This results in a strong stabilization of the spin-up sublattice, which acts only as a mean-field $`H_{MF}=z^{}J^{}`$ on the spin-down sublattice. Hence, the spin-down sublattice behaves like a 2d FM in an effective field $`H_{eff}=H+H_{MF}`$. The quality of the mean-field approximation crucially depends on a high coordination number $`z^{}`$, which is needed for integrating out thermal spin-fluctuations. ## Acknowledgments We would like to thank M. Acharyya and U. Nowak for helpful discussions and the Deutsche Forschungsgemeinschaft (Graduiertenkolleg ”Struktur und Dynamik heterogener Systeme”) for financial support.
no-problem/9906/hep-ph9906236.html
ar5iv
text
# Bicocca–FT–99–15Cavendish-HEP-99/06hep-ph/9906236 1/𝑄² corrections in DIS fragmentation functions ## 1 INTRODUCTION The work described in the following article is largely motivated by the recent experimental interest in the study of hadronic final state momentum distributions in the current hemisphere of the Breit frame of DIS. The observable in question is similar to the hadronic energy distribution in $`e^+e^{}`$ annihilation and is related in an analogous manner to the universal parton to hadron fragmentation functions. However in order to compare theoretical predictions for these DIS momentum distributions accurately with the HERA data, we must take into account the existence of $`1/Q^2`$ non-perturbative effects (power corrections) from both the singlet and non-singlet processes. We wish to study the distribution $`F^h(z)`$, defined by $$F^h(z;x,Q^2)=\frac{d^3\sigma ^h}{dxdQ^2dz}/\frac{d^2\sigma }{dxdQ^2},$$ (1) for a given hadron species $`h`$, as a function of the variable $`z=2p_hq/q^2`$, where $`p_h`$ is the four-momentum of the resultant hadron and $`q`$ the photon four-momentum. Since the fragmentation products of the remnant of the nucleon are expected to be travelling in directions close to that of the incoming nucleon, i.e. in the ‘remnant hemisphere’ $`p_z0`$, we consider only hadrons produced in the ‘current hemisphere’ $`p_z0`$. Such hadrons are expected to be fragmentation products of the scattered parton. Thus $`z`$ takes values $`0z1`$. The other quantities $`x`$, $`Q^2`$ have their usual DIS meanings. One has to consider power corrections to both the numerator and the denominator of Eq. (1) each of which have a leading power correction that goes as $`1/Q^2`$. The denominator is just the familiar DIS cross section which can be decomposed in terms of transverse and longitudinal structure functions for which the results are presented in Refs. and . For the numerator one has an identical decomposition in terms of generalised ($`z`$ dependent) structure functions to which we also compute power corrections. We present here, results for both singlet and non-singlet contributions to $`F^h(z;x,Q^2)`$. The calculations described here were done using the dispersive technique which we briefly comment on below. ## 2 CALCULATIONS AND RESULTS The calculation for the non-singlet piece follows the usual dispersive method first described in . In this method power corrections correspond to non-analytic pieces in a fake gluon mass $`ϵ`$, in the $`𝒪(\alpha _s)`$ calculation of a given observable, represented by the characteristic function $`(ϵ)`$. Importantly, the normalisation of the result is in terms of hopefully universal parameters, which turn out to be various moments of a universal strong coupling $`\alpha _s(\mu ^2)`$, defined for all scales $`\mu ^2`$. In order to calculate power-behaved contributions to QCD observables involving more than a single renormalon chain, like the singlet sector here, we need to generalise the standard dispersive treatment, the details of which can be found in references . Here we just mention that in cases involving for instance two gluons, like the forward scattering amplitude of the DIS singlet graphs, we need to introduce two dispersive variables (gluon masses) and our characteristic function $`(ϵ_1,ϵ_2)`$ is now an $`𝒪(\alpha _s^2)`$ quantity. However it turns out that the answer can be expressed in terms of functions $`\widehat{}`$ which are obtained from the characteristic function by setting one of the gluon masses to zero while retaining the other. Hence it is still the non-analytic terms in a single variable $`ϵ`$ which correspond to power corrections. In the present case (DIS fragmentation functions) we find the following types of non-analyticity which translate into $`1/Q^2`$ corrections as below: $$\widehat{}cϵ\mathrm{log}ϵ\delta F=c\frac{D_1}{Q^2}$$ (2) and $$\widehat{}\frac{1}{2}c^{}ϵ\mathrm{log}^2ϵ\delta F=c^{}\frac{D_1}{Q^2}\mathrm{log}\frac{D_2}{Q^2},$$ (3) where $`D_1`$ and $`D_2`$ are defined by $$D_1_0^{\mathrm{}}𝑑\mu ^2\left[2\alpha _s\delta \alpha _s(\delta \alpha _s)^2\right]$$ (4) $$\mathrm{log}D_2\frac{1}{D_1}_0^{\mathrm{}}𝑑\mu ^2\mathrm{log}\mu ^2\left[2\alpha _s\delta \alpha _s(\delta \alpha _s)^2\right].$$ (5) and $$\delta \alpha _s(\mu ^2)=\alpha _s(\mu ^2)\alpha _s^{\text{PT}}(\mu ^2)$$ represents a non-perturbative modification to the standard perturbative form of thr coupling. We emphasise the fact that in the above formula $`\alpha _s^{\text{PT}}`$ stands for the perturbative expansion of $`\alpha _s(\mu ^2)`$ in terms of $`\alpha _s(Q^2)`$, where this expansion is truncated at whatever order the perturbative result for a given observable has been computed. Hence in most cases of current interest this would be at $`𝒪(\alpha _s^2)`$. We calculate the singlet and non-singlet contributions to the fragmentation functions, retaining terms up to $`𝒪(ϵ)`$ non-analytic at $`ϵ=0`$. We find terms that diverge as $`\mathrm{log}ϵ`$ and $`\mathrm{log}^2ϵ`$ arising from collinear splitting; the gluon mass here behaves as a regulator, and the divergence is factored into the scale dependence of the parton distributions. The $`1/Q^2`$ corrections arise from terms proportional to $`ϵ\mathrm{log}ϵ`$ and $`ϵ\mathrm{log}^2ϵ`$. In the singlet case we obtain the correction $$\delta F^h=\frac{D_1}{Q^2}\frac{T_RC_F}{(2\pi )^2}\left[K_T+\frac{2(1y)}{1+(1y)^2}K_L\right]$$ (6) where $`K_T(z,x)`$ and $`K_L(z,x)`$ are shown in figure 2 and $`y`$ is the usual DIS variable. Note that if $`D_1`$ is positive, we would expect to see a large negative power correction at small $`z`$. In non-singlet scattering we find the relative correction $$\frac{\delta F^h}{F^h}=\frac{𝒜_2}{Q^2}H(z,x)$$ (7) where $`H(z,x)`$ is shown by figure 1. This gives a large positive correction at large $`z`$, but only a small correction at small $`z`$. The detailed calculations and algebraic expressions for results can be found in . ## 3 CONCLUSIONS AND PROSPECTS The results presented in this article clearly indicate that at small values of the fragmentation variable $`z`$, one would expect the singlet contribution to be dominant. Moreover at small values of the Bjorken variable $`x`$, we would anyhow expect the singlet scattering mechanism to start becoming important due to the rise in the gluon density. The parameter $`D_1`$ that appears in the singlet result has to be phenomenologically determined (extracted once and for all from the data). Results are relatively insensitive to the other parameter $`D_2`$. Unfortunately however, before one can compare the predictions made here with the experimental data one needs to clarify several issues. From a technical viewpoint one has to learn how to treat the singlet piece as a gluonic contribution rather than an $`𝒪(\alpha _s^2)`$ non-singlet contribution. In other words, the result presented here was obtained by treating the incoming gluon (that scatters via $`q\overline{q}`$ production off the photon) as an internal line radiated off an incoming quark. This is the only way to obtain the power correction. However to reinterpret this as a singlet contribution we have to find a meaningful way to factorise off the quark to gluon splitting and then convolute with the gluon density rather than the quark density. Then there is also the question of doing improved perturbative calculations. At present the NLO program CYCLOPS is used to obtain the perturbative prediction. However in actual fact, at small values of $`z`$ we would expect terms that vary as $`\mathrm{log}(1/z)`$ to spoil the convergence of the perturbative series. Hence the need arises for resummed perturbative estimates which take these terms into account at all orders. Another relatively minor consideration is that for the present results we calculate the distribution differential in $`z`$, the hadron longitudinal momentum fraction, rather than the energy fraction, since it turns out to be complicated to analytically compute the latter in DIS. Hence when comparisons are made with data one must make sure that we are using the same fragmentation variable although qualitative features can be expected to be similar for both choices. In fact at this conference we saw experimental results for the energy distributions where the deviation from NLO perturbation theory was indeed large and negative as we expect at small $`z`$ . Remarkably the data can be extremely well described using a simple power correction ansatz of Dokshitzer and Webber also presented here. However the ansatz is not a theoretical prediction in any strict sense but rather an educated guess at the shape of the correction. Hence for a full theoretical picture we need to understand the problems mentioned above. Once this is done we can proceed with extracting $`D_1`$ from the data and using it in fits to other singlet observables.
no-problem/9906/math9906181.html
ar5iv
text
# Lifting units modulo exchange ideals and 𝑪^∗-algebras with real rank zero ## Introduction The problem of lifting units from a quotient of a ring $`R`$ modulo a two-sided ideal $`I`$ has been of interest in several instances. The first extension of the classical index theory for Fredholm operators on a Hilbert space was directed to von Neumann algebras (see , and ). The class of self-injective rings and that of Rickart $`C^{}`$-algebras satisfying certain comparability conditions were considered by Menal and Moncasi (see ). The general case for Rickart $`C^{}`$-algebras was studied by Ara in . In all of the above cases, the extent to which a unit from a quotient $`R/I`$ can be lifted to a unit of $`R`$ is measured by a condition of $`K`$-theoretic nature, namely the vanishing of the connecting index map in $`K`$-Theory. Our aim here is to consider the class of exchange ideals of unital rings, which is known to contain both (not necessarily unital) von Neumann regular rings and $`C^{}`$-algebras with real rank zero. In fact, the exchange $`C^{}`$-algebras are exactly those having real rank zero (see Theorem 7.2 in ). A second unifying principle on which we will rely is that of separative cancellation of finitely generated projective modules, which can be regarded as a weak cancellation property. Separative unital exchange rings provide a framework in which a number of outstanding open problems are known to have solutions (see , ). Moreover, this weak cancellation condition holds widely (for instance, for all the known classes of regular rings - see , and also for the known classes of extremally rich $`C^{}`$-algebras - see ); it is therefore regarded as a condition that might hold for all exchange rings. Our main objective is to prove that if $`I`$ is a separative exchange ideal of a unital ring $`R`$, then the index is the only obstruction to lifting units modulo $`I`$, thus providing a common setting in which the results mentioned above can be handled. Hence we derive some consequences for both regular rings and $`C^{}`$-algebras with real rank zero (and their multiplier algebras). In order to develop our index theory, we benefit from results and techniques from , where a detailed analysis of elementary transformations on invertible matrices over unital exchange rings was carried out. Our present context is, however, different in that we deal with invertible matrices over unital rings, which are diagonal modulo an exchange ideal. It is remarkable that, in a different direction, it has been established by Brown and Pedersen (, Theorem 5.2) that the index is also the only obstruction to lifting invertibles modulo separative, extremally rich ideals of unital $`C^{}`$-algebras. We now fix some notations. As a general rule, $`R`$ will stand for a unital ring, whereas we shall use $`I`$ to denote a nonunital ring, generally sitting inside $`R`$ as a two-sided ideal. An elementary matrix is a matrix of the form $`\mathrm{𝟏}`$$`+re_{ij}`$, where $`\mathrm{𝟏}`$ is an identity matrix, $`e_{ij}`$ is one of the usual matrix units (with $`ij`$), and $`rR`$. We will denote by $`E_n(R)`$ the subgroup of $`GL_n(R)`$ generated by the elementary matrices. If $`x,yM_n(R)`$, then we use $`xy`$ to denote the matrix $`\left(\begin{array}{cc}x& 0\\ 0& y\end{array}\right)`$, and we will denote by $`1_n`$ the unit of $`M_n(R)`$. ## 1. Preliminary results Let $`M`$ be a right $`R`$-module. We say that $`M`$ satisfies the finite exchange property (see ) if for every right $`R`$-module $`A`$ and any decompositions $$A=M^{}N=A_1\mathrm{}A_n$$ with $`M^{}M`$, there exist submodules $`A_i^{}A_i`$ (which are in fact direct summands, by the modular law) such that $$A=M^{}A_1^{}\mathrm{}A_n^{}.$$ Following , we say that $`R`$ is an exchange ring provided that $`R_R`$ satisfies the finite exchange property. This notion is right-left symmetric (see , Corollary 2). In , Theorem 2.1, it is proved that $`M`$ has the finite exchange property if and only if $`\mathrm{End}(M)`$ is an exchange ring. Also, a useful ring-theoretic characterization was provided independently by Goodearl and Nicholson: ###### Lemma 1.1. (, Theorem 2.1 in ) A unital ring $`R`$ is an exchange ring if and only if for every element $`aR`$ there exists an idempotent $`eaR`$ such that $`1e(1a)R`$.∎ This characterization motivated the notion of an exchange ring for rings without unit (see ). Namely, a (possibly non-unital) ring $`I`$ is said to be an exchange ring if for each $`xI`$, there exist an idempotent $`eI`$ and elements $`r,sI`$ such that $`e=xr=x+sxs`$. As proved in Lemma 1.1 in , the ring $`I`$ is exchange if and only if, whenever $`xI`$ and $`R`$ is a unital ring containing $`I`$ as a two-sided ideal, then there exists an idempotent $`exI`$ such that $`1e(1x)R`$. Of course, the notions of unital exchange and non-unital exchange agree if the ring $`I`$ has a unit. Since we will usually have a non-unital exchange ring $`I`$ which is an ideal of a unital ring $`R`$, we will adopt the terminology in and say that, in this context, “$`I`$ is an exchange ideal of $`R`$”. The class of exchange rings is pleasantly large: it includes regular rings, $`\pi `$-regular rings, semiperfect rings (which are exactly the semilocal exchange rings), right self-injective rings (see Remark 2.9 (b) in ) and $`C^{}`$-algebras with real rank zero (by Theorem 7.2 in ). It is closed under natural ring constructions. For example, if $`I`$ is an exchange ideal of $`R`$ and $`eR`$ is an idempotent, then $`eIe`$ is an exchange ideal of $`eRe`$ (, Proposition 1.3 and also , Proposition 1.10). Also, if $`I`$ is an exchange ring, then $`M_n(I)`$ is an exchange ring for all $`n1`$ (by Theorem 1.4 in ). The behaviour of exchange rings under extensions is characterized by an idempotent-lifting condition (, Theorem 2.2): if $`I`$ is an ideal of a (possibly non-unital) ring $`L`$, then $`L`$ is exchange if and only if $`I`$ and $`L/I`$ are exchange and idempotents can be lifted modulo $`I`$. Let $`I`$ be an ideal of a unital ring $`R`$. We denote by $`FP(I,R)`$ the class of all finitely generated projective right $`R`$-modules $`P`$ such that $`P=PI`$, and we define $`V(I)`$ to be the set of all isomorphism classes of elements from $`FP(I,R)`$. Note that $`V(I)`$ becomes an abelian monoid under the operation $`[P]+[Q]=[PQ]`$. Even though $`V(I)`$ involves the unital ring $`R`$, it can be shown that it only depends on the ring structure of $`I`$ (see, for example, , or also ). It will be sometimes convenient to use an alternate description of $`V(I)`$ via idempotents (see ), so we identify $`V(I)`$ with the set of equivalence classes of idempotents in $`M_{\mathrm{}}(I)`$, the non-unital ring of $`\omega \times \omega `$ matrices with only finitely many nonzero entries from $`I`$. We will use $`[e]`$ to indicate the class of $`e`$ in $`V(I)`$. Viewing $`I`$ inside $`R`$, we can also identify $`V(I)`$ with $`\{[e]V(R)e\text{ an idempotent in }M_{\mathrm{}}(I)\}`$. Let $`M`$ be an (abelian) monoid. We can order $`M`$ by using the so-called algebraic ordering: for $`x,yM`$, write $`xy`$ provided that there exists $`zM`$ such that $`x+z=y`$. If $`S`$ is a submonoid of $`M`$ with the property that if $`xy`$ and $`yS`$, then $`xS`$, then $`S`$ is called an order-ideal of $`M`$. For example, if $`I`$ is a two-sided ideal of a ring $`R`$, then $`V(I)`$ is an order-ideal of $`V(R)`$. Notice that if $`e`$ and $`f`$ are idempotents in $`M_n(R)`$ for some $`n`$, then $`[e][f]`$ is equivalent to saying that $`eM_n(R)`$ is isomorphic to a direct summand of $`fM_n(R)`$. Let $`M`$ be a monoid, and let $`S`$ be an order-ideal of $`M`$. We say that $`M`$ has refinement with respect to $`S`$ if whenever $`x_1+x_2=y_1+y_2`$ in $`M`$ with at least one of $`x_1`$, $`x_2`$, $`y_1`$, $`y_2`$ in $`S`$, then there exist elements $`z_{ij}M`$ such that $`x_i=z_{i1}+z_{i2}`$ and $`y_i=z_{1i}+z_{2i}`$, for $`i=1,2`$. Observe that if $`S=M`$ then we are recovering the usual definition of a refinement monoid (see, for example ). Also, we see from the definition that in particular, $`S`$ is a refinement monoid. If $`I`$ is an exchange ring, it is known that $`V(I)`$ is a refinement monoid (see , Proposition 1.5, and also , Proposition 1.2). If now $`I`$ is an exchange ideal of a unital ring $`R`$, still some refinement persists in $`V(R)`$, as the following lemma shows: ###### Lemma 1.2. Let $`I`$ be an exchange ideal of a unital ring $`R`$. Then $`V(R)`$ has refinement with respect to $`V(I)`$. Proof: Let $`[A]+[B]=[C]+[D]`$, with $`[A]`$, $`[B]`$, $`[C]`$, $`[D]V(R)`$, and assume, for example, that $`[A]V(I)`$, that is, $`AFP(I,R)`$. Then $`\mathrm{End}(A)`$ is an exchange ring, so that $`A`$ has the finite exchange property, and hence we may use the proof of Proposition 1.2 in .∎ We say that an abelian monoid $`M`$ is separative provided that whenever $`a+a=a+b=b+b`$ in $`M`$, then $`a=b`$. Equivalently, $`M`$ is separative if the following weak cancellation condition holds: if $`a+c=b+c`$ and $`cna,nb`$ for some $`n`$, then $`a=b`$ (see Lemma 2.1 in ). Accordingly, we call a ring $`R`$ separative if $`V(R)`$ is a separative monoid (see ). The following lemma was stated in , Lemma 4.4, for full refinement monoids, and the proof used there can be used in our present context almost entirely. Since it will be an essential result later, we just indicate the major steps that lead to the conclusion. ###### Lemma 1.3. Let $`M`$ be a monoid and let $`S`$ be a separative order-ideal such that $`M`$ has refinement with respect to $`S`$. If $`a+e=b+e`$ for $`a,bM`$ and $`eS`$, and $`ena,nb`$ for some $`n`$, then $`a=b`$. Proof: Since $`M`$ has refinement with respect to $`S`$, and $`eS`$ with $`ena`$, we can decompose $`e=\underset{i=1}{\overset{n}{}}e_i`$, with $`e_ia`$ for all $`i`$. Hence we may assume that $`ea`$, and similarly $`eb`$. By refinement with respect to $`S`$ we get decompositions $`a=a_1+a_2`$, $`b=a_1+b_2`$, and $`e=b_2+c_2=a_2+c_2`$. Note that $`c_2S`$ and that $`c_2ea=a_1+a_2`$, hence we may use refinement with respect to $`S`$ again. It is not difficult to see that we can arrange the resulting decompositions and change notation in such a way that $`c_2a_2,b_2`$. Now separativity in $`S`$ entails $`a_2=b_2`$, so $`a=b`$.∎ Recall that an element $`x`$ in a ring $`R`$ is called von Neumann regular if there exists $`yR`$ such that $`x=xyx`$. If $`y`$ can be chosen to be a unit, then $`x`$ is called unit-regular. In this case $`x=(xy)y^1`$ is a product of an idempotent with a unit. ###### Proposition 1.4. Let $`I`$ be a separative exchange ideal of a unital ring $`R`$, and let $`dI`$ be such that $`dR=(1p)R`$ and $`Rd=R(1q)`$, for some idempotents $`1p,1qI`$. If $`RpR=RqR=R`$ then $`[p]=[q]`$ in $`V(R)`$. In particular, $`d`$ is unit-regular. Proof: Note that $`V(I)`$ is a separative monoid and that $`V(R)`$ has refinement with respect to $`V(I)`$ by Lemma 1.2. From the outset we evidently have that $`[1p]=[1q]`$ in $`V(I)V(R)`$. Also, since $`RpR=RqR=R`$, we get that $`[1p]n[p],n[q]`$ for some $`n`$. Now, in $`V(R)`$, we have that $`[p]+[1p]=[q]+[1p]`$. By Lemma 1.3, we get that $`[p]=[q]`$ in $`V(R)`$, so that $`pRqR`$. Now, note that $`R/dR=R/(1p)RpRqR=\mathrm{r}.\mathrm{ann}(d)`$, whence $`d`$ is unit-regular, by the proof of Theorem 4.1 in .∎ ## 2. Index theory for exchange rings ###### Lemma 2.1. (cf. , Lemma 2.1) Let $`I`$ be an exchange ideal of a unital ring $`R`$, and let $`e_1,e_2R`$ be idempotents such that $`e_1I`$. Then, there exists an idempotent $`ee_1R+e_2R`$ such that $`[e_i][e]`$ in $`V(R)`$, for all $`i`$. In particular, $`ReR=Re_1R+Re_2R`$. Proof: Considering that $`\mathrm{End}(e_1R)=e_1Re_1=e_1Ie_1`$ is a unital exchange ring, we have that $`e_1R`$ has the finite exchange property, and hence we get decompositions $`e_2R=AB`$ and $`(1e_2)R=A^{}B^{}`$ such that $`R=e_1RAA^{}`$. Then, choose $`eR`$ such that $`eR=e_1RA`$, and proceed as in the proof of Lemma 2.1 in .∎ In the next technical lemmas, we will be involved with performing several elementary row and column operations on an invertible $`2\times 2`$ matrix over a ring $`R`$. These operations will follow the lines of 2.3-2.7 in . However, our ring $`R`$ won’t be exchange, so we cannot apply the results in directly, and hence some different procedure is needed. Let $`I`$ be a two-sided ideal in a unital ring. We shall denote by $`\pi :RR/I`$ the natural quotient map. For any $`n>1`$, let $`E_n(I)`$ be the subgroup of $`E_n(R)`$ generated by the elementary matrices $`1_n+re_{ij}`$ for $`rI`$ and $`ij`$. Note that $`\pi (ϵ)=1_n`$ for all $`ϵE_n(I)`$. Observe also that multiplying a matrix $`\alpha M_n(R)`$ on the left or right by any matrices from $`E_n(I)`$ does not change $`\pi (\alpha )`$. ###### Lemma 2.2. Let $`I`$ be an exchange ideal of a unital ring $`R`$. Let $`\alpha =\left(\begin{array}{cc}a& b\\ c& d\end{array}\right)GL_2(R)`$ such that $`b,cI`$. * There exist $`\beta E_2(I)`$ and an idempotent $`1hI`$ such that $`\alpha \beta =\left(\begin{array}{cc}a^{}& b^{}\\ c^{}& d^{}\end{array}\right)`$, with $`c^{}Rc`$, $`c^{}R=(1h)R`$, $`d^{}R=hR`$ and $`RhR=R`$. * There exist $`\gamma E_2(I)`$ and an idempotent $`1kI`$ such that $`\gamma \alpha =\left(\begin{array}{cc}a^{\prime \prime }& b^{\prime \prime }\\ c^{\prime \prime }& d^{\prime \prime }\end{array}\right)`$, with $`b^{\prime \prime }bR`$, $`Rb^{\prime \prime }=R(1k)`$, $`Rd^{\prime \prime }=Rk`$ and $`RkR=R`$. Proof: $`(a)`$. First note that the row $`(c,d)`$ is right unimodular, so $`cR+dR=R`$. Hence there exist $`x,yR`$ such that $`cx+dy=1`$. Now, since $`cI`$ and $`I`$ is exchange, there exists an idempotent $`ecxRcR`$ such that $`1edR`$. Write $`e=cr`$, with $`re=r`$ and $`1e=ds`$, with $`s(1e)=s`$. As in , Lemma 2.4, we multiply $`\alpha `$ on the right by $`\alpha _1\alpha _2`$, where $`\alpha _iE_2(I)`$ are defined as follows: $$\alpha _1=\left(\begin{array}{cc}1& 0\\ sc& 1\end{array}\right),\alpha _2=\left(\begin{array}{cc}1& rd\\ 0& 1\end{array}\right).$$ We thus obtain as last row: $`(ec,(1e)d)`$. As in the proof of Lemma 2.4 in we get $`eccRc`$, $`(1e)ddRd`$, and $`R=ecR(1e)dR`$. Set $`w=ec+(1e)d`$, and note that $`ewr+(1e)ws=1`$, hence $`RwR=R`$. Similar to the proof of Corollary 2.5 in , we want to put $`w`$ in the $`(2,2)`$ position. By the exchange property again (applied to $`ewrI`$), we get an idempotent $`fewrRewR`$ such that $`1f(1ewr)R(1e)wR`$. Write $`f=eww_1`$, with $`w_1f=w_1`$, and $`1f=(1e)ww_2`$, with $`w_2(1f)=w_2`$, and set $`f_1=ww_1ewR`$, and $`f_2=ww_2(1e)wR`$. Note that $`f_i`$ are idempotents with $`f_1I`$, and that $`f_1RfR`$, whereas $`f_2R(1f)R`$. By Lemma 2.1, there is an idempotent $`gf_1R+f_2RwR`$ such that $`RgR=Rf_1R+Rf_2R`$. Since $`Rf_1R=RfR`$ and $`Rf_2R=R(1f)R`$, we see that $`RgR=R`$. Write $`g=ww^{}`$, for some $`w^{}R`$, and multiply $`\alpha \alpha _1\alpha _2`$ on the right by $`\beta _1\beta _2`$, where $`\beta _iE_2(I)`$ are defined as: $$\beta _1=\left(\begin{array}{cc}1& rc\\ 0& 1\end{array}\right),\beta _2=\left(\begin{array}{cc}1& 0\\ w^{}ec& 1\end{array}\right).$$ This gives as last row: $`((1g)ec,w)`$. At this point we start the first part of the procedure again with the current last row, so that after right multiplication by two matrices $`\gamma _iE_2(I)`$ for $`i=1,2`$, we get as last row $`(c^{},d^{})`$, with $`c^{}(1g)ecR(1g)ec(1g)Rc`$, and $`R=c^{}Rd^{}R`$. According to the direct sum decomposition, we see that there exists an idempotent $`1hI`$ such that $`c^{}R=(1h)R`$ and $`d^{}R=hR`$. Since $`g(1h)R=gc^{}R=0`$ and $`RgR=R`$, we conclude that $`RhR=R`$. Finally, set $`\beta =\alpha _1\alpha _2\beta _1\beta _2\gamma _1\gamma _2`$. (b). Since $`bI`$ and $`I`$ is exchange, we can perform the transpose version of the process carried out in $`(a)`$ to the last column. Hence we obtain matrices $`\alpha _i^{}`$, $`\beta _i^{}`$ and $`\gamma _i^{}`$ in $`E_2(I)`$, for $`i=1,2`$, such that after left multiplication by $`\gamma :=\gamma _2^{}\gamma _1^{}\beta _2^{}\beta _1^{}\alpha _2^{}\alpha _1^{}`$ we get a matrix $`\left(\begin{array}{cc}a^{\prime \prime }& b^{\prime \prime }\\ c^{\prime \prime }& d^{\prime \prime }\end{array}\right)`$ and an idempotent $`1kI`$ satisfying the desired properties.∎ ###### Lemma 2.3. Let $`I`$ be a separative exchange ideal of a unital ring $`R`$, and let $`\alpha =\left(\begin{array}{cc}a& b\\ c& d\end{array}\right)GL_2(R)`$ such that $`b,cI`$ and $`d1I`$. Then there exist units $`a^{},uGL_1(R)`$, and matrices $`\beta `$, $`\gamma `$ and $`ϵ`$ in $`E_2(R)`$ such that $$\gamma \alpha \beta (1u^1)ϵ=a^{}1,$$ and $`\pi (a^{})=\pi (au^1)`$. In particular, $`[\alpha ]=[a^{}u]`$ in $`K_1(R)`$. Proof: We remark again that $`\pi (\alpha )`$ remains unchanged after right or left multiplication by matrices from $`E_2(I)`$. Thus we apply Lemma 2.2 $`(a)`$, and without loss of generality there is an idempotent $`1hI`$ such that $`cR=(1h)R`$, $`dR=hR`$ and $`RhR=R`$. Now we proceed as in the proof of Lemma 2.7 in , so that we move $`c`$ to the $`(1,2)`$ position. This is achieved after right and left multiplication by the signed permutation matrix $`\sigma =\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)`$, which is a product of three elementary matrices. Hence, we get: $$\alpha ^{}:=\sigma \alpha \sigma =\left(\begin{array}{cc}d& c\\ b& a\end{array}\right).$$ By Lemma 2.2 $`(b)`$, there exist $`\gamma ^{}E_2(I)`$ and an idempotent $`1qI`$ such that $`\gamma ^{}\alpha ^{}=\left(\begin{array}{cc}a^{}& b^{}\\ c^{}& d^{}\end{array}\right)`$, with $`b^{}cR=(1h)R`$, $`Rb^{}=R(1q)`$ and $`RqR=R`$. Note that $`\pi (a^{})=1`$. Since $`b^{}`$ is regular, there exists an idempotent $`1pI`$ such that $`(1p)R=b^{}R`$. Using that $`h(1p)R=hb^{}RhcR=0`$ and $`RhR=R`$, we conclude that $`RpR=R`$. Hence, by Proposition 1.4 (and since $`I`$ is separative, by hypothesis), $`b^{}`$ is unit-regular. Write $`b^{}=fu`$, where $`f^2=fI`$ and $`uGL_1(R)`$. Now move the element $`b^{}`$ to the $`(2,2)`$ position, multiplying $`\gamma ^{}\alpha ^{}`$ on the left by $`\sigma ^1`$. Then we multiply $`\sigma ^1\gamma ^{}\alpha ^{}`$ on the right by $`\lambda =\left(\begin{array}{cc}1& 0\\ 0& u^1\end{array}\right)`$. The matrix we obtain is $`\alpha ^{\prime \prime }:=\left(\begin{array}{cc}r& s\\ t& f\end{array}\right)`$, where $`rI`$, and $`\pi (t)=1`$. We want to use now Lemma 2.3 in , in order to get $`1`$ in the $`(2,2)`$ position and zeros elsewhere in the last row and column. First, multiply $`\alpha ^{\prime \prime }`$ on the right by $`ϵ_1=\left(\begin{array}{cc}1& 0\\ ft& 1\end{array}\right)E_2(I)`$, so we get as last row $`((1f)t,f)`$. Since this row is right unimodular, we have that $`(1f)tR+fR=R`$, and hence $`(1f)tR=(1f)R`$. Choose $`vR(1f)`$ such that $`(1f)tv=(1f)`$. Since $`\pi (t)=1`$, we have that $`\pi (v)=1`$. After right multiplication by $`ϵ_2=\left(\begin{array}{cc}1& v\\ 0& 1\end{array}\right)`$, our second row becomes $`((1f)t,1)`$. Denote by $`z^{}`$ the element in the $`(1,2)`$ position of the resulting matrix. Finally, set $`\mu _1=\left(\begin{array}{cc}1& z^{}\\ 0& 1\end{array}\right)`$ and $`\mu _2=\left(\begin{array}{cc}1& 0\\ (1f)t& 1\end{array}\right)`$, the last routine matrices. Observe that $$\pi (\mu _1)=\left(\begin{array}{cc}1& \pi (z^{})\\ 0& 1\end{array}\right)\text{ and that }\pi (\mu _2)=\left(\begin{array}{cc}1& 0\\ 1& 1\end{array}\right).$$ After multiplying $`\alpha ^{\prime \prime }ϵ_1ϵ_2`$ on the left by $`\mu _1`$ and on the right by $`\mu _2`$, we get a matrix of the form $`a^{}1`$, where $`a^{}R`$. We need to compute $`\pi (a^{})`$. By all the calculations performed so far, we have that: $$\pi (a^{})1=\pi (\mu _1\sigma ^1\gamma ^{}\sigma \alpha \sigma \lambda ϵ_1ϵ_2\mu _2)=$$ $$\left(\begin{array}{cc}1& \pi (z^{})\\ 0& 1\end{array}\right)\left(\begin{array}{cc}\pi (a)& 0\\ 0& 1\end{array}\right)\left(\begin{array}{cc}0& 1\\ 1& 0\end{array}\right)\left(\begin{array}{cc}1& 0\\ 0& \pi (u^1)\end{array}\right)\left(\begin{array}{cc}1& 1\\ 0& 1\end{array}\right)\left(\begin{array}{cc}1& 0\\ 1& 1\end{array}\right).$$ After computing the right-hand side of the above equality, we obtain: $$\pi (a^{})1=\left(\begin{array}{cc}\pi (au^1)& \pi (au^1)\pi (z^{})\\ 0& 1\end{array}\right),$$ whence we get that $`\pi (a^{})=\pi (au^1)`$, as desired.∎ Let $`R`$ be a ring, and let $`I`$ be a two-sided ideal. We define the Fredholm elements relative to $`I`$ as the set $`F(I,R)=\pi ^1(GL_1(R/I))`$. Note that $`F(I,R)`$ is a multiplicative subsemigroup of $`R`$, such that $`GL_1(R)+IF(I,R)`$. Observe that if $`R`$ has stable rank one, then $`GL_1(R)+I=F(I,R)`$. Denote by $`\delta :K_1(R/I)K_0(I)`$ the connecting map in algebraic $`K`$-Theory, and recall that there is an exact sequence $$K_1(R)\stackrel{\pi _1}{}K_1(R/I)\stackrel{\delta }{}K_0(I)K_0(R)K_0(R/I)$$ (, Theorem 2.5.4). As usual, cf. , 6.1, we define the index map as the semigroup homomorphism $$\mathrm{index}:F(I,R)K_0(I),$$ given by the rule $`\mathrm{index}(x)=\delta ([\pi (x)])`$. We are now in position to prove our main result: ###### Theorem 2.4. Let $`I`$ be a separative exchange ideal of a unital ring $`R`$. Let $`xR`$ be a Fredholm element relative to $`I`$. Then there exists $`yGL_1(R)`$ such that $`xyI`$ if and only if $`\mathrm{index}(x)=0`$. In this case, for any $`\alpha K_1(R)`$ that is mapped to $`[\pi (x)]`$, we may find $`yGL_1(R)`$ such that $`[y]=\alpha `$ and $`xyI`$. Proof: Assume that $`\pi (x)`$ can be lifted to a unit of $`R`$. Then there exists $`yGL_1(R)`$ such that $`x=y+b`$, with $`bI`$. Hence $$\mathrm{index}(x)=\delta ([\pi (x)])=\delta ([\pi (y)])=\delta \pi _1([y])=0,$$ by exactness. Conversely, suppose that $`\mathrm{index}(x)=0`$. Again by exactness this means that there exists $`k`$ and $`y_1GL_k(R)`$ such that $`[\pi (y_1)]=[\pi (x)]`$. Hence, if $`m=2^n`$ is large enough, there is $`\pi (z)E_m(R/I)`$ such that $`\pi (x)1_{m1}=\pi (z)(\pi (y_1)1_{mk})`$. We may clearly assume that in fact $`zE_m(R)`$. Set $`w_1:=z(y_11_{mk})`$, and note that $`[w_1]=[y_1]`$ in $`K_1(R)`$. Denote by $`(a_{ij})`$ the entries of $`w_1`$, and observe that $`a_{ij}I`$ whenever $`ij`$, that $`a_{ii}1I`$ for all $`i1`$, and that $`a_{11}xI`$. We apply Lemma 2.3, replacing $`I`$ and $`R`$ by $`M_{m/2}(I)`$ and $`M_{m/2}(R)`$. (Notice that $`M_{m/2}(I)`$ is a separative exchange ideal of $`M_{m/2}(R)`$.) Thus we obtain matrices $`u,w_2^{}GL_{m/2}(R)`$ such that $`\pi (w_2^{}u)=(\pi (x)1_{m/21})`$, and $`[w_1]=[w_2^{}u]`$ in $`K_1(R)`$. Let $`w_2:=w_2^{}u`$. A recursive procedure shows that we get $`w_nGL_1(R)`$ such that $`\pi (w_n)=\pi (x)`$, and $`[w_n]=[y_1]`$ in $`K_1(R)`$.∎ ###### Remark 2.5. Notice that, according to Theorem 3.5 in , every ring has a largest exchange ideal with respect to the inclusion (which might be the zero ideal). ###### Corollary 2.6. Let $`R`$ be a (unital) separative exchange ring such that $`K_0(R)=0`$, and let $`I`$ be an ideal of $`R`$. Then the units of $`R/I`$ lift to those of $`R`$ if and only if $`K_0(I)=0`$. Proof: Clearly, if $`K_0(I)=0`$, then the connecting map $`\delta `$ vanishes and hence Theorem 2.4 applies. Conversely, suppose that the units of $`R/I`$ lift to units of $`R`$. By , Theorem 2.8, the natural map $`GL_1(R/I)K_1(R/I)`$ is surjective. It follows then that the map $`\pi _1:K_1(R)K_1(R/I)`$ in $`K`$-Theory is surjective. Therefore we get again that $`\delta =0`$, and since $`K_0(R)=0`$, we conclude by exactness that $`K_0(I)=0`$.∎ The previous corollary applies to the case where $`R`$ is a purely infinite, right self-injective ring and recovers some results of Menal and Moncasi (). If $`R`$ is a right self-injective ring, then by Theorem 1.22 in , the quotient $`R/J(R)`$ is regular and right self-injective, where $`J(R)`$ is the Jacobson radical of $`R`$. We say that a right self-injective ring is purely infinite if $`R/J(R)`$ is a purely infinite regular ring. (According to , a regular ring $`R`$ is purely infinite if $`RRR`$.) These rings are known to be exchange and separative. Further, $`K_0(R)=0`$, by the proof of Corollary 3.6 in . ## 3. $`C^{}`$-algebras with real rank zero Let $`A`$ be a $`C^{}`$-algebra. In order to distinguish between the algebraic and the topological $`K_1`$-groups of $`A`$, and according to more common usage, we will denote by $`K_1^{\mathrm{alg}}(A)`$ the algebraic $`K_1`$-group of $`A`$, and we shall use $`K_1(A)`$ to denote the topological $`K_1`$-group (see , Definition 8.1.1, , Definition 7.1.1). It is known that there is a natural surjective homomorphism $`\gamma :K_1^{\mathrm{alg}}(A)K_1(A)`$ (see, for example, ). Since idempotents in $`M_n(A)`$ are equivalent to projections (e.g., ), we may identify $`V(A)`$ with the abelian monoid of Murray-von Neumann equivalence classes of projections arising from $`M_{\mathrm{}}(A)`$. Recall that a (unital) $`C^{}`$-algebra $`A`$ has real rank zero provided that every self-adjoint element can be approximated arbitrarily well by self-adjoint, invertible elements. Other characterizations, including the original definition, may be found in . If $`A`$ is non-unital, then $`A`$ has real rank zero if and only if the minimal unitization $`\stackrel{~}{A}`$ of $`A`$ (see ) has real rank zero. Since the $`C^{}`$-algebras that are exchange rings are precisely those having real rank zero, we see that if $`A`$ is a $`C^{}`$-algebra with real rank zero, then $`V(A)`$ is a refinement monoid (see also , Theorem 5.3). Brown and Pedersen have introduced in the concept of weak cancellation for $`C^{}`$-algebras, meaning that if $`p`$ and $`q`$ are projections in $`A`$ that generate the same closed ideal $`I`$ of $`A`$, and $`[p]=[q]`$ in $`K_0(I)`$, then they are (Murray-von Neumann) equivalent in $`A`$. If this property holds for $`M_n(A)`$, for all $`n`$, then $`A`$ has stable weak cancellation. Notice that $`A`$ has stable weak cancellation if and only if $`A`$ is separative. In fact, if $`A`$ has real rank zero, then $`A`$ has weak cancellation if and only if $`A`$ is separative (that is, the property of weak cancellation is stable). This follows using Proposition 2.8 in and the fact that $`V(A)`$ is a refinement monoid. The property of (stable) weak cancellation is shown to hold widely within the class of extremally rich $`C^{}`$-algebras (see ), including those that have real rank zero (see , Theorem 2.11). As for the case of exchange rings, there are no examples known of extremally rich $`C^{}`$-algebras without weak cancellation (, Remark 2.12). Let $`A`$ be a $`C^{}`$-algebra, and let $`I`$ be a closed, two-sided ideal of $`A`$. Denote by $`:K_1(A/I)K_0(I)`$ the connecting map in topological $`K`$-Theory (see, e.g., , Definition 8.3.1, , Definition 8.1.1). We then define the index of a Fredholm element $`x`$ (relative to $`I`$) as $`\mathrm{index}(x)=([\pi (x)])`$, where $`\pi :AA/I`$ is the natural projection map. Now, since we have that $`\gamma =\delta `$, where $`\delta :K_1^{\mathrm{alg}}(A/I)K_0(I)`$ is the algebraic connecting map, we see that the two possible definitions of algebraic and topological indices for Fredholm elements coincide. From the observations made, it is clear that we can apply the result in the previous section to get the following theorem (which has been independently obtained by L.G. Brown \[unpublished\]). We remark that for extremally rich ideals with weak cancellation the same conclusion holds, as shown in , Theorem 5.2. ###### Theorem 3.1. Let $`A`$ be a $`C^{}`$-algebra and let $`I`$ be a closed ideal of $`A`$ with real rank zero and weak cancellation. Let $`xA`$ be a Fredholm element relative to $`I`$. Then there exists $`yGL_1(A)`$ such that $`xyI`$ if and only if $`\mathrm{index}(x)=0`$. In this case, for any $`\alpha K_1(A)`$ that is mapped to $`[\pi (x)]`$, we may find $`yGL_1(A)`$ such that $`[y]=\alpha `$ and $`xyI`$. Proof: As in the proof of Theorem 2.4, if there exists $`yGL_1(A)`$ such that $`xyI`$, then $`\mathrm{index}(x)=0`$. Conversely, if $`\mathrm{index}(x)=([\pi (x)])=0`$, then since $`\gamma ([\pi (x)])=[\pi (x)]`$ we have in fact that $`\delta ([\pi (x)])=0`$, and so Theorem 2.4 applies.∎ ###### Remark 3.2. As for the case of exchange rings (see Remark 2.5), any $`C^{}`$-algebra has a largest closed ideal of real rank zero, a description of which is given in , Theorem 2.3. We now give some applications to the multiplier algebras $`M(A)`$ of $`C^{}`$-algebras $`A`$ with real rank zero. Multiplier algebras of $`C^{}`$-algebras are relevant objects (since they can be used, for instance, to parametrize extensions) that have been intensively studied in the last years (to cite a few examples, among many others, see , , , , , , ). For the basic facts concerning multipliers see, for example, Chapter 2 in . The proof of the following corollary is derived entirely as the proof of Corollary 5.8 in , using Theorem 3.1 instead of , Theorem 5.2. For any unital $`C^{}`$-algebra $`A`$, we use $`U(A)`$ to denote the unitary group of $`A`$, whereas $`U_0(A)`$ stands for the connected component of the identity in $`U(A)`$. ###### Corollary 3.3. Let $`A`$ be a $`\sigma `$-unital and stable $`C^{}`$-algebra. Suppose that $`A`$ has real rank zero and weak cancellation. Then there is a short exact sequence of groups $$0U_0(M(A)/A)U(M(A)/A)K_0(A)0.\mathit{}$$ ###### Corollary 3.4. Let $`A`$ be a $`\sigma `$-unital and stable $`C^{}`$-algebra, with real rank zero and weak cancellation. Then the following are equivalent: * The units of $`M(A)/A`$ lift to units of $`M(A)`$; * $`K_0(A)=0`$; * $`U(M(A)/A)`$ is connected. Proof: Since $`A`$ is $`\sigma `$-unital and stable, we have that $`U(M(A))`$ is connected (see , or Theorem 16.8 in ). Granted this, and using also Corollary 4.3.3 in , it is obvious that $`(a)(c)`$. Now, $`(b)(c)`$ according to Corollary 3.3.∎ ###### Corollary 3.5. Let $`A`$ be a $`\sigma `$-unital (non-unital) purely infinite simple $`C^{}`$-algebra. Then the units of $`M(A)/A`$ lift to units of $`M(A)`$ if and only if $`K_0(A)=0`$. Proof: By , Theorem 1.2, $`A`$ is stable and has real rank zero, and by Theorem 1.4 and Proposition 1.5 in , $`A`$ has weak cancellation. Thus the result follows from Corollary 3.4.∎ We close by remarking the fact that if $`A`$ is simple, $`\sigma `$-unital (non-unital), with real rank zero and weak cancellation, then $`U(M(A))`$ is connected (and hence $`K_1(M(A))=0`$). Indeed, if all projections in $`A`$ are infinite, then $`A`$ is purely infinite simple, hence stable (, Theorem 1.2 (i)), and thus $`U(M(A))`$ is connected. On the other hand, if there is a nonzero finite projection $`pA`$, then $`pAp`$ has stable rank one, by Theorem 7.6 in , whence also $`A`$ has stable rank one, and then $`U(M(A))`$ is connected, by Lemma 3.3 in . Hence, for these algebras, the units of $`M(A)/A`$ can be lifted to units of $`M(A)`$ if and only if $`U(M(A)/A)`$ is connected. ## Acknowledgements The author is grateful to Gert Pedersen for inviting him to the Mathematics Institute of the University of Copenhagen, where this work was carried out. It is also a pleasure to thank Ken Goodearl for many helpful comments on an earlier draft of this paper.
no-problem/9906/physics9906058.html
ar5iv
text
# Why the Parity Violation ## 1 Hints 1) Let us consider the free lepton Lagrangian : $$=0.5i\left(\overline{\psi }\gamma ^\mu \left(_\mu \psi \right)\left(_\mu \overline{\psi }\right)\gamma ^\mu \psi \right)m\overline{\psi }\psi \text{.}$$ Hence: $$=0.5i\left(\psi ^{}\beta ^\mu \left(_\mu \psi \right)\left(_\mu \psi ^{}\right)\beta ^\mu \psi \right)m\psi ^{}\gamma ^0\psi \text{.}$$ This Lagrangian contains four matrices from the Clifford pentad $$\left\{\gamma _0\text{}\beta _1\text{}\beta _2\text{}\beta _3\text{}\beta _4\right\}\text{,}$$ but one does not contain $`\beta _4`$. 2) Let us consider the lepton current: $$j_\mu =\psi ^{}\beta ^\mu \psi \text{.}$$ for $`0\mu 3`$. Let us denote: $$J_\gamma =\psi ^{}\gamma ^0\psi \text{ and }J_4=\psi ^{}\beta ^4\psi \text{.}$$ In this case if $$\rho =j_0$$ then the average velocity vector is: $$\rho v_x=j_1\text{}\rho v_y=j_2\text{}\rho v_z=j_3\text{.}$$ Let us denote: $$\rho V_\gamma =J_\gamma \text{ and }\rho V_4=J_4\text{.}$$ In this case: $$v_x^2+v_y^2+v_z^2+V_\gamma ^2+V_4^2=1\text{.}$$ Hence of only all five elements of the Clifford pentad lends the entire kit of the velocity components. 3) In the Standard Model we have got the following entities: the right electron field vector $`e_R`$, the left electron field vector $`e_L`$, the electron field vector $`e`$ ($`e=\left[\begin{array}{c}e_L\\ e_R\end{array}\right]`$), the left neutrino fields vector $`\nu _L`$. the zero right neutrino fields vector $`\nu _R`$. the unitary $`2\times 2`$ $`SU(2)`$ matrix $`U`$ of the isospin transformation: $$U=\left[\begin{array}{cc}\mathrm{cos}\left(\epsilon \right)+in_3\mathrm{sin}\left(\epsilon \right)& \left(in_1+n_2\right)\mathrm{sin}\left(\epsilon \right)\\ \left(in_1n_2\right)\mathrm{sin}\left(\epsilon \right)& \mathrm{cos}\left(\epsilon \right)in_3\mathrm{sin}\left(\epsilon \right)\end{array}\right]\text{,}$$ $`\epsilon `$, $`n_1`$, $`n_2`$, $`n_3`$ are real and: $$n_1^2+n_2^2+n_3^2=1\text{.}$$ This matrix acts on the vectors of the kind:$`\left[\begin{array}{c}\nu _L\\ e_L\end{array}\right]`$. Therefore, in this theory (the (j,0)+(j,0) representation space: , , ): if $$U=\left[\begin{array}{cc}u_{1,1}& u_{1,2}\\ u_{2,1}& u_{2,2}\end{array}\right]$$ then the matrix $$\underset{¯}{U}=\left[\begin{array}{cccc}u_{1,1}1_2& 0_2& u_{1,2}1_2& 0_2\\ 0_2& 1_2& 0_2& 0_2\\ u_{2,1}1_2& 0_2& u_{2,2}1_2& 0_2\\ 0_2& 0_2& 0_2& 1_2\end{array}\right]$$ (1) operates on the vector $$\left[\begin{array}{c}\nu _L\\ \nu _R.\\ e_L\\ e_R\end{array}\right]$$ Because $`e_R`$, $`e_L`$, $`\nu _L`$, $`\nu _R`$ are the two-component vectors then $$\left[\begin{array}{c}\nu _L\\ \nu _R\\ e_L\\ e_R\end{array}\right]\text{ is }\left[\begin{array}{c}\nu _{L1}\\ \nu _{L2}\\ \nu _{R1}\\ \nu _{R1}\\ e_{L1}\\ e_{L2}\\ e_{R1}\\ e_{R1}\end{array}\right]$$ $`\underset{¯}{U}`$ has got eight orthogonal normalized eigenvectors $`s_1`$, $`s_2`$, $`s_3`$, $`s_4`$, $`s_5`$, $`s_6`$, $`s_7`$, $`s_8`$: $$s_1=\left[\begin{array}{c}0\\ 0\\ 1\\ 0\\ 0\\ 0\\ 0\\ 0\end{array}\right]\text{}s_2=\left[\begin{array}{c}0\\ 0\\ 0\\ 1\\ 0\\ 0\\ 0\\ 0\end{array}\right]\text{}s_3=\left[\begin{array}{c}a\\ 0\\ 0\\ 0\\ b+ic\\ 0\\ 0\\ 0\end{array}\right]\text{}s_4=\left[\begin{array}{c}0\\ a\\ 0\\ 0\\ 0\\ b+ic\\ 0\\ 0\end{array}\right]\text{,}$$ $$s_5=\left[\begin{array}{c}0\\ 0\\ 0\\ 0\\ 0\\ 0\\ 1\\ 0\end{array}\right]\text{}s_6=\left[\begin{array}{c}0\\ 0\\ 0\\ 0\\ 0\\ 0\\ 0\\ 1\end{array}\right]\text{}s_7=\left[\begin{array}{c}b+ic\\ 0\\ 0\\ 0\\ a\\ 0\\ 0\\ 0\end{array}\right]\text{}s_8=\left[\begin{array}{c}0\\ b+ic\\ 0\\ 0\\ 0\\ a\\ 0\\ 0\end{array}\right]$$ ($`a`$, $`b`$, $`c`$ are a real numbers) with the corresponding eigenvalues: $`1`$, $`1`$, $`\mathrm{exp}\left(i\lambda \right)`$, $`\mathrm{exp}\left(i\lambda \right)`$, $`1`$, $`1`$, $`\mathrm{exp}\left(i\lambda \right)`$, $`\mathrm{exp}\left(i\lambda \right)`$. Let: $`K`$ be the $`8\times 8`$ complex matrix, constructed by $`s_1`$, $`s_2`$, $`s_3`$, $`s_4`$, $`s_5`$, $`s_6`$, $`s_7`$, $`s_8`$ as the following: $$K=\left[\begin{array}{cccccccc}s_1& s_2& s_3& s_4& s_5& s_6& s_7& s_8\end{array}\right]\text{.}$$ Let for all $`k`$ ($`1k8`$): $$h_k=\underset{¯}{\gamma _0}s_k$$ and let: $$M=\left[\begin{array}{cccccccc}h_1& h_2& h_3& h_4& h_5& h_6& h_7& h_8\end{array}\right]\text{.}$$ Let: $$P_3=\left[\begin{array}{cccc}0_2& 0_2& 0_2& 0_2\\ 0_2& p_u& 0_2& 0_2\\ 0_2& 0_2& 0_2& 0_2\\ 0_2& 0_2& 0_2& 0_2\end{array}\right]\text{,}P_4=\left[\begin{array}{cccc}0_2& 0_2& 0_2& 0_2\\ 0_2& p_d& 0_2& 0_2\\ 0_2& 0_2& 0_2& 0_2\\ 0_2& 0_2& 0_2& 0_2\end{array}\right]\text{,}$$ $$P_7=\left[\begin{array}{cccc}0_2& 0_2& 0_2& 0_2\\ 0_2& 0_2& 0_2& 0_2\\ 0_2& 0_2& 0_2& 0_2\\ 0_2& 0_2& 0_2& p_u\end{array}\right]\text{}P_8=\left[\begin{array}{cccc}0_2& 0_2& 0_2& 0_2\\ 0_2& 0_2& 0_2& 0_2\\ 0_2& 0_2& 0_2& 0_2\\ 0_2& 0_2& 0_2& p_d\end{array}\right]\text{.}$$ In this case the projection matrices are: $$\begin{array}{c}Y_1=MP_3M^{}\text{,}\\ Y_2=MP_4M^{}\text{,}\\ Y_3=KP_3K^{}\text{,}\\ Y_4=KP_4K^{}\text{,}\\ Y_5=MP_7M^{}\text{,}\\ Y_6=MP_8M^{}\text{,}\\ Y_7=KP_7K^{}\text{,}\\ Y_8=KP_8K^{}\text{.}\end{array}$$ The vectors: $$\underset{¯}{e}=\left[\begin{array}{c}0\\ 0\\ 0\\ 0\\ e_{L1}\\ e_{L2}\\ e_{R1}\\ e_{R2}\end{array}\right]\text{}\underset{¯}{e_R}=\left[\begin{array}{c}0\\ 0\\ 0\\ 0\\ 0\\ 0\\ e_{R1}\\ e_{R2}\end{array}\right]\text{}\underset{¯}{e_L}=\left[\begin{array}{c}0\\ 0\\ 0\\ 0\\ e_{L1}\\ e_{L2}\\ 0\\ 0\end{array}\right]\text{.}$$ correspond to the vectors $`e`$, $`e_R`$ and $`e_L`$ resp. Let: $$\begin{array}{c}X_a=Y_1+Y_2+Y_3+Y_4\text{,}\\ X_b=Y_5+Y_6+Y_7+Y_8\text{,}\\ \underset{¯}{e_a}=X_a\underset{¯}{e}\text{,}\\ \underset{¯}{e_b}=X_b\underset{¯}{e}\text{.}\end{array}$$ In this case: $$\begin{array}{c}X_a+X_b=1_8\text{,}\\ X_aX_b=0_8\text{,}\\ X_aX_a=X_a\text{,}\\ X_bX_b=X_b\text{,}\\ X_a^{}=X_a\text{,}\\ X_b^{}=X_b\text{.}\end{array}$$ Let: $$\begin{array}{c}\rho _a=\underset{¯}{e_a}^{}\underset{¯}{e_a}\text{}\rho _b=\underset{¯}{e_b}^{}\underset{¯}{e_b}\text{,}\\ J_{\gamma ,a}=\underset{¯}{e_a}^{}\underset{¯}{\gamma _0}\underset{¯}{e_a}\text{}J_{\gamma ,b}=\underset{¯}{e_b}^{}\underset{¯}{\gamma _0}\underset{¯}{e_b}\text{,}\\ J_{4,a}=\underset{¯}{e_a}^{}\underset{¯}{\beta _4}\underset{¯}{e_a}\text{}J_{4,b}=\underset{¯}{e_b}^{}\underset{¯}{\beta _4}\underset{¯}{e_b}\text{,}\\ J_{\gamma ,a}=\rho _aV_{\gamma ,a}\text{}J_{\gamma ,b}=\rho _bV_{\gamma ,b}\text{,}\\ J_{4,a}=\rho _aV_{4,a}\text{}J_{4,b}=\rho _bV_{4,b}\text{.}\end{array}$$ Let: $$\begin{array}{c}\underset{¯}{e_a}^{}=U\underset{¯}{e_a}\text{}\underset{¯}{e_b}^{}=U\underset{¯}{e_b}\\ \rho _a^{}=\underset{¯}{e_a}^{}\underset{¯}{e_a}^{}\text{}\rho _b^{}=\underset{¯}{e_b}^{}\underset{¯}{e_b}^{}\text{,}\\ J_{\gamma ,a}^{}=\underset{¯}{e_a}^{}\underset{¯}{\gamma _0}\underset{¯}{e_a}^{}\text{}J_{\gamma ,b}^{}=\underset{¯}{e_b}^{}\underset{¯}{\gamma _0}\underset{¯}{e_b}^{}\text{,}\\ J_{4,a}^{}=\underset{¯}{e_a}^{}\underset{¯}{\beta _4}\underset{¯}{e_a}^{}\text{}J_{4,b}^{}=\underset{¯}{e_b}^{}\underset{¯}{\beta _4}\underset{¯}{e_b}^{}\text{,}\\ J_{\gamma ,a}^{}=\rho _a^{}V_{\gamma ,a}^{}\text{}J_{\gamma ,b}^{}=\rho _b^{}V_{\gamma ,b}^{}\text{,}\\ J_{4,a}^{}=\rho _a^{}V_{4,a}^{}\text{}J_{4,b}^{}=\rho _b^{}V_{4,b}^{}\text{.}\end{array}$$ In this case: $$V_{\gamma ,a}=V_{\gamma ,b}\text{}V_{4,a}=V_{4,b}$$ but: $$\begin{array}{c}V_{\gamma ,a}^{}=V_{\gamma ,a}\mathrm{cos}\left(\lambda \right)V_{4,a}\mathrm{sin}\left(\lambda \right)\text{,}\\ V_{4,a}^{}=V_{4,a}\mathrm{cos}\left(\lambda \right)+V_{\gamma ,a}\mathrm{sin}\left(\lambda \right)\text{,}\\ V_{\gamma ,b}^{}=V_{\gamma ,b}\mathrm{cos}\left(\lambda \right)+V_{4,b}\mathrm{sin}\left(\lambda \right)\text{,}\\ V_{4,b}^{}=V_{4,b}\mathrm{cos}\left(\lambda \right)V_{\gamma ,b}\mathrm{sin}\left(\lambda \right)\text{.}\end{array}$$ Hence, every isospin transformation $`U`$ divides a electron on two components which scatter on the angle $`2\lambda `$ in the space of ($`J_\gamma `$, $`J_4`$). Hence $`\beta ^4`$ must be inserted into Lagrangian. ## 2 Sufficient Conditions Let $`\underset{¯}{\psi }`$ be any field of the following type: $$\underset{¯}{\psi }=\left[\begin{array}{c}0\\ 0\\ 0\\ 0\\ \psi _{L1}\\ \psi _{L2}\\ \psi _{R1}\\ \psi _{R2}\end{array}\right]\text{.}$$ The value of the form $$\begin{array}{c}\left(\left(\underset{¯}{\psi }^{}\underset{¯}{\gamma ^0}X_a\underset{¯}{\psi }\right)^2+\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^4}X_a\underset{¯}{\psi }\right)^2\right)^{0.5}+\\ +\left(\left(\underset{¯}{\psi }^{}\gamma ^0X_b\underset{¯}{\psi }\right)^2+\left(\underset{¯}{\psi }^{}\beta ^4X_b\underset{¯}{\psi }\right)^2\right)^{0.5}\end{array}$$ (2) does not depend from the choice of the $`SU(2)`$ matrix $`U`$ and the Lagrangian: $$\begin{array}{c}_\psi =0.5i\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^\mu }\left(_\mu \underset{¯}{\psi }\right)\left(_\mu \underset{¯}{\psi }\right)^{}\underset{¯}{\beta ^\mu }\underset{¯}{\psi }\right)\\ m_\psi \left(\begin{array}{c}\left(\left(\underset{¯}{\psi }^{}\underset{¯}{\gamma ^0}X_a\underset{¯}{\psi }\right)^2+\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^4}X_a\underset{¯}{\psi }\right)^2\right)^{0.5}+\\ +\left(\left(\underset{¯}{\psi }^{}\underset{¯}{\gamma ^0}X_b\underset{¯}{\psi }\right)^2+\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^4}X_b\underset{¯}{\psi }\right)^2\right)^{0.5}\end{array}\right)\end{array}$$ is invariant for this $`SU(2)`$ transformation. Let us denote: $`{\displaystyle \frac{\left(\underset{¯}{\psi }^{}\underset{¯}{\gamma ^0}X_a\underset{¯}{\psi }\right)}{\sqrt{\left(\underset{¯}{\psi }^{}\underset{¯}{\gamma ^0}X_a\underset{¯}{\psi }\right)^2+\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^4}X_a\underset{¯}{\psi }\right)^2}}}`$ $`=`$ $`\mathrm{cos}\left(\alpha _a\right)\text{,}`$ $`{\displaystyle \frac{\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^4}X_a\underset{¯}{\psi }\right)}{\sqrt{\left(\underset{¯}{\psi }^{}\underset{¯}{\gamma ^0}X_a\underset{¯}{\psi }\right)^2+\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^4}X_a\underset{¯}{\psi }\right)^2}}}`$ $`=`$ $`\mathrm{sin}\left(\alpha _a\right)\text{,}`$ $`{\displaystyle \frac{\left(\underset{¯}{\psi }^{}\underset{¯}{\gamma ^0}X_b\underset{¯}{\psi }\right)}{\sqrt{\left(\underset{¯}{\psi }^{}\underset{¯}{\gamma ^0}X_b\underset{¯}{\psi }\right)^2+\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^4}X_b\underset{¯}{\psi }\right)^2}}}`$ $`=`$ $`\mathrm{cos}\left(\alpha _b\right)\text{.}`$ $`{\displaystyle \frac{\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^4}X_b\underset{¯}{\psi }\right)}{\sqrt{\left(\underset{¯}{\psi }^{}\underset{¯}{\gamma ^0}X_b\underset{¯}{\psi }\right)^2+\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^4}X_b\underset{¯}{\psi }\right)^2}}}`$ $`=`$ $`\mathrm{sin}\left(\alpha _b\right)`$ Let: $$\underset{¯}{\gamma }=\left(\mathrm{cos}\left(\alpha _a\right)\underset{¯}{\gamma ^0}+\mathrm{sin}\left(\alpha _a\right)\underset{¯}{\beta ^4}\right)X_a+\left(\mathrm{cos}\left(\alpha _b\right)\underset{¯}{\gamma ^0}+\mathrm{sin}\left(\alpha _b\right)\underset{¯}{\beta ^4}\right)X_b\text{.}$$ In this case: $$\underset{¯}{\gamma }\underset{¯}{\gamma }=1_8$$ and if $`1k3`$ then $$\underset{¯}{\gamma }\underset{¯}{\beta ^k}=\underset{¯}{\beta ^k}\underset{¯}{\gamma }$$ and the Euler-Lagrange equation for $`_\psi `$ is the following: $$\left(i\underset{¯}{\beta ^\mu }_\mu m_\psi \underset{¯}{\gamma }\right)\underset{¯}{\psi }=0\text{.}$$ Since $$\alpha _a=\alpha _b$$ then $$\underset{¯}{\gamma }=\mathrm{cos}\left(\alpha _a\right)\underset{¯}{\gamma ^0}+\mathrm{sin}\left(\alpha _a\right)\underset{¯}{\beta ^4}\text{.}$$ Let $`\underset{¯}{\psi }`$ be a plane wave electron spinor with a positive energy: $$\underset{¯}{\psi }=\left(a_1\left[\begin{array}{c}0\\ 0\\ 0\\ 0\\ 1\\ 0\\ \frac{p_z}{E+m_e}\\ \frac{p_x+ip_y}{E+m_e}\end{array}\right]+a_2\left[\begin{array}{c}0\\ 0\\ 0\\ 0\\ 0\\ 1\\ \frac{p_xip_y}{E+m_e}\\ \frac{p_z}{E+m_e}\end{array}\right]\right)\mathrm{exp}\left(ipx\right)\text{,}$$ here: $`a_1`$, $`a_2`$ are complex, and $`E=\sqrt{p^2+m_e^2}`$. In this case: $$\mathrm{cos}\left(\alpha _a\right)=1\text{.}$$ Hence $$\underset{¯}{\gamma }=\underset{¯}{\gamma ^0}$$ and the Euler-Lagrange equation is the following: $$\left(i\underset{¯}{\gamma ^\mu }_\mu m_\psi \right)\underset{¯}{\psi }=0\text{.}.$$ ## 3 Necessary Conditions Let $`U`$ be any $`8\times 8`$ complex matrix for which the Lagrangian $$_0=0.5i\left(\underset{¯}{\psi }^{}\underset{¯}{\beta ^\mu }\left(_\mu \underset{¯}{\psi }\right)\left(_\mu \underset{¯}{\psi }\right)^{}\underset{¯}{\beta ^\mu }\underset{¯}{\psi }\right)$$ is invariant. Hence $$\underset{¯}{U}^{}\underset{¯}{U}=1_8$$ and for all $`\mu `$ ($`1\mu 3`$): $$\underset{¯}{U}\underset{¯}{\beta ^\mu }=\underset{¯}{\beta ^\mu }\underset{¯}{U}\text{.}$$ $`\underset{¯}{U}`$ must be of the following type from this commutativity : $$\underset{¯}{U}=\left[\begin{array}{cccccccc}z_{1,1}& 0& 0& 0& z_{1,5}& 0& 0& 0\\ 0& z_{1,1}& 0& 0& 0& z_{1,5}& 0& 0\\ 0& 0& z_{3,3}& 0& 0& 0& z_{3,7}& 0\\ 0& 0& 0& z_{3,3}& 0& 0& 0& z_{3,7}\\ z_{5,1}& 0& 0& 0& z_{5,5}& 0& 0& 0\\ 0& z_{5,1}& 0& 0& 0& z_{5,5}& 0& 0\\ 0& 0& z_{7,3}& 0& 0& 0& z_{7,7}& 0\\ 0& 0& 0& z_{7,3}& 0& 0& 0& z_{7,7}\end{array}\right]$$ (here $`z_{j,k}`$ are a complex) and from the unitarity if $$z_{j,k}=x_{j,k}+iy_{j,k}$$ then $$\begin{array}{c}1x_{1,5}^2y_{1,5}^2y_{5,5}^20\text{,}\\ x_{1,1}=\sqrt{1x_{1,5}^2y_{1,5}^2y_{5,5}^2}\text{,}\\ x_{5,5}=x_{1,1}\text{,}\\ x_{5,1}=x_{1,5}\text{,}\\ y_{1,1}=y_{5,5}\text{,}\\ y_{5,1}=y_{1,5}\end{array}$$ and $$\begin{array}{c}1x_{3,7}^2y_{3,7}^2y_{7,7}^20\text{,}\\ x_{3,3}=\sqrt{1x_{3,7}^2y_{3,7}^2y_{7,7}^2}\text{,}\\ x_{7,7}=x_{3,3}\text{,}\\ x_{7,3}=x_{3,7}\text{,}\\ y_{3,3}=y_{7,7}\text{,}\\ y_{7,3}=y_{3,7}\text{.}\end{array}$$ If $$\begin{array}{c}x_{3,7}=0\text{,}\\ y_{3,7}=0\text{,}\\ y_{7,7}=0\end{array}$$ then $$z_{3,3}=1$$ and $`\underset{¯}{U}`$ is the matrix of type (1). In this case the mass form (2) is invariant for $`\underset{¯}{U}`$ and a right-handed particles do not interact by this transformation. Therefore if an electron has got a nonzero mass, provided with the mass form (2), then all neutrinos must be left-handed. Like this, for $`z_{1,1}=1`$, all antineutrinos must be right-handed to an positron has got a nonzero mass. If $`z_{1,1}1`$ and $`z_{3,3}1`$ then a mass form, invariant for $`\underset{¯}{U}`$, does not exist. ## 4 Acknowledgment Thanks very much to Prof. V. V. Dvoeglazov for his papers which he had sent me kindly in 1998.
no-problem/9906/cond-mat9906316.html
ar5iv
text
# Single particle and collective excitations in the one-dimensional charge density wave solid K0.3MoO3 probed in real time by femtosecond spectroscopy. ## Abstract Ultrafast transient reflectivity changes caused by collective and single particle excitations in the quasi one-dimensional charge-density wave (CDW) semiconductor K<sub>0.3</sub>MoO<sub>3</sub> are investigated with optical pump-probe spectroscopy. The temperature-dependence of non-equilibrium single particle excitations across the CDW gap and their recombination dynamics are reported for the first time. In addition, amplitude mode reflectivity oscillations are observed in real time. A $`T`$-dependent overdamped response is also observed which is attributed to relaxation of the phason mode. Molybdenum oxides like K<sub>0.3</sub>MoO<sub>3</sub> and Rb<sub>0.3</sub>MoO<sub>3</sub> are well known for their interesting electronic properties arising from their one-dimensional (1D) chain structure. At room temperature they are highly anisotropic 1D metals. Upon cooling, they become susceptible to a Peierls instability on the 1D chains causing fluctuating local CDW ordering. Upon further cooling, as fluctuations are reduced, inter-chain interactions cause the CDWs on individual chains to become correlated, eventually undergoing a second-order phase transition to a three-dimensionally (3D) ordered state below $`T_c=183`$ K. The formation of a 3D CDW ordered state is concurrent with the appearance of a gap $`\mathrm{\Delta }_{CDW}`$ in the single particle (SP) excitation spectrum, while the collective excitations of the 3D CDW state are described by an amplitudon mode (AM) and a phase mode (phason). In this Letter we report femtosecond time-domain transient reflectivity measurements on K<sub>0.3</sub>MoO<sub>3</sub> enabling for the first time real-time observation of the reflectivity modulations caused by collective CDW excitations. We report the $`T`$-dependence of the amplitude $`A(T),`$ frequency $`\omega _A(T),`$ and damping constant $`\tau _A(T)`$ of the AM and for the first time the $`T`$-dependence of the phason damping constant $`\tau _p(T)`$. We also report the $`T`$-dependence of electron-hole recombination lifetime $`\tau _s`$ across the CDW gap below, as well as above $`T_c`$. Up till now, enhaced coherent phonon oscillations associated with the formation of a CDW were observed below $`T_{c1}`$ in Mo<sub>4</sub>O<sub>11</sub>, but to our knowledge real-time observation of collective and SP excitations in CDW systems have not yet been reported. In these experiments, an ultrashort laser pump pulse first excites electron-hole pairs via an interband transition in the material (step 1 in Fig.1a)). In a process which is similar in most materials including metals, semiconductors and superconductors, these hot carriers very rapidly release their energy via $`ee`$ and $`eph`$ collisions reaching states near the Fermi energy within $`\tau _i=10100`$ fs (step 2 in Fig.1a)) acting as an ultrashort SP injection pulse. If a CDW or superconducting gap is present in the SP excitation spectrum, it inhibits the final relaxation step resulting in a relaxation bottleneck and photoexcited carriers accumulate above the gap. This causes a transient change in reflectivity $`\Delta /`$ due to change in dielectric constant arising from excited state absorption processes of the type shown in step 3 in Fig.1a). The density of these accumulated photoinduced (PI) carriers, $`n_{sp}^{}`$ can thus be determined as a function of temperature and time after photoexcitation from the transient reflectivity change $`\Delta _s/S(T)e^{t/\tau _s}`$, where $`\tau _s`$ is the characteristic SP recombination time. The amplitude is given by $`S(T)n_{sp}^{}\rho _2\left|M_{12}\right|^2`$, where $`M_{12}`$ is the matrix element for the $`E_1`$ $``$ $`E_2`$ optical transition (Fig.1a)) and $`\rho _2`$ is the density of states in level $`E_2`$. Whereas $`\rho _2`$ and $`M_{12}`$ can be assumed to be $`T`$-independent in first approximation, $`n_{sp}^{}`$ is strongly $`T`$-dependent when $`k_BT\mathrm{\Delta }_{CDW}.`$ A $`T`$-dependence of $`n_{sp}^{}`$ has recently been calculated for various gap situations, which we can now compare with experiments. In addition to the transient change of reflectivity due to SP excitations discussed above, a transient reflectivity signal is expected also from collective modes. The AM is of $`A_1`$ symmetry and involves displacements of ions about their equilibrium positions $`Q_0`$, which depend on the instantaneous surrounding electronic density $`n(t)`$. Since $`\tau _i<\mathrm{}/\omega _A,`$ the SP injection pulse may be thought of as a $`\delta `$-function-like perturbation of the charge density $`n_{sp}`$ and the injection pulse acts as a time-dependent displacive excitation of the ionic equilibrium position $`Q_0(t).`$ The response of the AM to this perturbation is a modulation of the reflectivity $`\Delta _A/`$ of the form $`A(T)e^{t/\tau _A}\mathrm{cos}(\omega _At+\varphi _0)`$ by the displacive excitation of coherent phonons (DECP) mechanism, known from femtosecond experiments on semiconductors and superconductors . The $`\delta `$-function-like SP injection pulse also gives rise to the displacement of charges with respect to the ions, directly exciting the CDW phason. Since this is infrared-active, we expect the resulting change of the dielectric constant $`\mathrm{\Delta }ϵ/ϵ`$ to lead to a directly observable reflectivity transient, which for small $`\mathrm{\Delta }ϵ`$ can be approximated as $`\Delta _p/\mathrm{\Delta }ϵ/ϵ.`$ In equilibrium, we expect the phason to be pinned and at a finite frequency, but in non-equilibrium situation such as here, where the excess carrier kinetic energy may easily exceed the de-pinning energy, the mode may be de-pinned. In this case we may expect an overdamped reflectivity transient written as $`P(T)e^{t/\tau _p}\mathrm{cos}(\omega _pt+\varphi ),`$ with $`\omega _p0`$, but with a damping constant which is expected to be similar to that of the AM $`\tau _p\tau _A`$, i.e. $``$10 ps. Summing all the contributions, in K<sub>0.3</sub>MoO<sub>3</sub> the photoinduced transient reflectivity signal is of the form: $`\mathrm{\Delta }(t,T)/`$ $`=`$ $`A(T)e^{t/\tau _A}\mathrm{cos}(\omega _At+\varphi _0)`$ (2) $`+P(T)e^{t/\tau _p}+S(T)e^{t/\tau _s}+B(T).`$ For completeness we have included an additional term $`B(T)`$ due to a long-lived background signal, which is also observed experimentally, and whose lifetime is longer than the inter-pulse separation of $`12`$ns. The different contributions to $`\Delta /`$ can be effecively distinguished experimentally by their very different time-dynamics, polarization- and $`T`$-dependences. In the experiments reported here a mode-locked Ti:Sapphire laser with pulselength $`\tau _L100`$ fs at 800 nm was used. The PI change in reflectivity $`\Delta /`$ was measured using a photodiode and lock-in detection. The pump laser power was kept below 5 mW, exciting approximately 10<sup>18</sup>-10<sup>19</sup> carriers per cm<sup>3</sup>, and the pump/probe intensity ratio was $``$100. The steady-state heating effect was accounted for as described in Ref.. The experiments were performed on freshly cleaved K<sub>0.3</sub>MoO<sub>3</sub> single crystals with the laser polarization in the a-b plane, a being the direction and b is the chain direction. The orientation of the crystal was determined by using an atomic force microscope, by the direction of the Mo-O chains. In Fig.1a) we show $`\mathrm{\Delta }/`$ as a function of time at different temperatures. Below $`T_c`$, an oscillatory component is observed on top of a negative induced reflection, the latter exhibiting a fast initial decay followed by a slower decay. As $`T_c`$ is approached from below, the oscillatory signal dissappears, while the fast transient signal remains observed well above $`T_c,`$ as shown by the trace at 210K. For a quantitative analysis, we separate the different components of the signal according to their $`T`$-dependence and probe polarization anisotropies. In Fig.1b) we show the signal at $`T=45`$ and 110K with the oscillatory component and background $`B(T)`$ subtracted. The logarithmic plot enables us to clearly identify two components with substantially different lifetimes, one with $`\tau _s`$ 0.5 ps, and the other with $`\tau _p10`$ ps at low $`T.`$ Their amplitudes and relaxation times are analyzed by fitting two exponentials (Eq.(1)). For reasons which will become apparent, we attribute them to the SP relaxation $`S(T)`$ and phason relaxation $`P(T)`$ respectively. We note that $`P(T)`$ displays no sign of oscillatory response, in accordance with the expectation that the phason relaxation is overdamped in this type of experiment. The insert shows the dependence of the amplitude of the fast signal on the probe pulse polarization, showing maximum amplitude for $`\stackrel{}{E}a`$. In Fig.1c) we have plotted only the oscillatory component with its FFT spectrum, showing a peak at $`\nu _A=`$ 1.7 THz. In contrast to the transient signal the amplitude of the oscillatory signal is independent of polarization (Fig.2a)). The $`T`$-dependences of the single oscillatory component frequency $`\nu _A`$ and damping $`\mathrm{\Gamma }_A=1/(\pi \tau _A)`$ derived from fits to the real-time oscillations are shown in Fig.2a). Since the oscillation frequency $`\nu _A`$ shows clear softening as T<sub>c</sub> is approached from below, the contribution from coherent phonons as observed in Ref. can be excluded. The measured $`\nu _A`$ and $`\mathrm{\Gamma }_A`$ closely follow the expected behaviour for the AM and are in good agreement with previous spectroscopic neutron and Raman data. The amplitude of the modulation $`A(T)`$ falls rather more rapidly with $`T`$ than $`\nu _A`$ and is rather isotropic in the $`ab`$ plane. In Fig.2b) and c) we have plotted the $`T`$-dependence of $`\tau _p`$ and $`P(T)`$. At $`T=`$50 K, $`\tau _p=12\pm 2`$ ps, in agreement with the $`\mathrm{\Gamma }=0.050.1`$ THz linewidths of the pinned phason mode in microwave and IR experiments. With increasing temperature $`\tau _p`$ is approximately constant up to 90 K and then falls rapidly as $`TT_c`$. The decrease of $`\tau _p`$ near $`T_c`$ is consistent with increasing damping due to thermal phase fluctuations arising from coupling with the lattice and SP excitations. The amplitude $`P(T)`$ exhibits somewhat different behaviour. It appears to first show an increase with increasing $`T`$, and then drops as $`TT_c`$. Such $`T`$-dependence behaviour has been previously observed - but not yet satisfactorily explained - for the threshold field in some non-linear conductivity experiments. Let us now turn to the transient reflectivity signal due to photoexcited SP excitations. The $`T`$-dependence of the PI signal amplitude below $`T_c`$ for a $`T`$-dependent gap $`𝚫(T)`$ \- for simplicity using a BCS-like $`T`$-dependence - is given by: $$S(T)n_{sp}^{}=\frac{_I/(𝚫(T)+k_BT/2)}{1+\gamma \sqrt{\frac{2k_BT}{\pi 𝚫(T)}}\mathrm{exp}(𝚫(T)/k_BT)},$$ (3) where $`_I`$ is the pump laser intensity per unit cell and $`\gamma `$ is a constant, depending on the materials’ parameters. Plotting Eq.(2) as a function of temperature in Fig.3a), we find that the amplitude $`S(T)`$ obtained from the fits to the data agrees remarkably well with the theory for $`T<T_c`$: $`S(T)`$ is nearly constant up to nearly 100 K, then increases slightly and then drops very rapidly near $`T_c`$. The value of the gap $`\mathrm{\Delta }(0)`$=850K$`\pm 100`$K obtained from the fit of Eq.(2) with $`\gamma =`$ 10 is in good agreement with other measurements . In contrast to the response of the collective modes $`A(T)`$ and $`P(T),`$ both of which dissappear within 10-20 K below $`T_c`$, $`S(T)`$ remains observable up to nearly 240 K, i.e. appears to show a pseudogap up to 50 K above $`T_c.`$ This is - in contrast to the behaviour below $`T_c`$ \- clearly incompatible with a BCS-like description of the gap and suggests the fluctuating presence of the gap well above $`T_c`$. The polarization anisotropy of the signal $`S(T)`$ for $`T>T_c`$ is the same as for $`T<T_c`$ (Fig.1b)), strongly suggesting that the origin of the signal $`S(T)`$ above $`T_c`$ is the same as below $`T_c`$ i.e. SP gap excitations. The $`T`$-dependence of the relaxation time $`\tau _s`$ (Fig.3b)) is qualitatively different to $`\tau _p`$ and $`\tau _A`$. As $`TT_c,`$ $`\tau _s`$ appears to diverge and then drops to $`\tau _s0.25`$ ps above $`T_c`$. Such behaviour is in agreement with expected $`T`$-dependence of SP relaxation across the gap. The dominant recombination mechanism across the gap is phonon emission via phonons whose energy $`\mathrm{}\omega _p>2\mathrm{\Delta }`$. As the gap closes near $`T_c`$, more low-energy phonons become available for reabsorption and the recombination mechanism becomes less and less efficient. The recombination lifetime near $`T_c`$ can be shown to be inversely proportional to the gap as $`\tau _s1/\mathrm{\Delta }(T)`$. The solid line in Fig.3b) shows a fit to the data using a BCS-like $`T`$-dependent gap $`\mathrm{\Delta }_{BCS}(T)`$ with $`T_c^{3D}`$=183 K. To complete the data analysis, we show in Fig.4 the amplitude of the slowly-decaying background signal $`B(T)`$ as a function of $`T`$. Its anomalous $`T`$-dependence clearly rules out a thermal origin. The lifetime $`\tau >10^8`$s deduced from the amplitude of $`B(T)`$ at ”negative times” i.e. from the preceding pulse, strongly suggests excitations involving localized states. From a fit to an Arrhenius law $`B(T)=B_0\mathrm{exp}[E_a/k_BT],`$ we obtain an activation energy $`E_a/\mathrm{\Delta }(0)0.6\pm 0.2`$, suggesting that the process involves the excitation of carriers from intra-gap states into the SP continuum $`E_1`$. As the gap closes, excitations from intra-gap states to the SP states are no longer possible, explaining the $`T`$-dependence of the signal above $`T_c`$. The microscopic origin of these states is most likely trapped defects, but the possibility of a collective excitation cannot be excluded at this stage. We note that a long-lived signal with a similar $`T`$-dependence was recently observed in time-resolved experiments on the cuprate superconductor YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7-δ</sub> and attributed to localized intra-gap states. The real-time optical data presents some qualitatively new information on the SP and collective excitations in quasi-1D materials. We have found that because of the qualitatively different time-, polarization- and $`T`$-characteristics, the responses of the different components can be effectively separated. Apart from directly extracting the $`T`$-dependences of $`A(T)`$, $`\nu _A(T)`$ and $`\tau _A(T)`$, we also observe an overdamped mode, which we have assigned to relaxation of the phason mode. In addition to the observation of the $`T`$-dependence of photoinduced SP population as predicted by theory, we find - also in agreement with calculations \- that the SP recombination time across the gap diverges as $`\tau _s1/\mathrm{\Delta }`$ as $`TT_c`$. From the fact that the SP population appears to persist above $`T_c`$, the data shows clear evidence for the existence of a pseudogap for SP excitations above $`T_c`$ and suggests the fluctuating presence of a SP gap, rather than fluctuations of the order parameter, which would appear as a tail also in the SP relaxation time $`\tau _s`$ above $`T_c`$ \- but does not. Finally, we should mention that many of the features, particularly the $`T`$-dependence of the SP excitation amplitude and the SP recombination lifetime is very similar to the behaviour recently reported in cuprates. We note that although coherent oscillations were reported in YBa<sub>2</sub>Cu<sub>3</sub>O<sub>7-δ</sub>, no $`T`$-dependence of the oscillation frequency was reported, and they were attributed not to any collective electronic mode, but to $`c`$-axis phonons. FIG 1. a) The transient reflection $`\mathrm{\Delta }R/R`$ from K<sub>0.3</sub>MoO<sub>3</sub> after photoexcitation by a 100 fs laser pulse at a number of temperatures above and below $`T_c^{3D}=183`$ K. The constant background signal $`B(T)`$ was substracted and offset for clarity. b) The time-evolution of transient signal with the oscillatory component subtracted shown at $`T`$=45 K and 110 K, displayed on a logarithmic scale to emphasise the difference in relaxation times $`\tau _p`$ and $`\tau _s`$. The insert shows the amplitude $`S(T)`$ as a function of probe pulse polarization below (solid squares), and above $`T_c^{3D}`$ (open circles) with respect to the crystal direction. c) The oscillatory transient signal $`\mathrm{\Delta }R_A/R`$ after subtraction of the decay components $`B`$, $`S`$ and $`P`$. The fit is made using the first term in Eq.(1). The insert shows the FFT spectrum of the signal. FIG 2. a) The amplitude $`A(T)`$ (diamonds), frequency $`\nu _A`$ (squares) and damping constant $`\mathrm{\Gamma }_A=1/(\pi \tau _A)`$ (full circles) as a function of $`T`$. Data from Refs. (open circles) and (open triangles) are also included. The insert shows the amplitude $`A(T)`$ as a function of probe polarization with respect to the crystal direction. b) $`\tau _p`$ as a function of $`T`$ (circles). The amplitudon decay time $`\tau _A`$ is plotted for comparison (squares). c) $`P(T)`$ as a function of $`T`$. FIG 3. a) The $`T`$-dependence of $`S(T)`$ . The fit to the data for $`S(T)`$ is shown using Eq.(2) with a BCS-like gap $`\mathrm{\Delta }_{BCS}(T)`$ opening at $`T_c^{3D}=183`$ K. b) The SP relaxation time $`\tau _s`$ as a function of $`T`$ and a fit using Eq.(25) of Ref.. FIG 4. The $`T`$-dependence of the long-lived signal $`B\left(T\right)`$. The line is an Arrhenius fit of the signal amplitude below $`T_{c\text{ }}`$with $`E_a=60`$meV. The insert shows the amplitude $`B(T)`$ as a function of probe pulse polarization with respect to the crystal direction. The authors wish to acknowledge L.Degiorgi and P.Monceau for very useful discussions and K.Kočevar for the AFM analysis. A part of this work was performed under the auspices of the EU ULTRAFAST project.
no-problem/9906/nucl-th9906072.html
ar5iv
text
# Excitation function of strangeness in A+A reactions from SIS to RHIC energies11footnote 1Supported by BMBF and GSI Darmstadt ## 1 INTRODUCTION The aim of high energy heavy-ion collisions is to investigate nuclear matter under extreme conditions, i.e. high temperature and density. The most exciting prospect is the possible observation of a signal for a phase transition from normal nuclear matter to a nonhadronic phase, where partons are the basic degrees of freedom. In this context strangeness enhancement in heavy-ion collisions compared to proton-proton collisions has been suggested as a possible signature for the phase transition . On the other hand, precursor effects might already be seen at SIS energies since densities up to 3$`\times \rho _0`$ can be achieved in central collisions of heavy nuclei and the effect of meson potentials can be studied with a higher sensitivity to the productions thresholds, respectively. In this contribution a brief survey is presented on the information gained so far in comparison of experimental data to nonequilibrium transport theory, here the Hadron-String-Dynamics (HSD) approach . For a more detailed discussion of the issues presented the reader is refered to . ## 2 ANALYSIS OF EXPERIMENTAL DATA Since the real part of the actual $`K^+`$ and $`K^{}`$ self-energy $`\mathrm{\Pi }_{K^\pm }`$ in hot and dense nuclear matter is quite a matter of debate we adopt a more practical point of view and as a guide for the analysis use a linear extrapolation of the kaon potential with density $`\rho _B`$ as $$m_K^{}(\rho _B)=m_K^0\left(1\alpha \frac{\rho _B}{\rho _0}\right),\mathrm{i}.\mathrm{e}.V_K=\alpha m_K^0\frac{\rho _B}{\rho _0},$$ (1) with $`\alpha _{\overline{K}}`$ 0.2-0.25 for antikaons and $`\alpha _K0.06`$ for kaons in line with Refs. . In (1) a momentum dependence of the kaon or antikaon potential has been neglected for reasons of numerical simplicity. The dispersion analysis of Sibirtsev et al. shows that this is roughly fulfilled for the kaon potential, however, the antikaon potential should be more strongly momentum dependent. The Lorentz invariant $`K^+`$ spectra for Ni + Ni at 0.8, 1.0 and 1.8 A$``$GeV are shown in Fig. 1 (l.h.s.) in comparison to the data from the KaoS Collaboration . Here the full lines reflect calculations including only bare $`K^+`$ masses ($`\alpha _K=0`$) while the dashed lines correspond to calculations with $`\alpha _K=0.06`$ in Eq. (1), which leads to an increase of the kaon mass at $`\rho _0`$ by about 30 MeV. The general tendency seen at all bombarding energies is that the calculations with a bare kaon mass seem to provide a better description of the experimental data for Ni + Ni than those with an enhanced kaon mass. This trend continues to hold also for the light system C + C as well as for the heavy systems Ru + Ru and even Au + Au within the cross sections for the $`\mathrm{\Delta }`$ induced channels used. The kaon flow in the reaction plane shows some sensitivity to the kaon potential in the nuclear medium as suggested by Li, Ko and Brown . Here due to elastic scattering with nucleons the kaons partly flow in the direction of the nucleons thus showing a positive flow in case of no mean-field potentials. With increasing repulsive kaon potential the positive flow will turn to zero and then become negative. In fact, experimental data on kaon flow indicate a repulsive potential for kaons in the nuclear medium at SIS as well as AGS energies . We now turn to the production of antikaons which do clearly show the effect from attractive potentials in the medium. We recall that for $`\alpha _{\overline{K}}`$ = 0 in Eq. (1) we recover the limit of vanishing antikaon self-energy, whereas for $`\alpha _{\overline{K}}`$ 0.2 we approximately describe the scenario of Refs. . The $`K^{}`$ spectra for Ni + Ni at 1.85 and 1.66 A$``$GeV are shown in Fig. 1 (r.h.s.) for $`\alpha _{\overline{K}}`$ = 0, 0.2 and 0.24 where the latter cases correspond to an attractive potential of $`100`$ and $`120`$ MeV at density $`\rho _0`$, respectively. We note, that due to the uncertainties involved in the elementary $`BB`$ production cross sections we cannot determine this value very reliably. With increasing $`\alpha _{\overline{K}}`$ not only the magnitude of the spectra is increased, but also the slope becomes softer. This is most clearly seen at low antikaon momenta because the net attraction leads to a squeezing of the spectrum to low momenta . Whereas the kaon and antikaon dynamics at SIS energies is reasonably described within the hadron-string transport approach when including meson potentials , this no longer holds at AGS energies . The heaviest system studied here is Au + Au at $``$ 11 A$``$GeV. In the ’bare mass’ scenario the data are underestimated strongly while for $`\alpha _K=0.06`$ and $`\alpha _{\overline{K}}`$ = 0.24 the situation improves significantly . Whereas the $`K^{}`$ yield is almost reproduced in the latter scheme, the $`K^+`$ yield is still underestimated as in case of the Si + Al and Si + Au system at 14.6 A$``$GeV . Without explicit representation we note that for all systems at SPS energies the $`h^{}`$, $`K^+`$ and $`K^{}`$ distributions are reproduced rather well showing even a tendency for an excess of kaons and antikaons in the calculations rather than missing strangeness. At RHIC energies we also find a sizeable enhancement of the $`K^\pm `$ yield in central Au + Au compared to $`pp`$ collisions or pure parton (VNI) cascade calculations due to a long lasting hadronic rescattering phase . The E866 and E895 Collaborations, furthermore, have measured Au + Au collisions at 2,4,6 and 8 A$``$GeV kinetic energy at the AGS . Thus it is of particular interest to look for a discontinuity in the excitation functions for pion and kaon rapidity distributions and to compare them to the hadron-string transport approach. In Fig. 2 the calculated $`K^+/\pi ^+`$ ratios (open squares) at midrapidity ($`|y_{cm}|`$ 0.25) for central (b=2 fm) Au + Au collisions at 1,2,4,6,8 and 11 A$``$GeV and Pb + Pb collisions at 160 A$``$GeV are shown together with the preliminary data (full dots). The ratio at midrapidity is slightly higher than the total $`K^+/\pi ^+`$ ratio, because the kaon rapidity distribution is narrower than that of the pions. While the scaled kaon yield at 1 and 2 A$``$GeV (SIS energies) is well described in the HSD approach within the errorbars, the experimental $`K^+/\pi ^+`$ ratio at 4 A$``$GeV is underestimated already by a factor of 2 and increases up to roughly 19% for 11 A$``$GeV. As mentioned before the calculated and measured ratio coincide again at 160 A$``$GeV. Data at 40 A GeV from the SPS as well as at 21.5 A TeV from RHIC are expected to come up soon and to complete the picture from the experimental side. ## 3 SUMMARY We find an enhancement of the $`K^+/\pi ^+`$ ratio in heavy-ion collisions relative to p + p reactions due to hadronic rescatterings both with increasing system size and energy. The excitation function in the $`K^+/\pi ^+`$ ratio from the HSD transport approach has a similar slope in nucleus-nucleus and p + p collisions (cf. Fig. 2) indicating a monotonic increase of strangeness production with bombarding energy. However, the experimental $`K^+/\pi ^+`$ ratio for central Au + Au collisions at midrapidity increases up to $`19\%`$ at 11 A$``$GeV – it is unknown if a local maximum will be reached at this energy – and decreases at SPS energies to $`16.5\%`$. Such a decrease of the scaled kaon yield from AGS to SPS energies is hard to obtain in a hadron-string transport model. On the contrary, the higher temperatures and particle densities at SPS energies allways tend to enhance the $`K^+/\pi ^+`$ yield closer to its thermal equlibrium value of $`2025\%`$ at chemical freezout and temperatures of $`T160`$ MeV. Thus the steep rise of the strangeness yield and its decrease suggests the presence of nonhadronic degrees of freedom which might become important already at about 4 A$``$GeV in central Au + Au collisions.
no-problem/9906/cond-mat9906208.html
ar5iv
text
# Binocular disparity can explain the orientation of ocular dominance stripes in primate V1 ## I INTRODUCTION The perception of depth in primates relies on recombining information coming from both eyes. This is accomplished by a retinotopic mapping , of the two retinal images onto the primary visual area V1. In many primates, neurons dominated by each eye are segregated into the system of alternating stripes known as the ocular dominance pattern ,. A complete reconstruction of the pattern in a macaque V1 is shown in Fig.1. Although this pattern may seem random, the orientation of the ocular dominance stripes on the cortical surface follows systematic trends found in other macaques and in Cebus monkeys . These trends are easiest to see when the ocular dominance pattern is transformed back into visual field coordinates by dividing all cortical distances by the local magnification factor. (The magnification factor gives the distance in millimeters on the cortex which corresponds to a $`1\mathrm{deg}`$ separation on the retina.) The transformed pattern shown in Fig.2 (as obtained by LeVay et al. (1985) following Hubel and Freeman (1977)) reveals two major trends: in the parafoveal region stripes tend to run horizontally, while farther from the fovea stripes follow roughly concentric circles. These trends in the orientation of the stripes call for explanation. Many theorists have successfully modeled the development of ocular dominance columns ,,, and some have addressed the trends in stripe orientation. It was suggested originally by LeVay et al. (1985) and later investigated by others ,, that the mapping from the two almost circular LGN layers to the more elongated representation in V1 requires least stretching (or anisotropy of the magnification factor within ocular dominance stripes) if the stripes run perpendicular to the long axis of V1. However, this theory does not explain the different orientation of ocular dominance stripes in the parafoveal region. Moreover, it is unlikely that the shape of V1 dictates its internal organization; more probably, its internal organization will dictate its shape. Goodhill et al. (1997) have instead proposed that the global pattern of ocular dominance stripes arises from anisotropic and spatially non-uniform correlations in the neural input from the retinae. This seems like a plausible developmental mechanism, although anisotropic correlations have yet to be demonstrated experimentally. Rather than modeling development, I have taken a different approach to explain the orientation of ocular dominance stripes. I propose that the stripe orientations follow from V1’s role in depth perception according to the principle of wiring economy. In other words, I focus on understanding why the stripes are arranged as they are, rather than how they become so arranged. The wiring economy principle amounts to the following ,,,, ,: because of limitations on head size, there is pressure to keep the volume of the cortex to a minimum. This implies that wiring, i.e. axons and dendrites, should be as short as possible, while maintaining function. In general, the function of a given cortical circuit specifies the connections between neurons. Therefore the problem presented by the wiring economy argument is to find, for a given set of connections, the spatial layout of the neurons that minimizes wiring length. This problem can be extremely difficult because of the large number of neurons and even larger number of interconnections within a cortical region. However, the columnar organization of the cortex allows me to consider the layout of cortical columns (each consisting of $`10^4`$ neurons) rather than individual neurons, reducing the problem to two dimensions. The wiring economy principle has been used to explain retinotopic map , and ocular dominance stripes , in V1. In the first case, the construction of receptive fields requires connections between neurons representing neighboring points in the visual field. Topographic mapping of the visual field minimizes the length of these connections. In the second case, each cortical neuron connects more often to cortical neurons dominated by the same eye than to neurons dominated by the opposing eye. Thus the segregation of neurons into alternating monocular stripes minimizes the total length of intra-cortical connections. Here I use the principle of wiring economy to find the optimal orientation of the ocular dominance stripes (given they exist) from the function of V1 in processing binocular disparity. Disparity arises when an object closer or farther than the point of fixation forms images on the two retinas that are in different positions relative to the fovea. Because the cortex is retinotopically mapped, the left and the right eye representations of this object in the cortex will fall some distance away from each other, as determined by the magnitude and direction of the retinal separation. Recombining these representations requires extensive wiring between cortical columns. In the next Section, I show that the left and right eye representations in V1 will be farther apart if the ocular dominance stripes run perpendicular, rather than parallel, to the direction of retinal separation. Therefore I argue that in order to minimize wirelength, the orientation of the stripes should correspond to the direction of binocular disparity. However, for any given point on the cortex the disparity in the corresponding point of the visual field depends on the viewing conditions, that is, direction of gaze and distance to the object. In Section III, I calculate a distribution of disparities averaged over viewing conditions. I obtain a map of typical disparities in the visual field which then determines the optimal orientation of the ocular dominance stripes. The orientation of stripes predicted by this theory agrees with actual patterns obtained from macaque and Cebus monkeys, see Section IV. My results show that the two major trends of stripe orientation result from two main contributions to disparity. In the parafoveal region, binocular disparity is due mainly to the horizontal displacement of the eyes, consistent with the horizontal stripes in the (transformed) ocular dominance pattern. Farther from the fovea, the pattern consists of isoeccentric lines. These are explained by binocular disparity due to unequal rotation of the eyeballs around the gaze line (cyclotorsion). ## II Disparity direction determines orientation of ocular dominance stripes Because the left and the right eye pathways do not converge before V1, the existence of binocular neurons in V1 suggests that the information from both eyes is recombined there. Moreover, many binocular neurons in V1 are disparity-tuned. , This requires intra-cortical wiring which connects cortical columns containing left/right eye representations of an object. To minimize the wirelength, the distance between these columns should be as small as possible. For a given magnitude of binocular disparity, the distance between the columns containing left/right representations of an object depends on the orientation of ocular dominance stripes relative to the separation of the columns. To see this consider two alternative arrangements: ocular dominance stripes oriented perpendicular, Fig.3a, or parallel, Fig.3b, to the separation of the columns. One can think of V1 as being composed of interleaved stripes cut from the two topographic maps belonging to the two eyes. If one were to move across the stripes the representation of every point in the visual field is encoutered twice: once in a right-eye column, once in a left-eye column. Therefore the separation between the two columns containing left/right eye representations of the object is twice as big if the stripes are perpendicular compared to parallel to the separation between the columns. Thus for a given magnitude of disparity in the visual field the length of inter-eye connections is minimized if the ocular dominance stripes run in the direction corresponding to the disparity direction. Several assumptions were made in this argument. First, I assumed the absence of stretching within the stripes which could change the distances across vs. along the stripes. To see whether this is a valid assumption I restate the argument in terms of the cortical magnification factor which has been measured experimentally. The separation between the two columns in V1 is given by the retinal disparity times the magnification factor. If the magnification factor across the stripes is greater than that along the stripes, the separation between the columns and hence the length of inter-eye connections is minimized by aligning the stripes with the disparity direction. Experiments show that the cortical magnification factor along the stripes is, indeed, about 1.5 times smaller than that across the stripes. Therefore, even though some stretching exists, my argument remains valid. Second, I did not include wirelength of intra-eye connections in the cortex. These connections are responsible for monocular functions of V1 such as processing of contour orientation and color. The reason for neglecting these connections is their isotropy, that is they do not depend on the direction in the visual field. Therefore, orientation of the ocular dominance stripes should not affect the length of intra-eye connection. Third, I neglected a possible specificity of inter-eye connections in respect to monocular functions of V1 such as orientational selectivity. Binocular neurons are likely to receive information from the neurons with the same preferred orientation. Moreover this preferred orientation should correlate with the disparity direction. However, this effect should not affect my argument because it averages out. Once all the possible orientations are included, the combined connections should be non-specific because different orientations are approximately equally represented. Fourth, I assumed retinotopic mapping in V1. Although there is scatter in the receptive field location in a given cortical column, the magnitude of the scatter does not exceed the period of the ocular dominance pattern. Because I rely on retinotopy on the scales of several stripe widths (Fig.3) the argument remains valid. Thus I showed that the orientation of ocular dominance stripes should follow the direction of disparity for the corresponding point of the visual field. To determine the optimal pattern of ocular dominance stripes I need to find disparity for all points in the visual field. Thus I need to calculate a binocular disparity map. ## III Calculation of the typical disparity map In the previous Section, I argued that the ocular dominance pattern should follow the map of binocular disparity in order to minimize the length of intra-cortical wiring. However the direction and magnitude of the disparity for a given point in the visual field depends on the viewing conditions such as the distance to the object and the direction of gaze. Therefore I need to average disparity over these variables. The typical direction of disparity should determine the optimal orientation of ocular dominance stripes. In order to find the disparity map I consider the origins of disparity in some detail. This is done in several steps by first considering a primate with the gaze direction fixed at straight ahead and then gradually adding degrees of freedom available to the eyeballs. Consider two eyes fixating at optical infinity. Then, by definition, the images of the fixation point fall on the foveae of the two eyes. Moreover, all objects at infinity are imaged on the retinal locations which are the same distance and direction from the fovea in both eyes. Such locations send afferents to adjacent cortical columns and are called corresponding. (In reality, images of infinite objects may not fall on exactly corresponding points. For example, there is a $`2\mathrm{deg}`$ tilt of the vertical meridians., In this paper I neglect these deviations because they do not alter the results qualitatively.) Physiologically, stimulation of corresponding points results in a single perception of the object. Objects at finite distance away, however, are imaged at different retinal locations relative to the fovea, called non-corresponding. Binocular disparity is defined as a displacement of the left-eye image from the location on the left retina corresponding to the right-eye image of the same point object. Physiologically, finite-distance objects may still appear single due to sensory fusion if the disparity falls within a range called Panum’s fusional area. Otherwise the doubling of the perception or diplopia is experienced. When the eyes fixate at optical infinity all the infinitely removed objects appear with zero disparity. To find disparity of other objects I use a geometrical construction illustrated in Fig.4. I fix the direction of gaze at straight ahead. The image of point $`O`$ in the right eye falls on the retinal point $`R`$ which belongs to the line passing through $`O`$ and the nodal point of the right eye $`N_R`$. The image of point $`O`$ in the left eye falls on the retinal point $`L^{}`$. A line passing through the left-eye nodal point $`N_L`$ and parallel to $`OR`$ intersects the retina at the point $`L`$, a point corresponding to $`R`$. The arc $`LL^{}`$ is the binocular disparity of point $`O`$. This arc belongs both to the retina and to a plane that passes through point $`O`$ and the nodal points of the two eyes, known as an epipolar plane. The common of the epipolar plane and the retina is called an epipolar line. Therefore the direction of disparity $`LL^{}`$ is along the epipolar line while its magnitude depends on the distance to point $`O`$. If point $`O`$ had a different elevation its disparity direction will be aligned with another epipolar line formed by the intersection of another epipolar plane and the retina. Therefore possible directions of disparity in the visual field are along epipolar lines formed by great circles passing through the interocular line. I call this disparity translational because it results from the horizontal displacement of the eyes. Now I allow eyes to change the direction of gaze in the horizontal plane, while assuming that the center of rotation of an eyeball coincides with its nodal point. Then point $`O`$ is projected onto the same locations $`L^{}`$ and $`R`$ in head-centered coordinates. Point $`L`$ remains corresponding to $`R`$. However the eyes, and hence the retinae rotate under those points. Therefore, the direction of disparity in the retinal coordinates changes depending on the gaze direction. I calculate the frequency distribution of disparity directions by averaging over a uniform distribution of gaze directions within $`\pm 30\mathrm{deg}`$ of straight ahead. The result is shown in Fig.5 as a grid of polar plots each of which corresponds to a particular point in the retinal coordinates. Each polar plot shows the frequency of different directions of disparity for a given point on the retina. The distribution of disparity exhibits strong anisotropy and the dominant disparity direction can be easily determined for all the retinal locations. If the only movements allowed to the eye were rotations around vertical axis then this would be a complete disparity map. Optimal ocular dominance pattern would be determined by transforming this map into cortical coordinates by using the magnification factor. This map agrees with the parafoveal stripe orientation in macaque but disagrees with it farther from the fovea. Inclusion of different gaze elevations eliminates this disagreement. Naively, one may expect that the disparity map remains intact because the interocular line is the axis of rotational symmetry. However, vertical eye movements are accompanied by cyclotorsion,, or rotation of the eyeballs around the direction of gaze. The amplitude of cyclotorsion depends on the direction of gaze as specified by Listing’s law. (Listing’s law states that to determine the amplitude of cyclotorsion for an arbitrary direction of gaze one has to rotate the eye into that direction from the primary position around an axis which lies in a (Listing) plane.) According to the recent measurements the amplitude of cyclotorsion is often unequal in the two eyes. Thus, although points $`L^{}`$ and $`R`$ remain fixed in the head-centered coordinates, point $`L`$ corresponding to $`R`$ rotates around the direction of gaze direction. This causes a cyclotorsional contribution to disparity which is oriented along concentric circles around the fovea. The full binocular disparity includes cyclotorsional and translational contributions. Because translational contribution depends on the distance to an object while cyclotorsional does not, the direction of disparity depends on the distance to an object. Therefore, finding the typical disparity direction requires specifying a range of distances to objects that are perceived stereoscopically. To define this range of distances I use the following argument. Because the cost of connections grows with their length, the pattern of connections in the cortex should not be far from isotropic. For a given location on the cortex, the circle of intra-cortical connections corresponds to an ellipse in the visual field because of the anisotropic magnification factor. The short axis of this ellipse $`a`$ is perpendicular to the direction corresponding to the stripes while the long axis $`b`$ is parallel. Information is recombined from all objects with disparity less than $`a`$ and is not recombined from any object with disparity greater than $`b`$. Whether information is recombined from objects with disparity in the interval $`[a;b]`$ depends on the orientation of ocular dominance stripes. In order to choose the optimal direction I need to calculate the frequencies of disparity directions for the object locations whose magnitude of disparity falls in the interval $`[a;b]`$. I choose the values of $`a`$ and $`b`$ as explained in Methods. Although, results of the calculation depend on the choice of parameters, they do not change qualitatively (see Discussion). The typical disparity map is shown in Fig.6. ## IV Discussion Orienting ocular dominance stripes in the direction locally corresponding to the typical disparity optimizes the length of intra-cortical wiring needed for the perception of depth. Therefore, the wiring economy principle predicts that the ocular dominance pattern follows the map of typical disparities. This prediction agrees with the data as can be seen by comparing the map of typical disparities in Fig.6 with the data of LeVay et al. (1985), Fig.2. The map of typical disparities reproduces correctly the two major trends in the data: in the parafoveal region stripes tend to run horizontally, farther from the fovea, the pattern consists of isoeccentric stripes. These trends result from the two major components of disparity: translational, due to the horizontal displacement of the eyes, and cyclotorsional, due to unequal rotation of the eyeballs around the gaze line (cyclotorsion). The relative magnitude of the two components depends on the distance to the fovea, Fig.7. In the parafoveal region, cyclotorsional disparity goes to zero linearly with eccentricity because rotational displacement is proportional to the radial distance from the axis of rotation passing through the fovea. At the same time, translational disparity remains finite for objects closer or farther than the point of fixation. Therefore, translational disparity dominates in the parafoveal region. Since the translational disparity is mostly horizontal, this explains the horizontal trend in the stripe orientation in parafoveal region. Farther from the fovea, cyclotorsional component of disparity may (or may not) become dominant depending on several parameters: the amplitude of cyclotorsion, the frequency of different viewing conditions, and the typical limit of depth perception. Along the horizontal meridian, cyclotorsional disparity is vertical, while translational is horizontal. Hence, the direction of the ocular dominance stripes must switch at the point where cyclotorsional component of disparity takes over the translational. This switch is evident in the macaque data, Fig.2 at about $`8\mathrm{deg}`$ eccentricity. These trends in the orientation of ocular dominance stripes are not qualitatively affected by assumptions made in the calculation. For example, I assumed a uniform distribution of the gaze directions. Any reasonable bell-shaped distribution should lead to the same two trends in the ocular dominance pattern. Although I considered fixation at infinity my results are qualitatively correct for fixation at nearby objects. A simple geometrical argument shows that the translational disparity has a greater vertical component for close distances. Also, cyclotorsional disparity is greater because cyclotorsion increases with the vergence angle.,, Qualitatively, the ocular dominance pattern displays the same two trends. However the switch in the stripe orientation may occur at different eccentricity. Generality of the trends in the stripe orientation confirms their functional significance. Ocular dominance patterns imaged in several macaques show the two trends in the stripe orientation. Rosa et al. (1992) transformed into visual field cooridinates a complete ocular dominance pattern of Cebus monkeys. They found that the pattern was qualitatively similar to macaque with the stripe orientation switching at $`6\mathrm{deg}`$ and the first trend sometimes lacking. Preliminary data on the ocular dominance pattern in humans is hard to analyze because precise topography in V1 is not known. Although the two trends are present, there is a significant interpersonal variability, possibly indicating varying significance of the two contributions to disparity from person to person. Although, this theory reproduces the two major trends in the data, there is an unexplained trend in the macaque and Cebus monkey data. In the foveal region, less than $`1\mathrm{deg}`$ eccentricity, the orientation of stripes differs from the typically horizontal disparity there. I speculate that this may be due to fixation point being projected away from the V1-V2 border. To verify this suggestion a combined topographic and ocular dominance mapping is needed. This theory relates ocular dominance pattern to the function of V1 allowing me to make several predictions. The location of the switch in the orientation of the ocular dominance stripes along the horizontal meridian should depend on the following parameters. Greater amplitude of cyclotorsion (or a greater frequency of gaze directions requiring cyclotorsion) increases cyclotorsional disparity and pushes the location of the switch in the stripe orientation towards the fovea. A greater extent of the Panum’s fusional area, $`b`$, achieves fusion for more objects with largely translational disparity. This should increase the eccentricity of the switch. Schwartz (1980) suggested that the size of the Panum’s fusional area and the width of the ocular dominance stripes are correlated between different species. If this is correct my theory implies that the species with greater width of ocular dominance stripes in the visual field should have the switch at higher eccentricity. According to the theory, the functional significance of the orientation of ocular dominance stripes is in accommodating the typical disparity direction. Then processing of binocular disparity for any point of the visual field should be more efficient in the direction corresponding to the stripe orientation at that point. This predicts a greater number of disparity selective neurons for the direction of disparity corresponding to the stripe orientation as can be verified electrophysiologically. Also the limits of depth perception (for example, Panum’s fusional area) should be greater in the direction along the stripes than across as can be tested psychophysically. In conclusion, I argued that the orientation of the ocular dominance stripes optimizes the length of intra-cortical wiring needed to process binocular disparity for depth perception. This argument supports the utility of the wiring economy principle as a powerful tool in relating organization of the cortex to its function. ## Methods I obtained the disparity maps numerically for a grid of points on the right retina using the following algorithm. I found points in the visual field that project onto a given point on the right retina and calculated the location of their image on the left retina. The difference between the images’ coordinates on the left and the right retinae is binocular disparity. Multiplying it by the magnification factor I found disparity in cortical coordinates. Including only the points in the visual field with cortical disparity magnitude in the interval $`[a;b]`$ I calculated the frequencies of disparity directions averaged over all possible directions gaze. I implemented the algorithm in MATLAB using the following parameters. The azimuth of the gaze directions as well as the elevation were uniformly distributed within $`30\mathrm{deg}`$ of straight ahead. The cyclotorsional misalignment was $`10\%`$ of the elevation angle. The point of fixation was at finite but large (2000 times interocular) distance. The expression for the cortical magnification factor was proportional to $`1/(e+e_2)`$ where $`e`$ is eccentricity and $`e2=2\mathrm{deg}`$. For the map shown in Fig.6 I chose $`a=0.5b`$ and $`b=0.13e`$ corresponding to the diameter of the Panum’s fusional area. ## Acknowledgments I thank M.R. DeWeese, A.M. Zador, J.D. Pettigrew, R.J. Krauzlis, J.C. Horton, B.G. Cumming, E.M. Callaway, T.D. Albright and, in particular, C.F. Stevens for helpful discussions. This research was supported by a Sloan Fellowship in Theoretical Neurobiology.
no-problem/9906/physics9906024.html
ar5iv
text
# 1 Introduction ## 1 Introduction Wave interpretation of quantum phenomena was started in 1924 by de Broglie hypotesis,which says that wavelength of any particle with P momentum is equal to $$\lambda =h/P$$ (1) where $`h`$ is Plank’s constant. This hypothesis was first proved by experiment conducted by K.Davisson and L.Djermer . In this experiment by observing reflection of electron beam from nickel single cristal they have obtained pattern absolutely corresponding to wave diffraction with $`\lambda `$, defined by (1) and satisfied by the condition $$bSin(\theta )=n\lambda $$ (2) where $`b`$ is cristal period,n is whole number,$`\theta `$ is observation angle of diffraction maxima. To develop a stable (observable) diffraction pattern in classic optics it is required to have simultaneous arrival to point of observation of 2 (interference) or more than 2 (diffraction) monocromatic and coherent waves. From the point of view of quantum mechanics, one will obtain interference of wave with itself if a barrier with 2 slits will be put on a way of single particle.By multiple repetition of the experiment full diffraction picture will be received on the screen.The probable nature of wave function F allows doing that (of course, on paper). Let’s assume that a microparticle passes through the first slit with probability $`\omega _1=|\psi _1|^2`$, and through the second slit with $`\omega _2=|\psi _2|^2`$.The superposition principle allows to write the following for the wave function $`\mathrm{\Psi }`$ $$\mathrm{\Psi }=c_1\psi _1+c_2\psi _2,$$ (3) where $`c_1`$ and $`c_2`$ are normalization coefficients. By choosing $`\psi _1=a_1Sin(\omega t\varphi _1)`$ and $`\psi _2=a_2Sin(\omega t\varphi _2)`$ we will have $$W=|\mathrm{\Psi }|^2=a_1^2+a_2^2+2a_1a_2Cos(\varphi _1\varphi _2),$$ (4) where $`W`$ is probability of particle to hit certain point of screen. One can see from (4) that depending on $`(\varphi _1\varphi _2)`$ the last (interference) term can be either positive or negative,i.e.in different points of the screen the hitting probability of particles will be different,and by long exposure of the screen by single particles the full diffraction picture can be received. Let’s see what logical contradictions will occur from superposition principle in interpretation of the diffraction pattern. No doubt that by passing a barrier with two slits the particle(wave) is not spliting into 2 parts, otherwise it would be visible by experiment.This means that each particle passes through one of these two slits(doesn‘t matter through wich one exactly).Then in expression (3) for wave function $`\mathrm{\Psi }`$ beyond the barrier either $`\psi _1=1`$, and consequently $`\psi _2=0`$ , or $`\psi _2=1`$, and $`\psi _1=0`$. In both cases the interference term disappears in (4).The interference maxima and minima observed in all experiments show that wave formalism introduced to explain diffraction pattern contradicts to experimental results.This conclusion becomes obvious if one looks at (3) from the point of view of simultaneous arrival of two waves to point of observation.To satisfy this condition , obligatory for the interference,the quantum mechanics divides tasitly the particle (wave) into 2 parts with the help of equition(3).Just in the same tacit way it is assumed that passing various paths these divided parts meet each other at a microscopic area $`\lambda ^2=10^{16}sm^2`$ ($`\lambda `$ is the de Broglie wavelengt equal to about $`1A^o`$ in the experiments and ) of the macroscopic screen surface $`10cm^2`$. In this work we show that the observed maxima and minima of the diffraction pattern can have other origin ,connected with the action discretness which served for the development of the quantum mechanics.In this way one can escape the wave presentation and the connected with them logical difficulties of explaining the interference of single particle(wave) with itself. To obtain an answer to this important question we propose a ,gedanken’ experiment considered by Feynman,which will give unambigeous answer on the validity of the wave presentations. The discussion of the corpuscular interpretation of the diffraction pattern is also given. ## 2 Two slits experiment According to the attempt to understand the phenomenon of single particle (wave) interference with itself is an unnecessary intellectual masochism. However,the absence of the unambigous test of the influence of the second slit through which the particle did not pass, makes such a statement groundless. The contemporary atomic interferometers allow one to realize such a test according to the folloving scheme:it is necessary to compare the interferometer pattern, obtained with the help of two slit interferometer , for instance, when both slits are open,with the summary pattern composed of two expositions when consequently one of the slits is closed. In the last case the interference term in (4) is absent and the maxima and minima must not be observed. If the summary pattern will differ from the one when both slits are open,then this will be anambigeous experimental confirmation of the fact that the open slit, through which the particle did not pass,is not equivalent to the close slit and it influences the interference pattern. And vice versa,if the summary pattern will be identical to the one when both slits are open,then this will mean that the wave interpretation of the appearance of maxima and minima is incorrect,and it is necessary to look for other explanation. ## 3 The corpuscular interpretation To overcome the above mentioned well known logical contradictions, connected with the interference of a single particle (wave) with itself,it is necessary to take into account the interaction of this particle with the slit matter (diffraction grating) as well as with subsidiary exiting electron and laser beams. At the first sight,the problem seems to have no solution. However, it can be solved using Planck’s constant $`h`$, common for all types of interactions quantum of action,together with the hypothesis that the action is multiple to $`h`$ for the unbound states,as for bound states. Let us follow how this can be done. Let a parallel beam of microparticles wit momentum $`P`$ moves along the direction $`x`$ and falls on a single cristal with period $`b`$ (distance between the cristallographic planes),or on a slit with width $`b`$. As a result of interaction with the cristal(slit) matter the particle gets a transverse momentum $`P_r`$ and will be scattered under an angle $`\theta `$, defined by the relation: $$Sin(\theta )=P_r/P$$ (5) It is necessary to find $`P_r`$ from the following differential equation $$dP_r=F_rdt$$ (6) where $`F_r`$ is the force acting on the particle by the cristal in a direction perpendicular to the beam direction. Multiplying (6) by $`dr`$ one obtains $$dP_rdr=F_rdrdt=dS_r$$ (7) where $`dS_r`$ is the action on the path $`dr`$ in the direction $`r`$ for the time $`dt`$. Using the hypothesis of multiplicity of action to $`h`$ and choosing the integration limits from 0 to $`t`$ for the time and from 0 to $`b`$ for $`r`$ one obtains $$bP_r=S_r=nh,$$ (8) or taking into account (5) $$bSin(\theta )=nh/P=n\lambda $$ (9) where $`\lambda `$ is the well known de Broglie wave length. Formula (9) shows that the agreement the experimental data with the de Broglie hypothesis is not accidental. In a hidden form the de Broglie hypothesis contains the hypothesis of the multiplicity of action to $`h`$ in an interaction used in this work. Just for this reason,as it is seen from (9) the diffractive scattering angles take discrete volues,imitating the diffraction pattern. ## 4 Conclusion To test the above proposed interpretation it is very important to perform the Feynmans,gedanken’ experiment with two slits completely with the help of atomic interferometers.
no-problem/9906/astro-ph9906178.html
ar5iv
text
# Long-Term X-ray Monitoring of 1E 1740.7–2942 and GRS 1758–258 ## 1 Introduction 1E 1740.7–2942 and GRS 1758–258 are by far the brightest persistent sources in the Galactic bulge above $``$50 keV (Sunyaev et al. 1991). Their spectra are typical of a black hole low (hard) state (Heindl et al. 1993; Sunyaev et al. 1991). Although variable over times of days to years, they spend most of their time near their brightest observed level. Both have a core-and-jet structure in the radio (Heindl, Prince, & Grunsfeld 1994; Mirabel et al. 1992; Rodriguez, Mirabel, & Martí 1992) and have therefore been described as microquasars. These characteristics make this pair of objects a subclass among the black hole candidates. This subclass shares features with other black hole candidates. Radio jets also appear in the much brighter, and spectacularly variable objects more usually called microquasars: GRS 1915+105 and GRO J1655-40 (Greiner, Morgan & Remillard 1996; Zhang et al. 1997a), whose jets, too, are brighter and more variable. Maximum luminosities around $`3\times 10^{37}`$ ergs sec<sup>-1</sup> are shared with Cyg X-1 and the recently discovered transient GRS 1737-31 (Cui et al. 1997). The property of being in the hard state at fairly high luminosities half the time or more is shared only with Cyg X-1. The property of having been observed only in the hard state is shared with GRS 1737-31, GS 2023+338, GRO J0422+32, and GRO J1719-24, although the total amount of time devoted to these objects varies widely (Zhang, Cui, & Chen 1997; Tanaka & Lewin 1995). Despite the hard spectra of the two objects, there has been some preliminary evidence of state changes: changes in the spectral shape of 1E 1740.7–2942 above 20 keV from BATSE data (Zhang et al. 1997b), and the detection of weak soft components from GRS 1758–258 (Mereghetti, Belloni, & Goldwurm 1994; Heindl & Smith 1998; Lin et al. 1999) and, with marginal significance, from 1E 1740.7–2942 (Heindl & Smith 1998). Both sources were observed by SIGMA and ART-P to vary between observations separated by 6 months from a hard X-ray flux of about 130 mcrab (40 mcrab in the 8-20 keV ART-P band) to a level less than 10 mcrab and consistent with zero (Churazov et al. 1994; Pavlinsky et al. 1994). BATSE has also observed both 1E 1740.7–2942 and GRS 1758–258 at a near-zero flux level (Zhang et al. 1997b). Day-to-day variability is also seen in these data, including a 1 day jump in ART-P flux from 1E 1740.7–2942 from $``$3 to $``$18 mcrab. Both sources show rapid variability with a flat-topped power spectrum, behavior typical of the hard states of both black holes and neutron stars (Smith et al. 1997). 1E 1740.7–2942 and GRS 1758–258 have high Galactic extinction in the optical, and counterparts have not been identified; only O stars and red supergiants have been ruled out as companions (Chen, Gehrels, & Leventhal 1994). There have therefore been no mass determinations; it has even been suggested (Bally & Leventhal 1991) that 1E 1740.7–2942 does not need a companion and could be accreting directly from a nearby molecular cloud. However, it has also been suggested that the lack of a 6.4 keV emission line in the spectrum of 1E 1740.7–2942 strongly constrains the amount of gas immediately surrounding the source (Churazov, Gilfanov, & Sunyaev 1996). ## 2 Observations We have observed 1E 1740.7–2942 and GRS 1758–258 in $``$1500 second intervals with the Rossi X-Ray Timing Explorer (RXTE). From 1996 February through 1996 October the observations were spaced one month apart. We have observed weekly since 1996 November. This report is based on data obtained through 1998 September. Because RXTE cannot point near the Sun, from late November to late January observations were not taken. All data reported here were taken with the Proportional Counter Array (PCA). The PCA (Jahoda et al. 1996) consists of five xenon proportional counters of $``$1300 cm<sup>2</sup> each, for a total of 6500 cm<sup>2</sup>, that are sensitive from 2 to 60 keV and share a 1 FWHM passively collimated field of view. We calculate instrumental background with the standard “Q6” model for consistency in a data set that spans the whole mission. Although this is not the most current model, we have found the differences among models to be negligible for these moderately bright sources. Our response matrices are those standard to FTOOLS release 4.1, including the time dependences of the gain and of the diffusion of xenon into the propane layer. The pointing directions were offset to avoid other nearby X-ray sources: A1742-294 and other Galactic Center sources near 1E 1740.7–2942 and GX 5-1 near GRS 1758–258. The instrumental effective areas resulting from the offsets were 43% and 46% of on-axis values for 1E 1740.7–2942 and GRS 1758–258, respectively. Since both sources lie in the Galactic plane, background observations were made to determine the Galactic diffuse emission. For 1E 1740.7–2942, the background field is opposite in Galactic longitude and equal in Galactic latitude to the source field. For GRS 1758–258, the background fields are at the same Galactic latitude as the source field and on either side in Galactic longitude. The source and background fields are shown along with some bright sources in the region in Figure 1. The coordinates of these pointings are given in Smith et al. (1997). For the spectral analyses, we used only the top layer of the PCA detectors, in the energy range 2.5 to 25 keV. In this mode, the diffuse X-ray background from the Galactic plane is 77 counts sec<sup>-1</sup> for 1E 1740.7–2942 and 32 counts sec<sup>-1</sup> for GRS 1758–258. The instrumental background is about 20 counts sec<sup>-1</sup>. Typical source count rates for both sources are about 100 counts sec<sup>-1</sup>. Although 1E 1740.7–2942 is somewhat brighter than GRS 1758–258 in the 2.5 to 25 keV band, it is also more absorbed below a few keV. All the background-subtracted energy spectra were fitted with an absorbed power law. The time histories of the PCA count rate, the rms variability, and the photon power-law index (PLI) are shown in Figure 2. Gaps from November to January of each year are due to the RXTE solar pointing constraint. A few observations of each source have been removed due to very high background when the observation was made immediately after exiting the South Atlantic Anomaly. This condition occurred more often for 1E 1740.7–2942. After removing one observation of 1E 1740.7–2942 which was contaminated by the new transient XTE J1739-302 (Smith et al. 1998), we present a total of 77 observations of 1E 1740.7–2942 and 82 observations of GRS 1758–258. ## 3 Subtle Changes and Hysteresis The histories of both 1E 1740.7–2942 and GRS 1758–258 in Figure 2 clearly show that neither source turned off during the past 3 years. The count rate has ranged from about 60 counts sec<sup>-1</sup> to 140 counts sec<sup>-1</sup> in both sources. This result is in conflict with a recent preliminary report on GRS 1758–258 by Cocchi et al. (1999) using data from the BeppoSAX Wide Field Camera. They report that on three occasions (1996 September, 1997 October, and 1998 March) the flux dropped by roughly a factor of 5, becoming so low as to be undetectable. No such large drops appear in the RXTE data at these times, which are marked by triangles in Figure 2. Although neither the BeppoSAX nor the RXTE data are continuous, the RXTE dataset has 59 short pointings during the range of time covered by the 16 short BeppoSAX observations in Cocchi et al. (1999). It is therefore highly unlikely that RXTE would miss the large variations reported by BeppoSAX if they occurred with a random distribution in time. The RXTE spectra show that both sources have remained in the hard state during the past 3 years, even though the PLI has occasionally softened slightly. All the variations discussed below are subtle changes within the hard state. Both sources exhibit events of brightening and softening in early 1998 (see Figure 2). The softening clearly lags the brightening. By using the cross-correlation function, we found that there is a $``$58 day lag between PLI and count rate in 1E 1740.7–2942 and a $``$36 day lag in GRS 1758–258. In both sources, we used only the data around the peaks in the count rate and PLI for computing the cross-correlation function. The peaks occurred between 22 January 1998 and 11 September 1998. The brightest periods are approximately from 2 February 1998 to 28 May 1998 for 1E 1740.7–2942 and 22 January 1998 to 12 March 1998 for GRS 1758–258. Figure 3 shows scatter plots of PLI vs. count rate for both GRS 1758–258 and 1E 1740.7–2942 during these events. When the points on the scatter plots are connected, a circle is clearly evident, showing hysteresis between the two parameters; i.e., the time lags described above can also be thought of as a phase lag of $`90^{}`$. Because of this hysteresis effect, our data show that the time lag could hide real correlations in scatter plots using data taken over long periods. We have considered the possibility that these events were instrumental. This is unlikely for two reasons: 1) the events are not simultaneous and 2) they are much greater changes than the known time evolution of the instrument parameters, such as efficiency and gain. The brightening/softening event in 1E 1740.7–2942 was preceded by a period in which the PLI was gradually hardening, beginning in March 1997 and lasting for $``$250 days. In GRS 1758–258 there was also a period of gradual hardening before the similar event, lasting $``$150 days. ## 4 Timing Results and Analysis The counts for each observation were summed into 31.25 ms bins for the timing analysis. The individual observation times ranged from 1000 to 1500 seconds. Typical power spectra for GRS 1758–258 and 1E 1740.7–2942 are shown in Figure 4. The full energy range of the PCA (2-60 keV) was used for all the power spectra, but the contribution above 20 keV is small. We fitted the power spectra with a broken power law (index 0 below the break and free above it). Typical values of the break frequency range from 0.1 to 0.8 Hz for both GRS 1758–258 and 1E 1740.7–2942. Typical values of the index above the break frequency are around -1. The rms variability integrated from 0.004 to 15.8 Hz ranged from 21.5% to 30.5% for GRS 1758–258 and from 13.9% to 28.6% for 1E 1740.7–2942. In these short observations, the statistics were insufficient to observe any quasi-periodic oscillations (QPOs). However, in longer duration observations, QPOs have been observed in both GRS 1758–258 and 1E 1740.7–2942 (Smith et al. 1997). Figure 5 shows the average power spectrum of 77 observations for 1E 1740.7–2942 and 82 for GRS 1758–258, and Figure 6 shows the results from Smith et al (1997). There are no obvious QPOs in our results. Since the QPOs exist in the long observations but do not appear when the short observations are summed, one may conclude that either the deep observations found a rare appearance of the QPOs, or that the QPOs drift in frequency with time. Wijnands & van der Klis (1998) have shown a correlation between the break frequency and QPO frequency in both neutron stars and black hole candidates. Since the break frequencies of GRS 1758–258 and 1E 1740.7–2942 are variable, then the QPO frequencies may also be variable. ## 5 Energy Spectra We fitted the energy spectra of GRS 1758–258 and 1E 1740.7–2942 with a power law absorbed by a column of neutral interstellar gas. The variations in PLI were shown in Figure 2. For 1E 1740.7–2942, the column depth varied between $`7.4\times 10^{22}`$ and $`11\times 10^{22}`$ atoms cm<sup>-2</sup> and the average value was $`9.2\times 10^{22}`$ atoms cm<sup>-2</sup>. For GRS 1758–258, the absorption column varied between $`0.71\times 10^{22}`$ and $`2.3\times 10^{22}`$ atoms cm<sup>-2</sup> with an average value of $`1.4\times 10^{22}`$ atoms cm<sup>-2</sup>. Sheth et al. (1996), measured the column depth for 1E 1740.7–2942 at $`(8.1\pm 0.1)\times 10^{22}`$ atoms cm<sup>-2</sup> with ASCA, which is consistent with the range we obtained. Another ASCA measurement (Mereghetti et al. 1997) measured the column depth for GRS 1758–258 at $`(1.5\pm 0.1)\times 10^{22}`$ atoms cm<sup>-2</sup>, again consistent with our range. The column depth is mostly determined by the spectral shape below $``$4 keV, and the PLI mostly by the higher-energy part of the spectrum. At the lowest energies, we are most vulnerable to uncertainties about the influence of diffuse emission and soft sources on the edges of both fields of view (see §2). We therefore do not claim that the range of absorption columns obtained is evidence of real variability. There is no correlation between the column depth and PLI in either source, and we are confident that the variations in PLI are real. The PLI ranged from 1.37 to 1.76 for 1E 1740.7–2942 and from 1.45 to 1.86 for GRS 1758–258. The values found by Heindl and Smith (1998) for the PLI of GRS 1758–258 and 1E 1740.7–2942, using their deep pointings to both sources in August and March of 1996, were 1.54 and 1.53, respectively. These values included HEXTE data and were derived with a model which included an exponential cutoff at high energies. With the simpler power-law model used here, the indices from the monitoring observations just before and after each deep pointing average to 1.67 and 1.68 for GRS 1758–258 and 1E 1740.7–2942, respectively. This is consistent with the expectation that, in the absence of a cutoff in the model, the effect of the cutoff will appear in a softening of the fitted PLI. The statistics in individual monitoring observations are not good enough to measure the cutoff and PLI independently. ## 6 Comparison to Cygnus X-1 The similarities between Cyg X-1 and 1E 1740.7–2942 and GRS 1758–258 suggest that these three sources are similar objects. Some of these similarities were mentioned in §1: the x-ray luminosities, hard spectra, and persistent activity of all three sources. Another similarity is the shape of the power spectra. When in the hard state, all three sources show white noise up to $``$.5 Hz (see §4) and break to a power law, with an index above the break of $``$ -1. Cyg X-1 displays a relationship between the break frequency and the rms variability integrated over frequency. This behavior was first illustrated by Belloni and Hasinger (1990). They showed that the low-state power spectrum for Cyg X-1 always had the same normalization above the break frequency, which varied. Miyamoto et al. (1992) noted that this held true from one black hole candidate to another. We searched for the same effect in GRS 1758–258 and 1E 1740.7–2942. We divided our data into three groups, those with the largest, smallest, and near-average rms and averaged the power spectra in each group. Because of this averaging, the break frequencies are more rounded than Belloni & Hasinger showed for Cyg X-1, but otherwise Figure 7 shows that GRS 1758–258 and 1E 1740.7–2942 are similar to Cyg X-1 in this respect. Unlike 1E 1740.7–2942 and GRS 1758–258, Cyg X-1 has been observed in a true soft state, in which a soft thermal component was dominant (e.g. Cui et. al. 1997b). When Cyg X-1 was in the soft state, the PLI was -2.2. A similar index was seen by BATSE above 20 keV in 1E 1740.7–2942 while that source was faint in the BATSE band (Zhang et al. 1997b). Simultaneous observations at lower energies during another occurrence of this state are needed to confirm that it is a soft state similar to that in Cyg X-1 and other black-hole candidates. ## 7 Discussion ### 7.1 Dynamical model for hysteresis We can qualitatively explain the hysteresis or time lag between brightening and softening in 1E 1740.7-2942 and GRS 1758-258 in the context of some recent models of black-hole accretion. These models (e.g. Chakrabarti & Titarchuk 1995, Esin et al. 1998) have two components in the outer regions of the flow: a standard Keplerian disk, physically thin and optically thick, and an optically thin, physically thick halo or corona. The mass in the halo is nearly in radial free-fall, and it advects most of its accretion energy into the black hole rather than radiating it as the Keplerian disk does (Ichimaru 1977; Rees et al. 1982; Narayan & Yi 1995; Abramowicz et al. 1995). In early disk-plus-corona models, the corona was produced locally by the Keplerian disk, and did not accrete independently and advectively (e.g. Liang & Price 1977, Bisnovatyi-Kogan & Blinnikov 1977). In the newer models, it is an equally valid and independent solution of the hydrodynamic equations. Within a certain radius, the Keplerian disk is unstable, and only a very hot solution remains (an advection-dominated flow in the model of Esin et al. 1998, and a shocked flow in the model of Chakrabarti & Titarchuk (1995)). The soft, thermal component of black-hole-candidate spectra is attributed to the inner part of the Keplerian disk, near this boundary. The hard, power-law component is attributed to inverse Comptonization of these soft photons in the very hot inner parts of the advective flow. While matter in the advective flow is nearly in free-fall, matter in the thin disk only accretes after a gradual loss of angular momentum via viscous torques. The timescale for this process is approximately (Frank, King & Raine 1992) $$t_{\mathrm{visc}}\mathrm{\hspace{0.33em}3}\times 10^5\alpha ^{4/5}\left(\frac{\dot{M}}{10^{16}\mathrm{g}/\mathrm{s}}\right)^{3/10}\left(\frac{M}{M_{\mathrm{}}}\right)^{1/4}\left(\frac{R}{10^{10}\mathrm{cm}}\right)^{5/4}\mathrm{s},$$ (1) where $`M`$ is the black-hole mass, $`\dot{M}`$ the accretion rate, $`R`$ the disk radius, and $`\alpha `$ the viscosity parameter $`(0<\alpha 1)`$. Chakrabarti & Titarchuk (1995) pointed out that if the mass accretion rate were increased at the outer edge of both flows simultaneously, it would arrive at the central regions of the advective flow first. Thus the hard component would brighten first, with the soft component only brightening after a delay approximately equal to the viscous time. We offer this delay as one explanation for the hysteresis we observe between brightening and softening. If we use $`M=10M_{\mathrm{}}`$, $`\dot{M}=10^{17}`$g/s, $`t_{\mathrm{visc}}`$ equal to the measured delays (see §3), and $`\alpha `$ = 0.3 (Esin et al. 1998) to solve equation (1) for $`R`$, we find $`R=5\times 10^{10}`$ cm ($`3\times 10^4GM/c^2`$) for 1E 1740.7–2942 and $`R=3\times 10^{10}`$ cm ($`2\times 10^4GM/c^2`$) for GRS 1758–258. These disk sizes are typical of low-mass x-ray binaries accreting by Roche-lobe overflow, and are larger than the disks expected in systems accreting winds from massive companions (Frank et al. 1992). ### 7.2 Disk-evaporation model for hysteresis An alternative explanation for the lag between brightening and softening is quasi-static rather than dynamic: the flows can be allowed to reach an equilibrium configuration after every infinitesimal increase in accretion rate. This explanation relies on a characteristic common to the models of Esin et al. (1998) and Chakrabarti & Titarchuk (1995): as the accretion rate rises, the inner edge of the Keplerian disk moves inwards. If this edge were sufficiently far out to begin with, the spectral changes due to its inward advance would at first be restricted to the EUV and soft x-ray ranges, which are not observable for sources deep in the Galactic bulge. The only effect on the hard x-rays of increasing the accretion rate would be a brightening. Eventually, the disk would move in so far that it would begin to replace the hard-x-ray-emitting region of the advective or shocked flow, resulting in a spectral softening. If the response to the reduction in accretion rate back to the normal level were equally quasi-static, one would expect to see a bright and hard phase on the decline, i.e. the peaks in Figure 2 would be symmetric in time. However, it may be that the thin disk, once established at smaller radii, takes a significant amount of time to evaporate. It has been noted that there is hysteresis in the hard/soft/hard transitions of soft x-ray transients (Miyamoto et al. 1995). In a typical outburst of this class of black-hole candidate, the system remains in the hard state as the luminosity rises quickly from quiescence to near maximum, then switches to the soft state, then returns to the hard state only when the luminosity is of order 1% of maximum. The quick rise in accretion rate and quick transition to the soft state are thought to be due to the rapid propagation of a thermal-ionization instability in the disk (Cannizzo et al. 1985). This mechanism is not relevant to 1E 1740.7–2942 and GRS 1758–258, since their usual accretion rates are high enough that the disk would remain ionized by the x-rays from the central regions of the accretion flow. The return of the transients to the hard state, however, may be relevant to our observations. Mineshige (1996) interpreted this return as a transition from a Keplerian disk to an advective flow. Both the disk and advective flows are stable over most of the accretion rates traversed during the decline, but the transition doesn’t take place until the disk solution becomes unstable at very low accretion rates. In other words, the disk, once it is established, tends to persist in regimes where both solutions are stable. The typical luminosity of both 1E 1740.7–2942 and GRS 1758–258 is $`2\times 10^{37}`$ erg s<sup>-1</sup> from 1-200 keV (Heindl & Smith 1998). Our data never deviate by more than about 50% from this value. This is roughly 1-7% of Eddington luminosity for black holes of 3-20 $`M_{\mathrm{}}`$, and is orders of magnitude higher than the luminosity where the idealized transient of Mineshige (1996) is forced to return to the hard state. Nonetheless, if the added regions of inner disk in the transients persist for a month or more at accretion rates where the advective flow would also be stable, then the more modest inward extensions of the disk which occur when 1E 1740.7–2942 and GRS 1758–258 brighten might persist as long. The time asymmetry in our data, in this interpretation, would be due to the evaporation time of the inner disk being longer than the time in which the accretion rate returns to normal. ## 8 Conclusions We have presented the most detailed long-term coverage of these black hole candidates to date. In this 3 year period, we have never seen either source at a flux level less than half its maximum. Although neither source has entered the soft (high) state in this time, we have seen variations in spectral index within the range usually associated with the hard or low state (photon PLI $`<2.0`$). There is hysteresis when GRS 1758–258 and 1E 1740.7–2942 brighten and soften within the hard state, with the softening lagging the brightening by 1-2 months. This hysteresis could be due to the different propagation times in a disk and halo of an increase in $`\dot{M}`$ (§7.1), or else to a persistence of the thin disk after $`\dot{M}`$ returns to normal (§7.2). If the former is the correct interpretation, the lag time implies accretion disks of the size usually associated with accretion from a low-mass companion overflowing its Roche lobe. We find that the weekly observations of GRS 1758–258 and 1E 1740.7–2942 reveal no QPOs when summed up over many weeks. This leads us to the conclusion that either our deep observations observed rare appearances of the QPOs, or, more likely, that they drift in frequency with time, consistent with the behavior described by Wijnands & van der Klis (1998) for other sources. Both objects show the same relationship between the break frequency of the power spectrum and the total rms variability as other hard-state sources. We would like to thank Ann Esin, Lars Bildsten, and Lev Titarchuk for useful discussions on the interpretation of the hysteresis results. This work was supported by NASA grants NAG5-4110, NAG5-7522, and NAG5-7265.
no-problem/9906/chao-dyn9906020.html
ar5iv
text
# The Bispectral Aliasing Test: A Clarification and Some Key Examples ## 1 The Bispectral Aliasing Test The domain of the discrete-time bispectrum is the two dimensional bifrequency $`\{\omega _1,\omega _2\}`$ plane. Assuming a real-valued discrete time series, the usual replication phenomenon dictates that all non-redundant information is confined to the square $`0\omega _1,\omega _2\pi `$. When one fully accounts for symmetries, the non-redundant information in the bispectrum is confined to a particular triangle inside this square . This triangle naturally divides into two pieces. One piece is an isosceles triangle and is unproblematic. The other piece, somewhat unusual in shape, is the source of the controversy under discussion. Naive consideration of this triangle shows that it involves frequencies higher than the Nyquist frequency and therefore must have something to do with aliasing. Hinich and Wolinsky considered this more carefully and showed that the naive intuition is correct: if the discrete time series arises from sampling a stationary, band-limited, continuous-time process, and if the sampling rate is sufficiently rapid to avoid aliasing, then the discrete bispectrum is non-zero only in the isosceles triangular subset of the fundamental domain. Conversely, if the bispectrum of a sampled stationary continuous-time process is non-zero in the outer triangle, then the sampling rate was too slow to avoid aliasing. It should be clearly understood that there is no assertion that aliasing in general can be detected. The statement is not “if a signal is aliased, then the outer triangle will have a non-zero bispectrum.” Rather, the assertion is the converse, “if the outer-triangle shows a non-zero bispectrum, the (underlying) continuous-time signal must have been aliased.” At one level, this result is obvious and, in fact, the result was initially so-regarded . However, doubt soon arose. Perhaps the most important source for suspicion is the argument based on reconstruction alluded to above. In light of this objection, one is led to reconsider the association of the outer triangle with aliasing. One can take the position that there is no relation, as in . One can decide that something is aliased, but that it is the bispectral estimator rather than the signal. There is some plausibility to this claim, for the frequencies that are involved in the outer triangle are $`\omega _1,\omega _2,`$ and $`\omega _1+\omega _22\pi `$. This seems to be the position of Pflug et al. . Or, one can try to delineate the conditions, if any, under which the test makes sense. This was done by Hinich and Messer in 1995. They confirmed the validity of the original argument and stated its conclusions more carefully. In particular they conclude that a non-zero bispectrum in the outer triangle indicates a non-random signal or one of the following: * a random, but non-stationary signal ; * a random, stationary, but aliased signal, or; * a random, stationary, properly-sampled signal which violates the mixing condition. We believe that the analysis of Hinich and Messer, while entirely correct, did little to persuade the detractors of the test. In particular their analysis did not address the reconstruction objection and may have left the impression that the circumstances for which the test applies are unlikely to be met in practice. In this paper, we show that the reconstruction objection is far from fatal. We further establish that stationarity is the only property which is crucial to the test. Since this property is required in order to define the bispectrum, one can legitimately apply the aliasing test whenever one is entitled to compute a bispectrum. Therefore the bispectral aliasing test is as theoretically sound as the bispectrum itself. ## 2 The Selection Rule and Brillinger’s Formula The bispectrum, defined to be the triple Fourier transform of the third-order autocorrelation, reduces to a function of two frequencies since stationarity confines the spectrum to the plane through the origin of the frequency domain perpendicular to the vector (1,1,1). $$𝔉_{123}(c_3(t_1,t_2,t_3))=b(\omega _1,\omega _2)\delta (\omega _1+\omega _2+\omega _3)$$ (1) Another way of computing the bispectrum is to switch the order in which one does the Fourier transforming and the ensemble averaging. This leads to the following result. $$b(\omega _1,\omega _2)=X(\omega _1)X(\omega _2)X(\omega _3=\omega _1+\omega _2)$$ (2) If the process is bandlimited and $`X(\omega )=0`$ for $`|\omega |>\pi `$, then the bispectrum is confined to the intersection of the $`(1,1,1)`$ plane and the $`\pi `$-cube (i.e. $`(\omega _1,\omega _2,\omega _3)(\pi ,\pi )(\pi ,\pi )(\pi ,\pi )`$ ). The plane and its projection onto the $`(\omega _1,\omega _2)`$ plane is shown in Figure 1. Upon sampling with unit time step, one obtains the usual replication in three dimensions. (Doing everything in 3-dimensions and projecting at the end keeps things simpler and makes it easier to avoid errors.) In particular, one gets that if the process is sampled at a frequency greater than twice the highest frequency component, then the bispectrum is confined to the replications of the tilted hexagon shown. The replication gives the discrete-time bispectrum $`b_d`$: $$\begin{array}{c}\hfill b_d(\lambda _1,\lambda _2,\lambda _3)=\underset{\omega _1+\omega _2+\omega _3=0}{}b(\omega _1,\omega _2,\omega _3).\end{array}$$ (3) where $`\omega _i=\lambda _i+2\pi k`$ for integer $`k`$. Since the replication does not cause any overlaps, the outer triangle remains empty. This is the Hinich and Wolinsky aliasing theorem. (Note that the outer triangle is equivalent to the bigger triangle with vertices $`(0,\pi ,0),`$ $`(0,0,\pi )`$ and $`(0,\pi ,\pi )`$ by symmetries. See for details.) ## 3 The Reconstruction Objection Suppose we have a stationary process $`x(t)`$ and we undersample it by sampling at $`tZ`$. Then by convolving $`x(t)`$ with the appropriate $`sinc`$ function we get a reconstructed process $`x_r(t)`$. This new process will have exactly the same samples as the original process and therefore exactly the same sampled bispectrum: yet it is not aliased. Therefore for any process that is undersampled, we have another process producing an identical sampled process which is not undersampled, showing that that one could not possibly detect aliasing via the bispectrum computed from samples! The rub here is the fact that the reconstructed signal will not necessarily be stationary. Processes reconstructed from aliased samples of continuous-time signals are generally cyclostationary but not stationary. Some aliased processes do, in fact, reconstruct into stationary processes. But in the class of stationary signals for which the bispectral aliasing test gives positive results, reconstruction from aliased samples produces non-stationary processes. To carefully illustrate this we will consider several stationary processes generated by taking a periodic signal with period T and giving it a random shift $`\theta `$ \[0,T). First consider a simple cosine process, $$x(t)=\mathrm{cos}(\alpha \pi t+2\pi \theta /T),$$ (4) where $`\alpha `$ = 1.5, T = 4/3, and $`\theta `$ is randomly chosen from \[0,4/3). Upon sampling and reconstruction we get the cosine process given by $$x_r(t)=\mathrm{cos}(\widehat{\alpha }\pi t+2\pi \theta /T),$$ (5) where $`\widehat{\alpha }=0.5`$. The key idea is that the signal appears at the lower frequency as dictated by its replication into the fundamental region of the frequency space, (-$`\pi ,\pi `$), but its reconstructed phase is the same as the “source” component phase. Now consider $$x(t)=\mathrm{cos}(\alpha \pi t+2\pi \alpha \theta )+\mathrm{cos}(\beta \pi t+2\pi \beta \theta ),$$ (6) where $`\alpha =1.0`$, $`\beta =3.0`$, and $`\theta `$ is chosen randomly from the interval $`[0,\alpha ^1)`$. The reconstructed process one gets is $$x(t)=\mathrm{cos}(\alpha \pi t+2\pi \alpha \theta )+\mathrm{cos}(\widehat{\beta }\pi t+2\pi \beta \theta ),$$ (7) (where $`\widehat{\alpha }`$ and $`\widehat{\beta }`$ are the aliased frequencies). Although the phase terms $`2\pi \alpha \theta `$ and $`2\pi \beta \theta `$ are still uniformly distributed over $`2\pi `$ they now correspond to different time shifts so we no longer have a single shifted waveform. This process can easily be seen to be cyclostationary, but not stationary. See Figure 2 . Computation of the required expectations requires that one can “average over the ensemble.” Since this stationary process is not ergodic, one can not get the result from a single realization of the process. It is at this point that some differences in perspective arise. Strictly speaking, in order to compute a bispectrum one must perform the ensemble average. A single realization does not suffice unless the process is ergodic. Finally, consider the process defined by $$\begin{array}{cc}\hfill x(t)=& \mathrm{cos}((10/20)\pi t+52\pi \theta )+\hfill \\ & \mathrm{cos}((12/20)\pi t+62\pi \theta )+\hfill \\ & \mathrm{cos}((22/20)\pi t+112\pi \theta )\hfill \end{array}$$ (8) where $`\theta `$ is chosen randomly from \[0,1). Because the phases of these components are in a fixed relation, this process has a spike in the bispectrum at $`\omega _1=10/20\pi `$, $`\omega _2=12/20\pi `$, i.e., in the outer triangle. The reconstructed signal is given by $$\begin{array}{cc}\hfill x(t)=& \mathrm{cos}((10/20)\pi t+52\pi \theta )+\hfill \\ & \mathrm{cos}((12/20)\pi t+62\pi \theta )+\hfill \\ & \mathrm{cos}((18/20)\pi t+112\pi \theta )\hfill \end{array}$$ (9) which is not stationary. Therefore we have a signal with nonempty outer triangle whose reconstruction is not stationary. This situation is exactly what the bispectral test implies happens whenever the outer triangle is nonempty. The loss of stationarity causes the Fourier transform of the triple autocorrelation to “move off” of the (1,1,1) plane. Therefore, if one knows (or is willing to assume) that the process which generated the observed samples was stationary, one can rule out the unaliased reconstruction as the source of the samples. In a sense, the continuous time signal reconstructed from aliased samples of an original time series is a “measure zero” object. This result is very surprising to most people’s intuitions. It is studied further in Vixie, Sigeti and Wolinsky . ## 4 The Replication objection Upon looking at Equation 2 one may observe that even if X($`\omega `$) = 0 for $`|\omega |>\pi `$, sampling effectively fills in the spectrum at higher frequencies. This is the basis for the objection appearing in Swami . This concern is addressed as follows. While the spectrum does indeed fill out upon sampling, the undesired expectations remain zero. Consider a (statistically stationary) ensemble constructed by uniformly translating a periodic or finite-duration waveform $`x(t)`$. Two operations are necessary to produce the discrete-time ensemble; a uniform shift in time over a period $`T`$, which introduces linear phase factors, and sampling, which produces spectrum replication. These operations do not commute: i.e., one wants to time-shift the waveform first and then sample, rather than to shift its samples. For the shifted samples $`x_s(t+\theta )`$ one finds $$𝔉(x_s(t+\theta ))=e^{i\theta \omega }𝔉(x_s(t))$$ (10) But for the sampled shifted waveforms $`x(t+\theta )_s`$, i.e., the waveforms needed to construct a stationary ensemble, the phase of the original signal is propagated to higher frequencies periodically rather than linearly. This difference leads to the vanishing of unwanted expectations. For example, consider the process given by the randomly shifted sum of unit amplitude cosine waves with frequencies at $`n/20`$ (rad/s) where $`n`$ takes integer values from $`1`$ to $`19`$. The sampled spectrum has components at $`\omega _1=10\pi /20,`$ $`\omega _2=11\pi /20`$ and $`\omega _3=21\pi /20`$ but the average $$X(\omega _1)X(\omega _2)X(\omega _3)$$ (11) reduces to $$e^{i10\theta \pi }e^{i11\theta \pi }e^{i19\theta \pi }_\theta $$ (12) where $`\theta `$ is chosen with uniform probability from \[0,1). This average vanishes. Therefore, the potential contribution in the outer triangle is zero because averaging kills it. This is in contrast to the case where the average is zero because the spectral amplitudes are themselves zero (as in the proof of the aliasing test). ## 5 Empirical counter-examples Other objections to the test have been made. Frequently these objections involve a (purported) counter-example to the bispectral aliasing test. A particularly clear example is provided by Frazer, Reilly and Boashash . Here the authors do two things. They present an example of an aliased signal which the aliasing test fails to mark as aliased. The example is unproblematic: neither the aliasing test nor any aliasing test we are aware of will detect all aliased signals. It is not, however, a counter-example to the test. Since there is nothing in the outer triangle, the bispectral aliasing test makes no assertion regarding the presence of aliasing. The other example the authors provide is more interesting. It consists of a signal involving coupled sinusoids at $`\omega _1=0.3125Hz`$, $`\omega _2=0.25Hz`$ and $`\omega _3=.4375Hz`$ and the authors show that there is a peak in the outer triangle under conditions which rule out aliasing. As the authors note these frequencies sum to 1 Hz ((the sampling rate). Under these conditions the authors are correct in asserting that the aliasing test gives a positive result, which they believe to be incorrect. However, what the aliasing test actually indicates is that this signal is non-stationary. The particular interaction which the authors have constructed is not one for which the continuous-time selection criteria is met, i.e., the frequencies involved do not sum to zero. Samples of this signal do meet the discrete-time stationarity condition and so a non-zero bispectrum is possible in the outer triangle. One can look at these results in various ways. Our position is that neither example constitutes a counter-example to the validity of the aliasing test in theory, though they both show that the test is limited in practice. The first example shows that there are aliased signals which the test does not see. This is obvious anyway since there are signals with zero bispectrum whose samples can be aliased. The second example shows that the term “aliasing test” must be restricted to stationary signals. As stated earlier, this restriction is inherent in the definition of the bispectrum. ## 6 Conclusions So, is this something for nothing? How can one get information about higher frequency amplitudes from what is usually thought of as Nyquist-limited data? The answer is of course that the assumption of stationarity is far from nothing. But, to exploit stationarity one must be able to perform the ensemble averaging indicated in the definition of the bispectrum. This implies that one must either have an ergodic process or have access to sufficiently many sample paths. It is certainly possible that, in practice, the bispectrum can be usefully applied to signals for which there is no theoretical justification. For such uses the aliasing test is silent. However, it is essential that a clear understanding of the fundamental properties of higher-order spectra be available. And the present authors believe that correct understanding of the outer triangle leads to deeper insight of the meaning of the bispectrum in general.
no-problem/9906/astro-ph9906192.html
ar5iv
text
# The Origin of Diversity of Type Ia Supernovae and Environmental Effects ## 1. Introduction It is widely accepted that Type Ia supernovae (SNe Ia) are thermonuclear explosions of accreting C+O white dwarfs (WDs), although the nature of the progenitor binary system and the detail of the explosion mechanism are still under debate. SNe Ia are good distance indicators, and provide a promising tool for determining cosmological parameters (e.g., Branch and Tammann 1992). From the observations of high redshift SNe Ia, both the SN Cosmology Project (Perlmutter et al. 1999) and the High-z SN Search Team (Riess et al. 1998) have suggested a statistically significant value for the cosmological constant. However, SNe Ia are not perfect standard candles, but show some intrinsic variations in brightness. When determining the absolute peak luminosity of high-redshift SNe Ia, therefore, these analyses have taken advantage of the empirical relation existing between the peak brightness and the light curve shape (LCS). Since this relation has been obtained from nearby SNe Ia only (Phillips 1993; Hamuy et al. 1995; Riess, Press & Kirshner 1995), it is important to examine whether it depends systematically on environmental properties such as metallicity and age of the progenitor system. This Letter addresses the issue of whether a difference in the environmental properties is at the basis of the observed range of peak brightness. There are some observational indications that SNe Ia are affected by their environment. The most luminous SNe Ia seem to occur only in spiral galaxies, while both spiral and elliptical galaxies are hosts for dimmer SNe Ia. Thus the mean peak brightness is dimmer in ellipticals than in spiral galaxies (Hamuy et al. 1996). The SNe Ia rate per unit luminosity at the present epoch is almost twice as high in spirals as in ellipticals (Cappellaro et al. 1997). Moreover, Wang, Höflich & Wheeler (1997) and Riess et al. (1999) found that the variation of the peak brightness for SNe located in the outer regions in galaxies is smaller. Höflich, Wheeler, & Thielemann (1998) examined how the initial composition of the WD (metallicity and the C/O ratio) affects the observed properties of SNe Ia. Umeda et al. (1999) obtained the C/O ratio as a function of the main-sequence mass and metallicity of the WD progenitors. In this Letter we suggest that the variation of the C/O ratio is the main cause of the variation of SNe Ia brightness, with larger C/O ratio yielding brighter SNe Ia ($`\mathrm{\S }`$ 2). We then show that the C/O ratio depends indeed on environmental properties, such as the metallicity and age of the companion of the WD ($`\mathrm{\S }`$ 3), and that our model can explain most of the observational trends discussed above ($`\mathrm{\S }4`$). We then make some predictions about the brightness of SN Ia at higher redshift ($`\mathrm{\S }5`$). ## 2. Explosion Model and the C/O Ratio of WD Progenitors For the progenitors of SNe Ia, we adopt the single degenerate (SD) Chandrasekhar mass model, in which an accreting C-O WD explodes when its mass reaches the critical mass $`M_{\mathrm{Ia}}1.371.38M_{}`$ (Nomoto, Thielemann, & Yokoi 1984). Merging of white dwarfs is likely to lead to accretion-induced-collapse rather than thermonuclear explosion (Saio & Nomoto 1998). Chandrasekhar mass models can reproduce well the spectrum and the light curves of SNe Ia, assuming either the explosion is induced by a deflagration or by a delayed detonation (Höflich & Khokhlov 1996; Nugent et al. 1997). In these models, the brightness of SNe Ia is determined mainly by the mass of <sup>56</sup>Ni synthesized ($`M_{\mathrm{Ni56}}`$). Observational data suggest that $`M_{\mathrm{Ni56}}`$ for most SNe Ia lies in the range $`M_{\mathrm{Ni56}}0.40.8M_{}`$ (e.g. Mazzali et al. 1998). This range of $`M_{\mathrm{Ni56}}`$ can result from differences in the C/O ratio in the progenitor WD as follows. In the deflagration model, a faster propagation of the convective deflagration wave results in a larger $`M_{\mathrm{Ni56}}`$. For example, a variation of the propagation speed by 15% in the W6 – W8 models results in $`M_{\mathrm{Ni56}}`$ values ranging between 0.5 and $`0.7M_{}`$ (Nomoto et al. 1984), which could explain the observations. The actual propagation of the deflagration depends on the highly non-linear behavior of the turbulent flame (Niemeyer & Hillebrand 1995), and so it may be very sensitive to the C/O ratio. Qualitatively, a larger C/O ratio leads to the production of more nuclear energy and buoyancy force, thus leading to a faster propagation and a larger $`M_{\mathrm{Ni56}}`$. Quantitatively, further studies of the turbulent flame are necessary to confirm that the expected range of C/O results in the required 15-20% variation of the flame speed. In the delayed detonation model, $`M_{\mathrm{Ni56}}`$ is predominantly determined by the deflagration-to-detonation-transition (DDT) density $`\rho _{\mathrm{DDT}}`$, at which the initially subsonic deflagration turns into a supersonic detonation (Khokhlov 1991). We reproduce the relation between $`\rho _{\mathrm{DDT}}`$ and $`M_{\mathrm{Ni56}}`$ in Figure 1 by performing hydrodynamical calculations for several values of $`\rho _{\mathrm{DDT}}`$, as in Nomoto et al. (1997), Kishimoto et al. (1999), and Iwamoto et al. (1999). The pre-explosive WD model is model C6 (Nomoto et al. 1984), and the flame speed of the initial deflagration is assumed to be 3% of the local sound velocity. Figure 1 shows that if the transition density varies in the range $`\rho _{\mathrm{DDT}}1.33\times 10^7`$ g cm<sup>-3</sup>, the resulting variation of $`M_{\mathrm{Ni56}}`$ is large enough to explain the observations. Possible mechanisms for DDT to occur have been studied by Arnett & Livne (1994), Niemeyer & Woosley (1997), and Khokhlov, Oran & Wheeler (1997): when the deflagration wave reaches a sufficiently low density, $`\rho 10^7`$ g cm<sup>-3</sup>, the turbulent motion associated with the flame may destroy the burning front. The resulting turbulent mixing between ashes and fuels efficiently heats up the fuel and could produce a region with a very shallow temperature gradient. In such a region, successive spontaneous ignitions cause the over-driven deflagration to propagate supersonically. This may induce a detonation wave if the mass of the region exceeds a critical mass $`\mathrm{\Delta }M_{\mathrm{DDT}}`$. This critical mass is quite sensitive to the carbon mass fraction $`X`$(C), e.g. $`\mathrm{\Delta }M_{\mathrm{DDT}}10^{19}`$ and $`10^{14}M_{}`$ at $`\rho =3\times 10^7`$ g cm<sup>-3</sup> for $`X`$(C) = 1.0 and 0.5, respectively (Niemeyer & Woosley 1997). Though the exact value of $`\rho _{\mathrm{DDT}}`$ is still debated, and its dependence on $`X`$(C) has not been studied, it is not unlikely that for a larger $`X`$(C) DDT can occur at larger $`\rho _{\mathrm{DDT}}`$. Hence a larger $`X`$(C) is likely to result in a larger $`\rho _{\mathrm{DDT}}`$ and $`M_{\mathrm{Ni56}}`$. In this Letter, therefore, we postulate that $`M_{\mathrm{Ni56}}`$ and consequently brightness of a SN Ia increase as the progenitors’ C/O ratio increases, as illustrated in Figure 1, where the range of $`M_{\mathrm{Ni56}}0.50.8M_{}`$ is the result of an $`X`$(C) range $`0.350.5`$, which is the range of $`X`$(C) values of our progenitor models described below (Figure 2). The $`X`$(C) – $`M_{\mathrm{Ni56}}`$ relation we adopt is still only a working hypothesis, which needs to be proved from studies of the turbulent flame during explosion. Höflich et al. (1998) considered the dependence on $`X`$(C) but they assumed in their DDT model that $`\rho _{\mathrm{DDT}}`$ and $`X`$(C) are independent parameters. They showed that for the same $`\rho _{\mathrm{DDT}}`$ a smaller $`X`$(C) leads to a slightly brighter SN Ia despite a slightly smaller $`M_{\mathrm{Ni56}}`$ produced, because a smaller fraction of the explosion energy goes into the kinetic energy. In this Letter we assume that $`X`$(C) is the primary parameter to determine $`\rho _{\mathrm{DDT}}`$ and thus $`M_{\mathrm{Ni56}}`$. The assumed variation of $`M_{\mathrm{Ni56}}`$ is so large that a smaller $`X`$(C) yields an intrinsically dimmer SN. ## 3. Metallicity and Age Effects In this section we discuss how the C/O ratio in the WD depends on the metallicity and age of the binary system. The C/O ratio in C+O WDs depends primarily on the main-sequence mass of the WD progenitor and on metallicity. According to the evolutionary calculations for 3$``$9 $`M_{}`$ stars by Umeda et al. (1999), the C/O ratio and its distribution are determined in the following evolutionary stages of the close binary. 1) At the end of central He burning in the 3$``$9 $`M_{}`$ primary star, C/O$`<1`$ in the convective core. The mass of the core is larger for more massive stars. 2) After central He exhaustion, the outer C+O layer grows via He shell burning, where C/O$`\text{ }>1`$ (Umeda et al. 1999). 3a) If the primary star becomes a red giant (case C evolution; e.g. van den Heuvel 1994), it then undergoes the second dredge-up, forming a thin He layer, and enters the AGB phase. The C+O core mass, $`M_{\mathrm{CO}}`$, at this phase is larger for more massive stars. For a larger $`M_{\mathrm{CO}}`$ the total carbon mass fraction is smaller. 4a) When it enters the AGB phase, the star greatly expands and is assumed here to undergo Roche lobe overflow (or a super-wind phase) and to form a C+O WD. Thus the initial mass of the WD, $`M_{\mathrm{WD}}^{(0)}`$, in the close binary at the beginning of mass accretion is approximately equal to $`M_{\mathrm{CO}}`$. 3b) If the primary star becomes a He star (case BB evolution), the second dredge-up in (3a) corresponds to the expansion of the He envelope. 4b) The ensuing Roche lobe overflow again leads to a white dwarf of mass $`M_{\mathrm{WD}}^{(0)}`$ = $`M_{\mathrm{CO}}`$. 5) After the onset of mass accretion, the WD mass grows through steady H burning and weak He shell flashes, as described in the WD wind model (Hachisu, Kato, & Nomoto 1996, 1999, and Hachisu et al. 1999; hereafter, HKN96, HKN99, and HKNU99, respectively). The composition of the growing C+O layer is assumed to be C/O=1. 6) The WD grows in mass and ignites carbon when its mass reaches $`M_{\mathrm{Ia}}=1.367M_{}`$, as in the model C6 of Nomoto et al. (1984). Because of strong electron-degeneracy, carbon burning is unstable and grows into a deflagration for a central temperature of $`8\times 10^8`$ K and a central density of $`1.47\times 10^9`$ g cm<sup>-3</sup>. At this stage, the convective core extends to $`M_r=1.14M_{}`$ and the material is mixed almost uniformly, as in the C6 model. In Figure 2 we show the carbon mass fraction $`X`$(C) in the convective core of this pre-explosive WD, as a function of metallicity ($`Z`$) and initial mass of the WD before the onset of mass accretion, $`M_{\mathrm{CO}}`$. Figure 2 reveals that: 1) $`X`$(C) is smaller for larger $`M_{\mathrm{CO}}`$. 2) The dependence of $`X`$(C) on metallicity is small when plotted against $`M_{\mathrm{CO}}`$, even though the relation between $`M_{\mathrm{CO}}`$ and the initial stellar mass depends sensitively on $`Z`$ (Umeda et al. 1999). Metallicity dependent wind during mass accretion: In the SD Chandrasekhar mass model for SNe Ia, a WD explodes as a SN Ia only when its rate of the mass accretion ($`\dot{M}`$) is in a certain narrow range (e.g., Nomoto & Kondo 1991). HKN96 showed that the accreting WD blows a strong wind if $`\dot{M}`$ exceeds the rate $`\dot{M}_\mathrm{b}`$ at which steady burning can process the accreted hydrogen into He. If the wind is sufficiently strong (i.e., the wind velocity $`v_\mathrm{w}`$ exceeds the escape velocity $`v_{\mathrm{esc}}`$ of the WD), the WD can avoid the formation of a common envelope and increase its mass continuously at a rate $`\dot{M}_\mathrm{b}`$ by blowing the extra mass away in a wind. In this model, which is adopted in the present study, an interesting metallicity effect has been found (Kobayashi et al. 1998; Hachisu & Kato 1999). The wind velocity is higher for larger $`M_{\mathrm{WD}}`$ and larger Fe/H because of higher luminosity and larger opacity, respectively. In order to blow sufficiently strong wind (i.e., $`v_\mathrm{w}>v_{\mathrm{esc}}`$), $`M_{\mathrm{WD}}`$ should exceed a certain mass $`M_\mathrm{w}`$ (Fig. 6 of HKN99). As seen in Figure 1 of Kobayashi et al. (1998), $`M_\mathrm{w}`$ is larger for lower metallicity; e.g., $`M_\mathrm{w}=`$ 0.65, 0.85, and 0.95 $`M_{}`$ for $`Z=`$ 0.02, 0.01, and 0.004, respectively. In order for a WD to grow its mass at $`\dot{M}>\dot{M}_\mathrm{b}`$, its initial mass $`M_{\mathrm{CO}}`$ should exceed $`M_\mathrm{w}`$. In other words, $`M_\mathrm{w}`$ is the metallicity-dependent minimum $`M_{\mathrm{CO}}`$ required for a WD to become an SN Ia (strong wind condition in Fig.2). The upper bound $`M_{\mathrm{CO}}1.07M_{}`$ is imposed by the condition that carbon should not ignite and is almost independent of metallicity. As shown in Figure 2, the range of $`M_{\mathrm{CO}}`$ can be converted into a range of $`X`$(C). From this we find the following metallicity dependence for $`X`$(C): 1) The upper bound of $`X`$(C), which is determined by the lower limit on $`M_{\mathrm{CO}}`$ imposed by the metallicity-dependent conditions for a strong wind, e.g., $`X`$(C) $`\text{ }<0.51`$, 0.46 and 0.41, for $`Z`$=0.02, 0.01, and 0.004, respectively. 2) On the other hand, the lower bound, $`X`$(C) $`0.350.33`$, does not depend much on $`Z`$, since it is imposed by the maximum $`M_{\mathrm{CO}}`$. 3) Assuming the relation between $`M_{\mathrm{Ni56}}`$ and $`X`$(C) given in Figures 1 and 2, our model predicts the absence of brighter SNe Ia in lower metallicity environment. Age effects: In our model, the age of the progenitor system also constrains the range of $`X`$(C) in SNe Ia. In the SD scenario, the lifetime of the binary system is essentially the main-sequence lifetime of the companion star, which depends on its initial mass $`M_2`$. HKNU99 and HKN99 have obtained a constraint on $`M_2`$ by calculating the evolution of accreting WDs for a set of initial masses of the WD ($`M_{\mathrm{WD}}^{(0)}M_{\mathrm{CO}}`$) and of the companion ($`M_2`$), and the initial binary period ($`P_0`$). In order for the WD mass to reach $`M_{\mathrm{Ia}}`$, the donor star should transfer enough material at the appropriate accretion rates. The donors of successful cases are divided into two categories: one is composed of slightly evolved main-sequence stars with $`M_21.73.6M_{}`$ (for $`Z`$=0.02), and the other of red-giant stars with $`M_20.83.1M_{}`$ (for $`Z`$=0.02) (HKN99, HKNU99; also Li & van den Heuvel 1997). If the progenitor system is older than 2 Gyr, it should be a system with a donor star of $`M_2<1.7M_{}`$ in the red-giant branch. Systems with $`M_2>1.7M_{}`$ become SNe Ia in a time shorter than 2 Gyr. Likewise, for a given age of the system, $`M_2`$ must be smaller than a limiting mass. This constraint on $`M_2`$ can be translated into the presence of a minimum $`M_{\mathrm{CO}}`$ for a given age, as follows: For a smaller $`M_2`$, i.e. for the older system, the total mass which can be transferred from the donor to the WD is smaller. In order for $`M_{\mathrm{WD}}`$ to reach $`M_{\mathrm{Ia}}`$, therefore, the initial mass of the WD, $`M_{\mathrm{WD}}^{(0)}M_{\mathrm{CO}}`$, should be larger. This implies that the older system should have larger minimum $`M_{\mathrm{CO}}`$ as indicated in Figure 2. Using the $`X`$(C)$`M_{\mathrm{CO}}`$ and $`M_{\mathrm{Ni56}}`$X$`(C)`$ relations (Figs. 1 and 2), we conclude that WDs in older progenitor systems have a smaller $`X`$(C), and thus produce dimmer SNe Ia. ## 4. Comparison with Observations The first observational indication which can be compared with our model is the possible dependence of the SN brightness on the morphology of the host galaxies. Hamuy et al. (1996) found that the most luminous SNe Ia occur in spiral galaxies, while both spiral and elliptical galaxies are hosts to dimmer SNe Ia. Hence, the mean peak brightness is lower in elliptical than in spiral galaxies. In our model, this property is simply understood as the effect of the different age of the companion. In spiral galaxies, star formation occurs continuously up to the present time. Hence, both WD+MS and WD+RG systems can produce SNe Ia. In elliptical galaxies, on the other hand, star formation has long ended, typically more than 10 Gyr ago. Hence, WD+MS systems can no longer produce SNe Ia. In Figure 3 we show the frequency of the expected SN I for a galaxy of mass $`2\times 10^{11}M_{}`$ for WD+MS and WD+RG systems separately as a function of $`M_{\mathrm{CO}}`$. Here we use the results of HKN99 and HKNU99, and the $`M_{\mathrm{CO}}X`$(C) and $`M_{\mathrm{Ni56}}X`$(C) relations given in Figure 2. Since a WD with smaller $`M_{\mathrm{CO}}`$ is assumed to produce a brighter SN Ia (larger $`M_{\mathrm{Ni56}}`$), our model predicts that dimmer SNe Ia occur both in spirals and in ellipticals, while brighter ones occur only in spirals. The mean brightness is smaller for ellipticals and the total SN Ia rate per unit luminosity is larger in spirals than in ellipticals. These properties are consistent with observations. The second observational suggestion is the radial distribution of SNe Ia in galaxies. Wang et al. (1997) and Riess et al. (1998) found that the variation of the peak brightness for SNe Ia located in the outer regions in galaxies is smaller. This behavior can be understood as the effect of metallicity. As shown in Figure 2, even when the progenitor age is the same, the minimum $`M_{\mathrm{CO}}`$ is larger for a smaller metallicity because of the metallicity dependence of the WD winds. Therefore, our model predicts that the maximum brightness of SNe Ia decreases as metallicity decreases. Since the outer regions of galaxies are thought to have lower metallicities than the inner regions (Zaritsky, Kennicutt, Huchra 1994; Kobayashi & Arimoto 1999), our model is consistent with observations. Wang et al. (1997) also claimed that SNe Ia may be deficient in the bulges of spiral galaxies. This can be explained by the age effect, because the bulge consists of old population stars. ## 5. Conclusions and Discussion We have suggested that $`X`$(C) is the quantity very likely to cause the diversity in $`M_{\mathrm{Ni56}}`$ and thus in the brightness of SNe Ia. We have then shown that our model predicts that the brightness of SNe Ia depends on the environment, in a way which is qualitatively consistent with the observations. Further studies of the propagation of the turbulent flame and the DDT are necessary in order to actually prove that $`X`$(C) is the key parameter. Our model predicts that when the progenitors belong to an old population, or to a low metal environment, the number of very bright SNe Ia is small, so that the variation in brightness is also smaller. In spiral galaxies, the metallicity is significantly smaller at redshifts $`z\text{ }>1`$, and thus both the mean brightness of SNe Ia and its range tend to be smaller. At $`z\text{ }>2`$ SNe Ia would not occur in spirals at all because the metallicity is too low. In elliptical galaxies, on the other hand, the metallicity at redshifts $`z13`$ is not very different from the present value. However, the age of the galaxies at $`z1`$ is only about 5 Gyr, so that the mean brightness of SNe Ia and its range tend to be larger at $`z\text{ }>1`$ than in the present ellipticals because of the age effect. We note that the variation of $`X`$(C) is larger in metal-rich nearby spirals than in high redshift galaxies. Therefore, if $`X`$(C) is the main parameter responsible for the diversity of SNe Ia, and if the LCS method is confirmed by the nearby SNe Ia data, the LCS method can also be used to determine the absolute magnitude of high redshift SNe Ia. We wish to thank P. Mazzali, K. Iwamoto and N. Kishimoto for useful discussion and suggestions. This work has been supported in part by the grant-in-Aid for Scientific Research (0980203, 09640325) and COE research (07CE2002) of the Ministry of Education, Science, Culture and Sports in Japan.
no-problem/9906/hep-ex9906020.html
ar5iv
text
# Inclusive Jet Cross Sections in 𝑝̄⁢𝑝 Collisions at √𝑠= 630 and 1800 GeV ## 1 Introduction Within the framework of quantum chromodynamics (QCD), inelastic scattering between a proton and antiproton is described as a hard collision between their constituents (partons). After the collision, the outgoing partons manifest themselves as localized streams of particles or “jets”. Predictions for the inclusive jet cross section have improved in the early nineties with next-to-leading order (NLO) perturbative QCD calculations and new, accurately measured parton density functions (pdf). DØ has recently measured and published the cross section for the production of jets as a function of the jet energy transverse to the incident beams, $`E_T`$. The measurement is based on an integrated luminosity of about 92 pb<sup>-1</sup> of $`\overline{p}p`$ hard collisions collected with the DØ Detector at the Fermilab Tevatron Collider. This result allows a stringent test of QCD, with a total uncertainty substantially reduced relative to previous results . We also measure the ratio of jet cross sections at two center-of-mass energies: 630 (based on an integrated luminosity of about 0.537 pb<sup>-1</sup>) and 1800 GeV . Experimental and theoretical uncertainties are significantly reduced in the ratio. This is due to the large correlation in the errors of the two cross section measurements, and the suppression of the sensitivity to parton distribution functions (pdf) in the prediction. The ratio of cross sections thus provides a stronger test of the matrix element portion of the calculation than a single cross section measurement alone. Previous measurements of cross section ratios have been performed with smaller data sets by the UA2 and CDF experiments. ## 2 Jet Reconstruction and Data Selection Jets are reconstructed using an iterative jet cone algorithm with a cone radius of $``$=0.7 in $`\eta `$$`\varphi `$ space, (pseudorapidity is defined as $`\eta =\mathrm{ln}[\mathrm{tan}\frac{\theta }{2}]`$. The offline data selection procedure, which eliminates background caused by electrons, photons, noise, or cosmic rays, follows the methods described in Refs. . ## 3 Energy Corrections The jet energy scale correction, described in , removes instrumentation effects associated with calorimeter response, showering, and noise, as well as the contribution from spectator partons (underlying event). The energy scale corrects jets from their reconstructed $`E_T`$ to their “true” $`E_T`$ on average. An unsmearing correction is applied later to remove the effect of a finite $`E_T`$ resolution . ## 4 The Inclusive Jet Cross Section The resulting inclusive double differential jet cross sections, $`d^2\sigma /(dE_Td\eta )`$, for $`|\eta |0.5`$ and $`0.1|\eta |0.7`$ (the second region for comparison to Ref. ), are compared with a NLO QCD theoretical prediction . Discussions on the different choices in the theoretical calculation: pdfs, renormalization and factorization scales ($`\mu `$), and clustering algorithm parameter ($`R_{sep}`$) can be found in Refs. . Figure 1 shows the ratios $`(DT)/T`$ for the data ($`D`$) and JETRAD NLO theoretical ($`T`$) predictions based on the CTEQ3M, CTEQ4M and MRST pdf’s for $`|\eta |0.5`$. (The tabulated data for both $`|\eta |0.5`$ and $`0.1|\eta |0.7`$ measurements can be found in Ref. .) The predictions are in good quantitative agreement with the data, as verified with a $`\chi ^2=_{i,j}(D_iT_i)(C^1)_{ij}(D_jT_j)`$ test, which incorporates the uncertainty covariance matrix $`C`$. Here $`D_i`$ and $`T_i`$ represent the $`i`$-th data and theory points, respectively. The overall systematic uncertainty is largely correlated. Table 1 lists $`\chi ^2`$ values for several JETRAD predictions using various parton distribution functions . The predictions describe both the $`|\eta |0.5`$ and $`0.1|\eta |0.7`$ cross section very well. The measurement by DØ and CDF are also in good quantitative agreement within their systematic uncertainties . ## 5 $`\eta `$ Dependence of $`d^2\sigma /(dE_Td\eta )`$ DØ has made a preliminary measurement of the pseudorapidity dependence of the inclusive jet cross section. Figure 2 shows the ratios $`(DT)/T`$ for the data ($`D`$) and JETRAD NLO theoretical ($`T`$) predictions using the CTEQ3M pdf set for $`0.5|\eta |1`$ and $`1|\eta |1.5`$. The measurements and the predictions are in good qualitative agreement. A detailed error analysis is currently being completed. ## 6 Ratio of Scale Invariant Jet Cross Sections A simple parton model would predict a jet cross section that scales with center-of-mass energy. In this scenario, $`E_T^4E\frac{d^3\sigma }{dp^3}`$, plotted as a function of jet $`x_T\frac{2E_T}{\sqrt{s}}`$, would remain constant with respect to the center-of-mass energy. Figure 3 shows the DØ measurement of $`E_T^4E\frac{d^3\sigma }{dp^3}`$ (stars) compared to JETRAD predictions (lines). There is poor agreement between data and NLO QCD calculations using the same $`\mu `$ in the numerator and the denominator (probability of agreement not greater than 10%). The agreement improves for predictions with different $`\mu `$ at the two center-of-mass energies . In conclusion, we have made precise measurements of jet production cross sections. At $`\sqrt{s}`$=1800 GeV, there is good agreement between the measurements and the data. The ratio of cross sections at $`\sqrt{s}`$=1800 and 630 GeV, however, differs from NLO QCD predictions, unless different renormalization scales are introduced for the two center-of-mass energies.
no-problem/9906/quant-ph9906014.html
ar5iv
text
# 1 INTRODUCTION — NORMAL SCIENCE ## 1 INTRODUCTION — NORMAL SCIENCE There was difficulty during these discussions reaching any consensus on what was meant by “paranormal”. In the end we did not try. I suspect that part of the problem was that our diverse group does not agree on what is “normal science”, making a sharp, contrasting definition of “paranormal” phenomena impossible for us in the first place. I have therefore decided that, before I can explain to you how I try to think about paranormal phenomena, I must first explain how I think about ordinary science. I am a physicist. For me, as for many others, physics is an empirical science based on quantitative measurements mutually agreed on by a community of practitioners of physics. That such a community exists, but has come into existence only since the “scientific revolution” of the seventeenth century, I take to be an established historical fact. In this sense, I take agreed upon laboratory protocol and practice to be primary and the mathematical language and other technical terms used in describing how, up to a point, agreement between members of the community is achieved to be secondary. Both evolve over time, and bring in other communities, as is well illustrated by Peter Galison’s incisive examination of the objects on the laboratory floor which constitute the material culture of particle physics in this century . What concerns me here is not so much particle physics per se, but how its conclusions are extended to provide a framework with which to describe the past. Since I have presented at this meeting the cosmological framework that comes out of Program Universe and its connection to bit-string physics, I will be brief. The basic assumptions are: a) the Galilean assumption that processes we observe occurring here and now will — until we have evidence to the contrary — occur a similar way under similar circumstances elsewhere in the cosmos; b) the assumption that (except under special circumstances described by the General Theory of Relativity) light travels at the limiting velocity $`c`$ if unimpeded by matter; c) on a large enough scale (which has now been achieved, thanks to the Hubble Space Telescope) the universe at any epoch is homogeneous and isotropic, leading to the Friedman-Robertson-Walker metric for the macroscopic framework into which we fit our observations. Extrapolating back to 13 billion years from the present (thanks to a number of recent developments ) now provides a consistent description of the evolution of the cosmos within our event horizon, with a number of detailed cross-checks. This sounds like a departure from my commitment to an operational stance about space and time. So I emphasize that this picture only refers to physical phenomena we can measure and/or observe here and now. I remain skeptical, even doubtful, as to whether these successes establish the “reality” of space and time in any deep sense. Clearly, as with “common sense” space and time, they form a useful descriptive framework, if we do not commit the error of casting it in concrete. I consider it a real triumph of the ANPA program that we can arrive at this framework from the combinatorial hierarchy construction via program universe or any similar algorithm without postulating any a priori space time. Granted this background, the older story of the origin of the solar system, of biomolecular chirality and biopoesis , and of terrestrial biological evolution falls into its appropriate niche. Recent work, which I will not bother to cite, has enormously deepened and enriched this description and (for me, at least) strengthened my conviction that no major lacunae remain. I stress that the “here and now” sciences — physics, chemistry, biology, … — are a necessary background for understanding the historical sciences in the broad sense: cosmology, stellar and solar system evolution, terrestrial biological evolution, evolution of human intelligence and language, social evolution, political evolution. As we proceed up the chain from physics to politics, the scientific disciplines become more and more contingent on unique, local events whose prevalence in the rest of the cosmos we can currently only guess at. However, the recent discovery of many extra-solar planetary systems in our immediate neighborhood makes it possible that, in the not too distant future, some of these guesses about exobiology may be replaced by hard fact. We may also be on the threshold of understanding the co-evolution of language and the brain in the human species if Deacon, among others, is to be believed. ## 2 WHAT ABOUT PARANORMAL PHENOMENA? Much recent work on “paranormal phenomena” has amounted to getting large statistical samples with small deviations from “chance” which are unexplained. Much of this work has considerably higher methodological standards than most scientific work. However, for those familiar with experimental physics (and presumably in many other fields as well) this will never be convincing. We are all too familiar with unexplained effects that cannot be attributed to “chance”. For us these are examples of systematic error, and if they cannot be brought under control, simply characterize a bad experiment. One has to understand the sources of systematic error, show that they vary in a systematic way with changes in experimental conditions, and do one’s best to bring them down below the effects of statistical error. For this, of course, one needs a theory, not only of the phenomenon being investigated, but also a theory of what is (or is likely to be) interfering with the measurement. I do not see how this situation can be achieved in investigations of paranormal phenomena without much more theoretical work using a framework that allows for the testing of hypothesis and their rejection. In this I agree with what Etter and Shoup have already said at this meeting, and in this discussion. But I would go further and say that one needs not only a quantitative theory for the phenomena themselves, but also a theory for sources of systematic error in a form which can also be tested. The impetus for research into paranormal phenomena has not come, and does not now come, from small, inexplicable effects. Judging by material presented in this discussion, and from my own contacts with scientists interested in the subject, I assert that this interest usually arises from personal experience. I have never had any “paranormal” experience. But people I respect, including some at this meeting, tell me they have. So I take the possibility that some people have this capacity seriously. I also do not get much out of listening to music. But I have plenty of evidence that many people do. In both respects I am not unusual. I start with an incident I heard of three decades ago, which was told to me by an anthropologist. In brief, while working one day in the Pacific Northwest with a shaman he had known for several months, the shaman asked suddenly if the anthropologist would like to know what the anthropologist’s friend in Chicago was doing just then. Of course he said yes. Equipped with the shaman’s response, the anthropologist documented it, wrote to his friend in Chicago and got a statement of what he was doing at that time. The correspondence between the shaman’s report and the friend’s statement was so close that the anthropologist, twenty years later, was still afraid to publish for fear it would damage his professional reputation. I didn’t know what to do with this story at the time. However a year or so later I proved that when a system with two quantum mechanical particles interacting via short range forces is augmented by a third particle with similar interactions, the behavior of the pair changes no matter how far away the third particle is. I called this example of the extreme non-locality of quantum mechanics the eternal triangle effect, and compared it analogically with the above instance and other behavioral examples. The analysis I subsequently published provides a good starting point for discussing my current position. I quote: > It is not necessary for you to believe the story in order to ask the question, as I do, of how such a remarkable ‘communication’ might occur. After much rumination on the event, and after the discovery of the eternal triangle effect and its behavioral analog, I have come to a tentative model, or rather explanatory framework. Since the anthropologist and the shaman had reached a mutual level of confidence and trust, they could to a certain extent ‘share each other’s thoughts’ — \[a\] phenomenon known to all of us, and not necessarily involving any paranormal phenomena It is relevant here that the anthropologist was one of the founders of kinesics; he once told me that given only a minute or so of the start of a filmed psychotherapeutic session, he could predict what would happen during the rest of the hour.. Further, the anthropologist knew his distant friend well, and might by \[a\] similar process anticipate (unconsciously) what his friend would be doing at the time. We know of many instances when such unconscious deductions come to us in dreams — sometimes accurate and sometimes not. For the shaman to ‘pick up’ this knowledge or conjecture from the anthropologist need involve only the types of ‘non-verbal communication’ discussed in this volume. and which, though often difficult to understand, model, or demonstrate, are again familiar aspects of human behavior. Granted only the postulate that a human mind makes makes many accurate deductions about present \[and future\] happenings from past experience — which would shock no psychoanalyst — the whole incident can be fitted into the framework of explanatory models that, separately, are often accepted. > > It is interesting to speculate on whether many phenomena which are called ‘paranormal’ might not fit into such an explanatory framework. The ‘framework’ does not really explain anything, of course. To account for an unexplained occurrence by saying that the human mind can make, unconsciously, very accurate deductions about what will occur (‘precognition’), what another person is thinking (‘telepathy’), or how an unstable system will behave (predictive ‘telekinesis’) is only to replace one problem with another — namely how to explain this extraordinary computational ability. But it does have the aspect of explaining a fact that is troublesome in ‘paranormal research’, namely that the ability is not 100% and closely tied to the emotional state of the individual<sup>§</sup><sup>§</sup>§In the light of our discussion of systematic error above, it occurs to me that ‘emotional state’ of both subject and experimenter is one factor that cries out for quantitative assessment and investigation in this field — perhaps an impossible task?. This is what we would expect, from psychoanalytic theory, of a process deeply buried in the unconscious. Coming back to the theme of this volume, such unconscious processes clearly can have an important bearing on non-verbal communication of more conventional sorts, and it is perhaps reassuring that the underlying physics warns us we should include them in our thinking about how such communications work. > > My intention in this essay is not to say that quantum mechanics ‘explains’ paranormal phenomena by some such route. What I do claim is that quantum mechanics, in the simplest case where the phenomena can occur (the three particle problem with finite range interactions), does require both an extreme nonlocality of description when forced into an ‘instantaneous’ or ‘static’ form, and the inclusion (in principle) of all past events in the discussion of the current situation. I hope that this fact can provide an ‘explanatory framework’ within which it is easier to contemplate correlations between events so distant in space and time from each other as to make models drawn from classical physics seem inadequate or implausible. My first criterion for the establishment of a scientific study of paranormal phenomena is that it be capable of convincing skeptics like me that meaningful experimental investigation is possible in the first place. If the investigations are statistical, it is all too easy to dismiss their results as due to unexplained systematic error. If they are anecdotal, it is all to easy to fall back, as I have done in the analysis just quoted, on some form of unexplained “unconscious” effect that falls more properly in the domain studied by psychiatrists than in a new discipline. I am afraid that all too many “scientists” are uncomfortable living in a world in which most of the important things in life are unexplained, and grasp at facile explanations or rejections. For me the true scientist lives with uncertainty as his constant companion, and never expects that situation to change. But that does not mean that new facts and methods are to be avoided; rather, they should be eagerly pursued. I look at one new possibility in the next section. ## 3 A NEW METHODOLOGY? We have already heard from Etter and Shoup about a new approach to the study of paranormal phenomena based on new theoretical insights that have come out of Etter’s work on the foundations of quantum mechanics. Since the up to date material will not be available for a while in written form, I refer you to an older paper of Tom’s, which is now available on the web. What Tom does is to show that the core laws of quantum mechanics (Born’s probability rule that gives probabilities as the squares of “amplitudes”, and the unitary evolution of the quantum state called Schroedinger’s equation) are simply a piece of mathematics which has no physics in it. This allows him to formalize the Markov chains (which are irreversible) with either past or future boundary conditions, and use the same framework to describe the time-reversible Schroedinger evolution. Thus classical (statistical) systems peacefully coexist with quantum systems, as they must in quantum measurement theory. Hopefully, his discussion of quantum measurement theory will make this subject less paradoxical for some who have trouble with it. Although his approach provides a new way of looking at quantum mechanics, at this stage no new predictions are made. What makes Etter’s analysis exciting from the point of view of this paper is that in addition to quantum mechanics, the formalism allows a clean description of phenomena, such as “future causation”, which appear to occur in many reports of paranormal phenomena. But this descriptive framework, being general, is not tied to Planck’s constant. Thus it provides for the possibility of macroscopic acausality which, as I indicated in the last section, is analogically suggested by quantum mechanics, but without giving a clue as to how to make a systematic theory for it. Even having a theory is useless, except as an aid to imagination, until a way is found to fit experimental results into the theoretical framework. The payoff is when experimental results thus formulated lead to a reliable technology which can join the everyday world of fact. I must confess that I am skeptical whether this can be done for paranormal phenomena, but I enthusiastically support Etter and Shoup’s efforts to take this step. ## 4 CONCLUSION I conclude by turning to Ted Bastin’s proposed definition of the paranormal. He started with the proposition that paranormal phenomena show no dependence on space and time. He then coupled this to his further assumption that current physics begins with space and time. These two propositions in conjunction make a clash with normal science inevitable. This need not be the case. I am not the only contemporary physicist who feels the need to construct space and time as part of the foundations of physics. Since I think of myself as doing “normal science”, or posibly as encouraging a paradigm shift which will turn out to be acceptable by normal scientists, I cannot accept the second half of Ted’s position. I quite agree with Ted that many scientists do start by uncritically accepting either the continuum space-time of physics or the cruder space-time of “common sense” as the given theatre in which the dramas they study take place. But I do not go along with them. In fact many people accept the fact that demonstrated macroscopic quantum phenomena such as supraluminal correlation without supraluminal signalling over distances of 20 kilometers, and the teleportation of photons (destruction at one position and recreation of the same photon at a separate space-time location) show that “space-time” is more complicated than the Maxwellian picture allows for. Similar remarks could be made about black holes and modern cosmology. Thus, for me, it comes down to whether the best strategy for getting on with the job of obtaining a better understanding of “paranormal phenomena” is to follow a course that inevitably leads to confrontation, or to find a way to expand “normal science” so that it can include such phenomena. Obviously, from what I have said in this paper, I currently favor the latter course. But I am ready to be convinced that this is a mistake. I end by giving my heartfelt thanks to the Epiphany Philosophers for making possible these two days of very interesting discussion.
no-problem/9906/astro-ph9906453.html
ar5iv
text
# Does magnetic pressure affect the ICM dynamics? ## 1 Introduction Since the work of Loeb & Mao (1994), the possibility of explaining the discrepancies on mass determinations, found by Miralda-Escudé & Babul (1995), via non-thermal pressure support has been widely discussed (see also Wu & Fang 1996, 1997; Wu et al. 1998). The discrepancy arises from the two most promising techniques to obtain clusters of galaxies masses. On one hand, the determination of masses in clusters of galaxies, via X-ray data, is based on the hypothesis that the ICM is in hydrostatic equilibrium with the gravitational potential, using the radial profiles of density and temperature. There are uncertainties in the determination of temperature profiles, particularly for radii $`>1\mathrm{Mpc}`$, and for most systems only a mean emission-weighted X-ray temperature is available (radial temperature profiles are available only for a few clusters ,e.g., Allen & Fabian 1994; Nulsen & Böhringer 1995). On the other hand, gravitational lensing measures the projected surface density of matter, a method which makes no assumptions on the dynamical state of the gravitating matter (Fort & Mellier 1994; Miralda-Escudé & Babul 1995; Smail et al. 1997). One can find in the literature some attempts to resolve the discrepancy between X-ray and gravitational lensing mass measurements of clusters of galaxies. For instance, Allen (1998) studied in detail a sample of 13 galaxy clusters (including cooling flows, intermediate and non-cooling flows systems) with the goal of comparing X-ray and lensing mass measurements. His conclusions pointed out that, at least for cooling flows systems, being more relaxed systems, this discrepancy is completed resolved, and therefore, non-thermal pressures can be discarded in these systems. The magnetic field of the ICM can be obtained via Faraday rotation, due to the effect of magnetic field on the polarized radio emission from the cluster or the background radio sources. The polarization plane of linearly polarized radiation is rotated during the passage through a magnetized plasma. The angle of rotation is $`\varphi =(RM)\lambda ^2`$, where $`RM`$ is the rotation measure and $`\lambda `$ the radiation wavelength (Sarazin 1992, for a review). In clusters with diffuse radio emission, X-ray observations can give a lower limit to the strength of the magnetic field. Typically, this limit is $`B0.1\mu \mathrm{G}`$ (Rephaeli et al. 1987) on scales of $`1\mathrm{Mpc}`$. In the case of Faraday rotation the information obtained is the upper limit on the intensity of the field, and the measured values are $`RM100\mathrm{rad}/\mathrm{m}^2`$, that is more or less consistent with a intracluster field of $`B1\mu \mathrm{G}`$, with a coherence length of $`l_B10\mathrm{kpc}`$. This strength of the magnetic field corresponds to a ratio of magnetic to gas pressure of $`p_B/p_{gas}`$ $`10^3`$, implying that $`B`$ does not influence the cluster dynamics (at least on large scales). At inner regions, of the cooling flow clusters, the magnetic fields are expected to be amplified due to the gas compression (Soker & Sarazin 1990). If they are frozen in the mass flow flux, and if this flux is homogeneous and spherically symmetric, $`Br^1`$ and $`RMr^1`$, ($`p_Br^2`$ and the gas pressure increases slowly). Even in this case $`p_B`$ reaches equipartition at a radius $`r_B`$ of $`r_B1\mathrm{kpc}\left(\frac{B}{1\mu \mathrm{G}}\right)^{1/2}\left(\dot{M}/100\mathrm{M}_{}\mathrm{yr}^1\right)^{1/3}`$. In these inner regions many sources with very strong Faraday rotations were observed, in which the rotation measure can reach values of $`RM4000\mathrm{rad}/\mathrm{m}^2`$ (radio sources associated with the central galaxies of the clusters with very strong cooling flows (M87/Virgo, Cyg A, Hydra A, 3C 295, A1795)), implying, $`B10\mu \mathrm{G}`$ at $`l_B1\mathrm{kpc}`$ (Taylor & Perley 1993; Ge & Owen 1993, 1994). These observations strongly suggest that the Faraday rotation is created by magnetic fields within the cooling flow clusters. Another promising method to estimate the cluster scale magnetic field, as cited above, is the detection of co-spatial inverse Compton X-ray emission with the synchrotron plasma emission (the $`3\mathrm{K}`$ background photons scattering off the relativistic electrons can produce a diffuse X-ray emission) (Rephaeli & Gruber 1988). Therefore, this method provides limits on the cluster scale magnetic fields, in addition of limits on the non-thermal amount of X-ray emission (or even on the relativist electrons energy) in galaxy clusters. Such a kind of detection of clusters magnetic fields leads, using ROSAT PSPC data and also $`327\mathrm{MHz}`$ radio map of Abell 85 (a cooling flow cluster, with a central dominant cD galaxy and about $`100\mathrm{M}_{}/\mathrm{yr}`$), to an estimate of $`(0.95\pm 0.10)\mu \mathrm{G}`$ (Bagchi et al. 1998). However, even non-cooling flows clusters present this diffuse, relic radio source which can be used to estimate magnetic field strength. For instance Ensslin & Biermann (1998) studied limits on the Coma cluster magnetic field strength, using this multifrequency observations. They shown that the central magnetic field limit is $`B>0.3\mu \mathrm{G}`$. Others have determined the strength of the magnetic field for Coma cluster, using different techniques and obtaining similar values: $`B1.2\mu \mathrm{G}`$ (Lieu et al. 1996); $`B>0.4\mu \mathrm{G}`$ (Sreekumar et al. 1996). For the same cluster (Coma), but using Faraday rotation measure, Feretti et al. (1995) estimated magnetic fields of $`6.0`$$`\mu \mathrm{G}`$ (at scales of $`1\mathrm{kpc}`$), and of $`1.7\mu \mathrm{G}`$ (at scales of $`10\mathrm{kpc}`$) was estimated by Kim et al. (1990). The above scenario allow us conclude that for both methods the observational resolution of the telescope limits the detection of smaller scales magnetic fields, implying that at scales smaller than $`1\mathrm{kpc}`$ the magnetic field strength can be higher (Ensslin et al. 1997). Another point to be noted is that Faraday rotation measures always gives values higher than inverse Compton/CBM measures. Anyway, these fields are present in the ICM and therefore justify the such a kind of study we present here. Other theoretical works concerning the magnetic pressure on the ICM are available (for instance Soker & Sarazin 1990; Tribble 1993; Zoabi et al. 1996) and we briefly compare our results with those obtained by these authors. Our goal in this paper is trying to answer the question whether or not magnetic support can be relevant in cooling flow clusters, using a more realistic treatment of the magnetic field geometric evolution. The scope of the paper is the following: in Section 2 we present the hydrodynamical equations and the method applied for their solution; Section 3 describes our models and results compared to the available observations; and in Section 4 we discuss our results in the light of others obtained in previous works, as well our main conclusions. ## 2 Evolution of the ICM with Magnetic Pressure The evolution of the intracluster gas is obtained by solving the hydrodynamic equations of mass, momentum and energy conservation: $$\frac{\rho }{t}+\frac{1}{r^2}\frac{}{r}\left(r^2\rho u\right)=\omega \rho $$ (1) $$\frac{u}{t}+u\frac{u}{r}=\frac{1}{\rho }\frac{p_t}{r}\frac{GM(r)}{r^2}$$ (2) $$\frac{U}{t}+u\frac{U}{r}=\frac{p_t}{\rho ^2}\left(\frac{\rho }{t}+u\frac{\rho }{r}\right)\mathrm{\Lambda }\rho $$ (3) where $`u`$, $`\rho `$, $`p_t`$, $`U`$ are the gas velocity, density, total pressure and the specific internal energy. The equation of state relates $`U`$ and the temperature, $$U=\frac{3}{2}\frac{k_BT}{\mu m_H}$$ (4) ($`k_B`$ is the Boltzmann’s constant, $`m_H`$ is the hydrogen atom mass and $`\mu =0.62`$ is the mean molecular weight of a fully ionized gas with $`10\%`$ helium by number). The mass distribution, $`M(r)`$, is due to the contribution of the X-rays emitting gas plus the cluster collisionless matter (which is the sum of the contributions of galaxies and dark matter – the latter being dominant), i.e, $`M(r)=M_g(r)+M_{cl}(r)`$. $`M_{cl}(r)`$ follows $$\rho _{cl}(r)=\rho _0\left(1+\frac{r^2}{a^2}\right)^{3/2}$$ (5) in which $`\rho _0`$ and $`a`$ (the cluster core radius) are related to $`\sigma `$ (the line-of-sight velocity dispersion) via: $`9\sigma ^2=4\pi Ga^2\rho _0`$. The total pressure $`p_t`$ is the sum of thermal and magnetic pressure, e.g., $`p_t=p+p_B`$. The constraints to the magnetic pressure come from observations, from which $`p_B=B^2/8\pi 4\times 10^{14}\mathrm{erg}\mathrm{cm}^3\mathrm{s}^1`$ (cf. Bagchi et al. 1998) for a diffuse field located at $`700h_{50}^1\mathrm{kpc}`$ from the cluster center. Along this paper we will use mostly the ratio between magnetic and thermal pressures, or the $`\beta `$-parameter, $`\beta =p_B/p`$. The sink term $`\omega \rho `$ in the mass equation describes the removal of mass from the gas flow by thermal instabilities. The importance of the gas removal was studied in detail by Friaça (1993) following the q-description described by White & Sarazin (1987). The sink is particularly important when one searches for a steady state solution of the cooling flow without an implausible huge accumulation of mass at the center. In fact, the condensations formed by the sink will probably give rise to stars, planetary bodies or cold dense clouds which in turn will constitute a halo surrounding the central dominant galaxy. We assume isobaric removal, so that the sink does not introduce any additional term in the energy equation. Summing up the physics contained in this term one can say that the specific mass removal rate is $`\omega =q/t_c`$ where the denominator is the instantaneous isobaric cooling time, such that the removal efficiency $`q`$ relates the cooling time to the growth time scale of the thermal instability in the cooling flow. We assume $`q`$ between $`1.0`$ and $`1.5`$, which are the $`q`$-values found to be more consistent with the observations (Friaça 1993). The cooling function adopted $`\mathrm{\Lambda }(T)`$ is the cooling rate per unit volume. Since there is no ionization equilibrium for temperatures lower than $`10^6`$ K, we adopt a non-equilibrium cooling function for the gas at $`T<10^6`$ K (the recombination time of important ions is longer than the cooling time at these temperatures). The cooling function was calculated with the atomic database of the photoionization code AANGABA (Gruenwald & Viegas 1992). The adopted abundances are sub-solar as appropriate for the ICM (Edge & Stewart 1991; Fabian 1994; Grevesse & Anders 1989). Despite the presence of steep temperature gradients we did not consider thermal conduction in our models. This can be justified using the fact that on a global scale cooling flow clusters contain cooler gas near the center and hotter gas further out. Therefore, the presence of cooling flows is itself a proof that thermal conduction effect is, at least, reduced in the ICM. Models show that thermal conduction would erase the observed density and temperature gradients in cooling flows, unless it is inhibited (see, for instance, Friaça 1986; David & Bregman 1989). It is well known that even weak magnetic field, if it is tangled, can inhibit the thermal conduction perpendicular to the field lines. More recently it has been argued that electromagnetic instabilities driven by temperature gradients (or electric currents in other situations) also can cause this inhibition in cooling flows (Pistinner et al. 1996), even for non-tangled field lines. A spherically symmetric Eulerian code is employed for the calculations, which are solved via the finite-difference scheme based on Cloutman (1980). The grid points are spaced logarithmically, with a grid of 100 cells, with the first being $`50\mathrm{pc}`$ wide. The innermost cell edge is located at $`100\mathrm{pc}`$ and the outer boundary at twice the tidal radius of the cluster. The artificial viscosity for the treatment of the shocks follows the formulation of Tscharnuter & Winkler (1979) based on the Navier-Stokes equation. The outer boundary conditions on pressure and density are derived by including an outer fictitious cell, the density and pressure in which are obtained from extrapolation of power laws over the radius fitted to the five outermost real cells. The inner boundary conditions are adjusted according to whether inflow (velocity at the inner boundary is extrapolated from the velocities at the innermost cell edges) or outflow (velocity is set zero) prevails locally. The initial conditions for the gas are an isothermal atmosphere ($`T_0=10^7\mathrm{K}`$) with $`30\%`$ solar abundances and density distribution following that of the cluster dark matter. The evolution is followed until the age of $`14\mathrm{Gyr}`$. The initial $`\beta `$ value used here was derived from the magnetic field observations (using, for instance, Bagchi et al. 1998; Ge & Owen 1993, 1994; Ensslin & Biermann 1998; Ruy & Biermann 1998). We assume: frozen-in field; spherical symmetry for the flow and the cluster itself; and that at $`r>r_c`$ (the cooling radius, see below), the magnetic field is isotropic, i.e., $$B_r^2=B_t^2/2=B^2/3$$ and $`l_r=l_tl`$ (where $`B_r`$ and $`B_t`$ are the radial and transversal components of the magnetic field $`B`$ and $`l_r`$ and $`l_t`$ are the coherence length of the large-scale field in the radial and transverse directions). In order to calculate $`B_r`$ and $`B_t`$ for $`r<r_c`$ we modified the calculation of the magnetic field of Soker & Sarazin (1990) by considering an inhomogeneous cooling flow (i.e. $`\dot{M}_i\dot{M}`$ varies with $`r`$). Therefore, the two components of the field are then given by $$\frac{D}{Dt}\left(B_r^2r^4\dot{M}^2\right)=0$$ and $$\frac{D}{Dt}\left(B_t^2r^2u^2\dot{M}^1\right)=0.$$ In our models we take as reference radius the cooling radius $`r_c`$. In fact we modify the geometry of the field when and where the cooling time comes to be less than $`10^{10}\mathrm{yr}`$ (usually adopted as the condition for the development of a cooling flow). Therefore, our condition to assume a non-isotropic field is $`t_{coo}3k_BT/2\mu m_H\mathrm{\Lambda }(T)\rho 10^{10}\mathrm{yr}`$. After the formation of the cooling flow, in the inner regions of the ICM, the magnetic field geometry is changed, following the enhancement of the radial component of the field, due to the enhancement of the density. ## 3 Models and Results In this section we present the results of our models. There are four parameters to consider in each one of the models: $`\sigma `$, the cluster velocity dispersion; $`\rho _0`$, the initial average mass density of the gas; $`a`$, the cluster core radius; and $`\beta _0`$, the initial magnetic to thermal pressure ratio. We adopted the removal efficiency $`q=1.5`$. The most important results of our models are shown on figures we describe below, for which we assume: $`\sigma =1000\mathrm{km}\mathrm{s}^1`$ and $`a=250\mathrm{kpc}`$. First of all, the evolution we follow here is characteristic of cooling flow clusters and in this scenario we discuss the evolution of the basic thermodynamics parameters. Considering the overall characteristics of our models, we will compare the results with the very recent study based on ROSAT observations of the cores of clusters of galaxies, by Peres et al. (1998), focusing on cooling flows in a X-rays flux-limited sample (containing the brightest 55 clusters over the sky in the $`210\mathrm{keV}`$ band). Comparing the present models with Peres et al. (1998) deprojection results, we see that the central cooling time here adopted as our cooling flow criterion, e.g. $`t_{coo}10^{10}\mathrm{yr}`$, is typical for a fraction between $`70\%`$ and $`90\%`$ of their sample. They also discuss briefly the cooling flow age, remembering that in hierarchical scenarios for the formation of structures in the Universe, clusters are formed by smaller substructures by mergers, and therefore the estimation of the cooling flows ages (and the cluster ages themselves) is complicated. Anyway they determine the fraction of cooling flow clusters in their sample considering a factor of two in the ages and concluding that the fraction do not vary that much (from $`13\mathrm{Gyr}`$ to $`6\mathrm{Gyr}`$, the fraction varies from $`70\%`$ to $`65\%`$). This allow us conclude that our models, which present cooling flows since the cluster has the age of $`79\mathrm{Gyr}`$, are typical for their sample. As a matter of fact, the time in which the cooling flow structure is formed depends strongly on the initial density we adopted. For models with $`\rho _0=1.25\times 10^{28}\mathrm{g}\mathrm{cm}^3`$ it rises on $`9\mathrm{Gyr}`$, while the models with $`\rho _0=1.5\times 10^{28}\mathrm{g}\mathrm{cm}^3`$ have it formed on $`7\mathrm{Gyr}`$. We will come back to this point later while analyzing the field anisotropy. The characteristics of our models are summarized using four typical set of initial parameters, and discussing some details which came up of the study of a larger grid of parameters. Therefore, each model is characterized by its position in the $`(\rho _0,\beta _0)`$ parameter space: model I $`(\rho _0=1.5\times 10^{28}\mathrm{g}\mathrm{cm}^3,\beta _0=10^2)`$; model II $`(\rho _0=1.5\times 10^{28}\mathrm{g}\mathrm{cm}^3,\beta _0=10^3)`$; model III $`(\rho _0=1.25\times 10^{28}\mathrm{g}\mathrm{cm}^3,\beta _0=10^2)`$; and model IV $`(\rho _0=1.25\times 10^{28}\mathrm{g}\mathrm{cm}^3,\beta _0=10^3)`$. Figure 1 shows the evolution of density and temperature profiles corresponding to model I, from which the presence of the cooling flow on later stages of the ICM evolution and at inner regions is remarkable if one notices the steep gradients of these quantities. In order to better understand how the magnetic field geometry is modified after the cooling flow formation, e.g., after the steepness on the temperature and density gradients, we follow the evolution of the degree of anisotropy, using the concepts previously defined on Section 2, concerning the geometry of the magnetic field. Hereafter we called ‘degree of anisotropy’ the ratio $`B_t/B_r`$, noting that for the isotropic case it results $`\sqrt{2}`$ and the more anisotropic the field geometry the smaller is this ratio. Therefore, we present on Figure 2 the evolution of the degree of anisotropy since $`3.3\mathrm{Gyr}`$, comparing models I and III, in which one can see, clearly, that the anisotropy begins decreasing on earlier times for models with higher $`\rho _0`$ (model I) than for the ones with lower values of $`\rho _0`$ (model III). From Figure 2 we are allowed to conclude that the degree of anisotropy can be seen as a sensor of the presence of the cooling flow. In another words, the change in the degree of anisotropy can be used as another criterion to indicate the epoch, on the ICM evolution, in which the cooling flow appear. These results can also be discussed in the light of some observational works in which the limits to the magnetic field strength on large and small scales of the cooling flow clusters are given. Following such a kind of observations, as previously seen in the introduction section, we chose two values of magnetic field strength derived by the authors below. The first one is presented in Bagchi et al. (1998) who estimated, from inverse Compton X-ray emission with the synchrotron emission plasma, a cluster-scale ($`700\mathrm{kpc}`$) magnetic field strength of $`(0.95\pm 0.10)\mu \mathrm{G}`$ for Abell 85 (a cooling flow cluster with a central dominant cD galaxy and $`\dot{M}`$ $`100\mathrm{M}_{}/\mathrm{yr}`$). The second one is presented in two papers of Ge & Owen (1993, 1994), in which they present and discuss rotation measures and the related intensity of the magnetic field, giving a range of this intensity at scales of $`10\mathrm{kpc}`$. Therefore, our results for the magnetic field strength and also for pressures, on large and small scales, are compared to the chosen observed ones, in Figures 3 and 4. Reminding that the time on which the cooling flow arises is closely related to $`\rho _0`$, one can expect distinct results on the evolution of the field intensity from, for instance, model I to model III. However this evolution can be better explained comparing model I (Figure 3) to model II (Figure 4), since these two models have the same initial density but distinct $`\beta _0`$. Our best model, in terms of the magnetic field strength compared with observations, is model I $`(\rho _0=1.5\times 10^{28}\mathrm{g}\mathrm{cm}^3,\beta _0=10^2)`$. From Figure 3 it is possible to see that on scales of $`700\mathrm{kpc}`$ the magnetic field expected for the model is higher than the observed one (considering, of course, the profile correspondent to redshift zero, or evolution times on the order of $`1314\mathrm{Gyr}`$), while on scales of $`10\mathrm{kpc}`$ the model gives a value lower than the observed one. Meanwhile, at least on scales of $`700\mathrm{kpc}`$, the situation is inverted if one takes a look on Figure 4, for which $`\rho _0=1.5\times 10^{28}\mathrm{g}\mathrm{cm}^3`$, but $`\beta _0=10^3`$. Given the uncertainties characteristics of the observations, we can say that our models are in agreement with the magnetic field estimations available. On Figures 5 and 6 we show the magnetic and thermal pressures evolution, or in another words, $`\beta `$-evolution, for models I and II respectively, on later times of the ICM evolution, in order to analyze when and where magnetic pressure reaches equipartition. Obviously the magnetic pressure is compatible with the magnetic field intensities and may be compared to the values determined by, for instance, Bagchi et al. (1998), $`p_B=B^2/8\pi 4\times 10^{14}\mathrm{erg}\mathrm{cm}^3\mathrm{s}^1`$, at scales of $`700\mathrm{kpc}`$, on the present time. From the analysis of the magnetic pressures expected from our models it is clear that they agree, as well as the magnetic field strength, with the observations. Here again model I appears being the best one, with $`(\beta _0=10^2)`$, but the values expected from model II are not far away from the observed ones as well. Noting also that magnetic pressure and/or magnetic intensity does not change very much after $`12\mathrm{Gyr}`$, for both cases. Results presented on Figures 3 - 6 would indicate that we should adopt an intermediate initial value for $`\beta `$ (like $`\beta _0=5\times 10^3`$) in order to obtain a magnetic field intensity in better agreement with the observations, at least on scales of $`700\mathrm{kpc}`$. Nevertheless such an exercise should not solve the match of models and observations on smaller scales, since $`\beta _0`$ $`5\times 10^3`$ should decrease magnetic pressure on scales of $`10\mathrm{kpc}`$, at the present time, as a result of the present modelling assumptions (see Figure 4). Other proposals for the amplification of the magnetic field in the center of the cooling flow clusters are: i) rotational driven mechanisms, in which the twisting of the magnetic flux tubes and/or the operation of fast $`\alpha \omega `$ dynamo are the responsible for the increase of the magnetic strength (Godon et al. 1998); ii) turbulence induced amplification (Eilek 1990; Mathews & Brighenti 1997). However, these processes can not account for the strong magnetic fields observed in the center regions, confirming the expectations previously discussed by authors like Goldshmidt & Rephaeli (1993) and Carvalho (1994). ## 4 Discussions and Conclusions The present models are in many aspects similar to the one of Soker & Sarazin (1990). However there are two important differences between our model and theirs: i) they take into account only small-scale magnetic field effects; and ii) they consider homogenous cooling flow. Since we consider inhomogeneous cooling flow (i.e. $`\dot{M}`$ decreases with decreasing $`r`$) the amplification of $`B`$ is smaller in our models. As a matter of fact the magnetic pressure reaches equipartition only at radius as small as $`1\mathrm{kpc}`$ (model I) or $`0.5\mathrm{kpc}`$ (model II), because the central increase of the $`\beta `$ ratio is moderate in our model. Our more realistic description of the field geometry is crucial. This implies that the effect of the magnetic pressure on the total pressure of the intracluster medium, even on regions as inner as few kpc, is small. Tribble (1993) studying the formation of radio haloes in cooling flow clusters from the point of view of the cluster evolution via mergers, suggested typical magnetic field strengths of $`1\mu \mathrm{G}`$. In addition, Zoabi et al. (1996), studying a completely different characteristic of the ICM (magnetic fields on the support of X-rays clumps and filaments), adopted the usually assumed magnetic to pressure ratio, at few scales of $`1020\mathrm{kpc}`$, of $`0.1`$, and following a simple geometry of the field in which it is amplified by the radial inflow, this ratio become $`1`$ at $`5\mathrm{kpc}`$. Again our results are more or less compatible with the above ones (for the cluster scale magnetic field), but the equipartition condition is reached at smaller scales. There are a number of papers discussing heating processes on the inner part of the cooling flow clusters, in particular mechanisms to power the emission lines of optical filaments, which use the magnetic energy transformed in optical emission via magnetic reconnection (Jafelice & Friaça 1996) or dissipation of Alfvén waves (Friaça et al. 1997). These works are based in the enhancement of the magnetic pressure on scales smaller than $`10\mathrm{kpc}`$, where the filaments are observed (Heckman et al. 1989). Finally, our results suggest that the effect of the magnetic fields on the ICM dynamics can be relevant only on very small scales: $`\beta 10^1`$, $`r10\mathrm{kpc}`$, and $`\beta 1`$, $`r1\mathrm{kpc}`$, depending on the model adopted (see Figure 7). From Figure 7 one can see quite clearly that the equipartition condition is reached at smaller radii for models in which $`\beta _0`$ is equal to $`10^3`$ (model II and model IV), emphasizing the agreement between our models, another theoretical models, and observations. It is also quite relevant noting that the general agreement of our models and the available data can be emphasized by the fact that observations give us only limits on the magnetic field intensities. In the case of rotation measures the limit is the upper one, in contrast with the data coming from inverse Compton scattering which give the lower limit of this quantity. Therefore, from our best model (model I, see Figure 3) the expected field intensity is lower than the observed value (provided via rotation measures), on scales of $`10\mathrm{kpc}`$, and higher than the field intensity derived from X-ray inverse Compton scattering, on larger scales ($`700\mathrm{kpc}`$). That the discrepancy found in the determination of mass, from gravitational lensing and from X-rays observations (Loeb & Mao, 1994; Loewenstein, 1994; Miralda-Escudé & Babul, 1995), and in the mass distribution of the gravitating matter, mostly dark matter (Eyles et al. 1991), can be consequences of the standard description of the ICM, in which it is assumed hydrostatic equilibrium driven by thermal pressure (Fabian 1994), is a subject of discussion. Allen (1998) argued that, at least for cooling flow clusters, the above discrepancy is resolved and, therefore, the effect of non-thermal pressures on the hydrostatic equilibrium of these systems could be completed discarded. However, it is important to point out that the radius in which magnetic pressure reaches equipartition is much smaller than the core or arc radii obtained by Allen in his analysis ($``$ 50 kpc, in average), implying that, despite Allen’s results, at smaller scales the non-thermal pressures can be important. Theoretical models, like the one here presented, point out that magnetic pressure does affect the hydrostatic equilibrium of the ICM, but only in the inner radius, as small as $`1\mathrm{kpc}`$. In addition, it is important to remind that there are other sources of non-thermal pressures that could be considered jointly to the magnetic pressure before to close the discussion on whether or not non-thermal pressures can explain the discrepancies on the mass estimations of the galaxy clusters. ### Acknowledgements One of the authors (D.R.G.) would like to thank the Brazilian agency FAPESP (97/05246-3) for support, and the other author (A.C.S.F.) would like to thank the Brazilian agency CNPq for partial support. We also would like to acknowledge partial support by Pronex/FINEP (41.96.0908.00).
no-problem/9906/cond-mat9906029.html
ar5iv
text
# Interplay of disorder and nonlinearity in Klein-Gordon models: Immobile kinks ## I Introduction An extensive research work on the static and transport properties of nonlinear excitations in various soliton-bearing disordered systems has been undertaken in the last decade (see Refs. and ). It is well known that when taken separately both the nonlinearity and disorder contribute to the localization effects, the character of localization being however essentially different. The nonlinearity results in the possibility of existence of nonlinear localized excitations usually referred to as solitons that are rather robust and can propagate through the system undistortedly. At the same time the disorder (in linear systems) evokes the Anderson localization, which manifests itself in the behavior of the transmission coefficient of a plane wave decaying exponentially with the system width. These two localization mechanisms are competitive to some extent; taken together, the nonlinearity and disorder may lead to a number of qualitatively new effects, namely, the transmission coefficient tending to zero with increasing the system length according to a power law rather than exponentially (which would be otherwise the property of a linear system); there can arise a multistability in the wave transmission through a disordered slab; excitations in highly nonlinear or multidimensional nonlinear Schrödinger (NLS) systems (which would either disperse or collapse otherwise) can be stabilized by disorder. In the present paper we study the static properties of a one-kink solution (or, equivalently, of a diluted kink gas) of disordered Klein-Gordon models where the disorder is assumed to be a $`\delta `$-correlated Gaussian spatial noise. The disorder of the kind is akin, for example, to Josephson junctions, where it is caused by the fluctuations of the gap between two superconductor plates. Quite recently Mints proposed a similar model with randomly alternating critical current density to account for a self-generated magnetic flux observed experimentally. We discuss his results in more detail in the Conclusion. In general, Klein-Gordon models have been repeatedly studied for the disorder represented by a lattice of $`\delta `$-like impurity potentials with random positions of the impurities and either equal or randomly distributed intensities. It was found out that in a number of cases the kink dynamics in disordered systems could be adequately described within the framework of the collective coordinates approach (Refs. , and references therein). This is the background of our restricting ourselves to the said approach in the scope of the present paper providing however at each step the validation of analytical results comparing them to the numerical simulations of the original system. We emphasize that in opposite to, for example, Refs. and , we use Rice’s collective coordinates approach with the kink width being the variational parameter. A similar approach has been recently applied for two other one-dimensional (1D) systems: Bussac et al. investigated the effects of the polaron ground state in a deformable chain, while Christiansen et al. considered the stabilization of nonlinear excitations by disorder in the NLS model. It should be indicated that investigation of these two (in fact closely related) systems leads to one and the same result: the width of stationary solitons decreases with growing intensity of the disorder. The importance of this conclusion resides in its prediction that the disorder can stabilize otherwise unstable solitons in 2D and 3D NLS models. Quite recently this prediction was borne out numerically for the 2D case. It must be emphasized that for NLS models the conclusion does not depend on the averaging procedure: one can equally perform averaging either on absolute ground states or over all local minima of the effective random potential with equal weights. We show that it is not the case for the Klein-Gordon models: their properties exhibit strong dependence on the assumptions as to the statistical distribution of kinks over the minima of the effective random potential. For the purely dynamical problem these statistics are left beyond consideration and should be thus imposed as an additional assumption. We use Jaynes’s maximum entropy inference for this purpose. The outline of the paper is the following. In Sec. II we present the model and derive the equations for the collective coordinates of the kink taking into account the disorder via the effective random potential. In Sec. III we investigate both analytically and numerically the case of immobile kinks and demonstrate the existence of the crossover between monotonic and nonmonotonic dependence of the average kink width as a function of the disorder intensity. In Sec. IV we summarize the exposed results. ## II Collective coordinates approach We consider a Klein-Gordon (KG) model in the presence of space disorder. The Hamiltonian of the system has the form $`H={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑x\left\{{\displaystyle \frac{1}{2}}(\varphi _t^2+\varphi _x^2)+[1\eta ϵ(x)]\mathrm{\Phi }(\varphi )\right\},`$ (1) where the subscripts stand for partial derivatives with respect to the indicated variables and units are chosen so that the Hamiltonian is already in the scaled form. The potential $`\mathrm{\Phi }(\varphi )`$ has the form $$\mathrm{\Phi }(\varphi )=1\mathrm{cos}\varphi $$ (2) for the sine-Gordon (SG) model, and $$\mathrm{\Phi }(\varphi )=\frac{1}{4}(\varphi ^21)^2$$ (3) for the $`\varphi ^4`$ model. We assume that $`ϵ(x)`$ is delta-correlated spatial disorder $$ϵ(x)ϵ(x^{})=\delta (xx^{}),$$ (4) (the brackets $`\mathrm{}`$ denote averaging over all realizations of the disorder) with the Gaussian distribution $$p[ϵ(x)]=\frac{1}{\sqrt{\pi }}\mathrm{exp}[ϵ^2(x)].$$ (5) We have studied both SG (2) and $`\varphi ^4`$ (3) models. Although properties of these two models show many similarities, they also exhibit a number of interesting distinctions related, in particular, to the existence of a breather state in the SG model and of Rice’s internal mode in the $`\varphi ^4`$ model. So, it was important to compare the effects of disorder in both these models. But since the qualitative features of the exposed models turned out to coincide in the scope of our present investigations, we dare not overload the paper with unnecessary repetitions and restrict ourselves with presenting in detail the sine-Gordon model only, keeping in mind although that every stage of the calculations applies for the $`\varphi ^4`$ model as well. The SG system is governed by the equation of motion $$\varphi _{tt}\varphi _{xx}+[1\eta ϵ(x)]\mathrm{sin}\varphi +\gamma \varphi _t=0,$$ (6) where the damping term with the damping constant $`\gamma `$ has been included. It is well known that in the absence of disorder and damping ($`\eta =\gamma =0`$) Eq. (6) is completely integrable and possesses a topologically stable solution in the form of a kink given by $$\varphi _K(x,t)=4\mathrm{arctan}\mathrm{exp}\left(\frac{xX(t)}{L(t)}\right),$$ (7) where $`X(t)=X_0+vt`$ is the kink coordinate, $`v`$ is its velocity, and $`L=\sqrt{1v^2}`$ is the kink width. In the general case of Eq. (6) for a number of situations the kink emission is exponentially small, so that the kink dynamics can be studied by the collective coordinate approach. In the framework of this approach the variables $`X(t)`$ and $`L(t)`$ are understood as time-dependent variational parameters. Inserting Eq. (7) into Hamiltonian (1) as a trial function, we obtain the effective Hamiltonian $$H_{eff}=\frac{L}{16}p_X^2+\frac{3L}{4\pi ^2}p_L^2+U(L)+V(\{ϵ\},L,X),$$ (8) with momenta $$p_X=\frac{8}{L}\frac{dX}{dt},p_L=\frac{2\pi ^2}{3L}\frac{dL}{dt}.$$ (9) Here $$U(L)=\frac{4}{L}+4L$$ (10) is the potential function in the case of no disorder and $`V(\{ϵ\},L,X)=2\eta {\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑xϵ(x)\text{sech}^2\left({\displaystyle \frac{xX}{L}}\right)`$ (11) is the effective random potential arising because of the disorder term. Then, taking into account the damping, we arrive at the following equations of motion: $$\frac{dp_X}{dt}+\gamma p_X+\frac{d}{dX}V(\{ϵ\},L,X)=0,$$ (12) $`{\displaystyle \frac{dp_L}{dt}}`$ $`+`$ $`\gamma p_L+{\displaystyle \frac{3p_L^2}{4\pi ^2}}+{\displaystyle \frac{p_X^2}{16}}`$ (13) $`+`$ $`{\displaystyle \frac{d}{dL}}[U(L)+V(\{ϵ\},L,X)]=0.`$ (14) In the following section we solve approximately these equations of motion for immobile kinks and compare the results to the results of direct numerical integration of Eq. (6). ## III Results for immobile kinks As a consequence of the damping $`\gamma `$ the kink will eventually stop at some stable or metastable stationary position along the system. Here we do not consider this transient stage and assume that the kink is already immobile \[$`dp_L/dt=p_L=p_X=0`$\]. In this case the equations of motion (12) and (13) take on the form $$\frac{d}{dX}V(\{ϵ\},L,X)=0,$$ (15) $$\frac{d}{dL}[U(L)+V(\{ϵ\},L,X)]=0.$$ (16) Considering the center-of-mass motion described by Eq. (12) we observe that for each realization of the random potential $`ϵ(x)`$ the stable stationary position $`X=X_m(\{ϵ\},L)`$ of the kink is defined by the point where $`V(\{ϵ\},L,X)`$ has a minimum with respect to $`X`$. Thus we can now insert the value $`X=X_m(\{ϵ\},L)`$ into Eq. (16) and, solving the resulting equation $$L=\left(1+\frac{1}{4}\frac{d}{dL}V(\{ϵ\},L,X_m)\right)^{1/2},$$ (17) get the value of the stationary kink width $`L(\{ϵ\})`$ at given position of the kink. Encountering a similar problem for the case of NLS system, Christiansen et al. invoked the mean-field approximation $$\frac{d}{dL}V(\{ϵ\},L,X_m)\frac{d}{dL}V(\{ϵ\},L,X_m).$$ (18) Further the estimation of the quantity $`V`$ was performed using Rice’s averaging theorem. However, being rather good for the NLS model the mean-field approximation (18) fails for the KG models. Thereby we were forced to use a more precise averaging procedure calculating $`dV/dL`$ directly. Expanding Eq. (17) into series up to the second order in $`\eta `$, $$L1\frac{1}{8}\frac{dV}{dL}+\frac{3}{128}\left(\frac{dV}{dL}\right)^2,$$ (19) and solving by iterations we get after averaging \[which by means of Eq. (4) can be performed for the terms containing $`\eta ^2`$ exactly\] the average kink width $$L_{var}L(\{ϵ\})1+\frac{\eta }{2}\lambda (\{ϵ\},X_m)+\frac{\pi ^2}{180}\eta ^2,$$ (20) where the averaging of the function $$\lambda (\{ϵ\},X)=_{\mathrm{}}^{\mathrm{}}𝑑xϵ(x)(xX)\frac{\mathrm{sinh}(xX)}{\mathrm{cosh}^3(xX)}$$ (21) is performed over all realizations of the disorder in the points $`X=X_m`$ in which the potential $$V(\{ϵ\},L,X)2\eta \mu (\{ϵ\},X)$$ (22) with $$\mu (\{ϵ\},X)=_{\mathrm{}}^{\mathrm{}}𝑑xϵ(x)\text{sech}^2(xX)$$ (23) takes on its minima on $`X`$. Thus we arrive at the problem of performing the average of $`\lambda (\{ϵ\},X_m)`$ over the minima $`X_m`$ of the function $`\mu (\{ϵ\},X)`$. It is convenient for later use to perform this averaging in two steps calculating at the outset $$\mathrm{\Lambda }(\stackrel{~}{\mu })=_{\mathrm{}}^{\mathrm{}}\stackrel{~}{\lambda }P_l(\stackrel{~}{\lambda }\stackrel{~}{\mu })𝑑\stackrel{~}{\lambda }$$ (24) and thereafter $$\lambda (\{ϵ\},X_m)=_{\mathrm{}}^{\mathrm{}}\mathrm{\Lambda }(\stackrel{~}{\mu })P_m(\stackrel{~}{\mu })𝑑\stackrel{~}{\mu }.$$ (25) Here $`P_l(\stackrel{~}{\lambda }\stackrel{~}{\mu })`$ is the conditional probability that $`\lambda (\{ϵ\},X_m)`$ has the value $`\stackrel{~}{\lambda }`$ if $`\mu (\{ϵ\},X_m)`$ equals to $`\stackrel{~}{\mu }`$. Correspondingly, $`\mathrm{\Lambda }(\stackrel{~}{\mu })`$ is the value of $`\lambda (\{ϵ\},X_m)`$ averaged over all realizations of the disorder for which $`\mu (\{ϵ\},X_m)`$ is equal to $`\stackrel{~}{\mu }`$. It is difficult to calculate $`\mathrm{\Lambda }(\stackrel{~}{\mu })`$ analytically but numerical simulations show (see Fig. 1) that up to very good accuracy the dependence $`\mathrm{\Lambda }(\stackrel{~}{\mu })`$ is linear: $$\mathrm{\Lambda }(\stackrel{~}{\mu })0.344\stackrel{~}{\mu }0.500.$$ (26) Substituting it into Eq. (25) we obtain that $$\lambda (\{ϵ\},X_m)0.344\mu (\{ϵ\},X_m)0.500,$$ (27) where $$\mu (\{ϵ\},X_m)=_{\mathrm{}}^{\mathrm{}}\stackrel{~}{\mu }P_m(\stackrel{~}{\mu })𝑑\stackrel{~}{\mu }$$ (28) is the average value of the function $`\mu (\{ϵ\},X)`$ over its minima $`X_m`$. Here the probability density $`P_m(\stackrel{~}{\mu })`$ is a product of two factors. The first one is the probability density that some arbitrary chosen minimum $`X_m`$ of the function $`\mu (\{ϵ\},X)`$ will be equal to $`\stackrel{~}{\mu }`$. We denote this probability density as $`p_{min}(\stackrel{~}{\mu })`$. The second factor is the conditional probability that if the minimum is equal to $`\stackrel{~}{\mu }`$ it will be actually occupied by the kink. It is evident that in a real system the kink is more likely to occupy the deeper minimum than the shallow one. So to be consequential one must ascribe to every minimum of the function $`\mu (\{ϵ\},X)`$ some probability weight and average taking into account those probabilities. But the values of these probabilities are in general determined by the whole prehistory of the kink. These values are not contained in the dynamical equations of motion that state only that the kink should take on some minimum regardless to its depth. It would be a cumbersome problem to calculate them appropriately. That is why two limit cases are in general considered: either the kink seats itself into the deepest well or it rather occupies any of them with equal probabilities. As it was already remarked in the Introduction, both limit cases lead to qualitatively the same results for the NLS model. It is not the case of KG models where, as it will be shown later, different assumptions as to the a priori weights of minima lead to qualitatively different behavior of the kink width. Since we are merely lacking information sufficient enough to reconstruct these weights in an objectivistic fashion, a remedy would be Jaynes’s maximum entropy inference, according to which the simplest self-consistent unbiased choice is to assume that the kink will occupy a potential well corresponding to the minimum $`X_m`$ of the function $`\mu (\{ϵ\},X)`$ \[for given profile $`ϵ(x)`$\] with probability proportional to $`e^{\beta \mu (\{ϵ\},X_m)}`$ thus introducing an additional parameter $`\beta `$ (following Jaynes we shall call it a conjugate parameter). By this expedience we present some natural interpolation covering two mentioned limit cases: of equiprobable distribution ($`\beta =0`$) and of averaging over the deepest minima only ($`\beta \mathrm{}`$). So we can write $$P_m(\stackrel{~}{\mu })=\frac{1}{𝒵}e^{\beta \stackrel{~}{\mu }}p_{min}(\stackrel{~}{\mu }),$$ (29) where $`𝒵(\beta )={\displaystyle _{\mathrm{}}^{\mathrm{}}}𝑑\stackrel{~}{\mu }e^{\beta \stackrel{~}{\mu }}p_{min}(\stackrel{~}{\mu })`$ (30) plays part of the partition function. To calculate $`p_{min}(\stackrel{~}{\mu })`$ we follow Ref. and make use of the Rice’s averaging theorem \[valid for the case, well attested by the numerics, of $`\mu (\{ϵ\},X)`$ being a stationary centered Gaussian process\] stating that the probability density of some given minimum of the function $`\mu (\{ϵ\},X)`$ to be equal to $`\stackrel{~}{\mu }`$ is $$p_{min}(\stackrel{~}{\mu })=\frac{1}{\sqrt{2\pi M_0}}\sigma (\frac{\stackrel{~}{\mu }}{\sqrt{2M_0}},\sqrt{1\frac{M_2^2}{M_0M_4}}),$$ (31) where the function $`\sigma (y,\kappa )`$ $`=`$ $`\kappa e^{(y/\kappa )^2}`$ (32) $``$ $`2\sqrt{1\kappa ^2}ye^{y^2}{\displaystyle _{(y/\kappa )\sqrt{1\kappa ^2}}^{\mathrm{}}}e^{t^2}𝑑t,`$ (33) and the spectral momenta $`M_0=[\mu (\{ϵ\},X)]^2={\displaystyle \frac{4}{3}},`$ (34) $`M_2=[\mu _X(\{ϵ\},X)]^2={\displaystyle \frac{16}{15}},`$ (35) $`M_4=[\mu _{XX}(\{ϵ\},X)]^2={\displaystyle \frac{64}{21}},`$ (36) were introduced. Hence the partition function takes on the form $$𝒵(\beta )=\frac{1}{\sqrt{\pi }}_{\mathrm{}}^{\mathrm{}}𝑑xe^{\sqrt{8/3}\beta x}\sigma (x,\frac{3}{5}\sqrt{2})$$ (37) and can be expanded into series in $`\beta `$ yielding the average value $`\mu (\{ϵ\},X_m)={\displaystyle \frac{d}{d\beta }}\mathrm{ln}𝒵(\beta )`$ (38) $`={\displaystyle \frac{1}{5}}\sqrt{{\displaystyle \frac{7\pi }{6}}}{\displaystyle \frac{2}{75}}(647\pi )\beta +𝒪(\beta ^2).`$ (39) And now, substituting it into Eqs. (27) and (20) we obtain for the average kink width $`L_{var}1+(0.193\beta 0.059)\eta +{\displaystyle \frac{\pi ^2}{180}}\eta ^2.`$ (40) Thus it is seen that there is a qualitative change of the kink width behavior as function of the disorder intensity $`\eta `$ according to whether the value of conjugate parameter $`\beta `$ is below or above some critical value $`\beta _{cr}0.3`$. At small $`\beta `$ the average kink width is a nonmonotonic function of the disorder intensity $`\eta `$: it decreases at small intensities but starts to increase thereafter. This result is well attested by the direct numerical calculations of stationary kink solutions of the initial equation of motion (6). In Fig. 2 we compare the analytical prediction given by Eq. (40) with the numerical results for the case of equiprobable distribution of the kinks over the potential wells ($`\beta =0`$). The numerical results have been obtained as an average of 1000 realizations of the disorder. Two different expressions for the kink width were calculated: $$L_{cos}=\frac{1}{4}_{\mathrm{}}^{\mathrm{}}𝑑x\{1\mathrm{cos}[\varphi (x)]\},$$ (41) and $$L_{der}=8\left\{_{\mathrm{}}^{\mathrm{}}𝑑x\left[\frac{d\varphi (x)}{dx}\right]^2\right\}^1.$$ (42) From the point of view of the collective coordinate approach $`L_{cos}`$ and $`L_{der}`$ should coincide with the value of $`L`$ introduced in Eq. (7). Indeed, as it is seen from the figures, it does take place for $`0\eta 0.2`$; thus, these are limits where the collective coordinate approach works well. Going on with the analysis of Eq. (40) we see that at big values of the conjugate parameter ($`\beta >\beta _{cr}`$) when the kink rather occupies deep potential wells, the kink width should grow monotonically with the disorder. But it is evident that in this case the analytical approach discussed above is applicable to systems of infinite length only. For the finite system the case of big $`\beta `$ represents the situation when the kink sits in the deepest potential well. Obviously its average depth essentially depends on the system length. We can make an estimation of this dependence drawing on the formula for the average number of minima of the function $`\mu (\{ϵ\},X)`$ on an interval $`R`$ whose values lie below some $`\stackrel{~}{\mu }`$: $$N_{min}=\frac{R}{2\pi }\sqrt{\frac{M_2}{M_0}}e^{\stackrel{~}{\mu }^2/(2M_0)}.$$ (43) Inserting there $`N_{min}=2`$ ($`\stackrel{~}{\mu }`$ is an absolute minimum on the interval $`R`$ but not on the longer interval) one can estimate its average value $`\mu (\{ϵ\},X_{abs.m})`$ $``$ $`\left(2M_0\mathrm{ln}\left\{{\displaystyle \frac{R}{4\pi }}\sqrt{{\displaystyle \frac{M_2}{M_0}}}\right\}\right)^{1/2}`$ (44) $`=`$ $`\left({\displaystyle \frac{8}{3}}\mathrm{ln}\left\{{\displaystyle \frac{R}{2\sqrt{5}\pi }}\right\}\right)^{1/2}`$ (45) and, substituting it into Eqs. (20) and (27) one can find that the average kink width equals $$L_{var}1+\left(0.28\mathrm{ln}^{1/2}\left\{\frac{R}{2\sqrt{5}\pi }\right\}0.25\right)\eta +\frac{\pi ^2}{180}\eta ^2.$$ (46) It is seen that for finite-size systems the character of dependence $`L_{var}(\eta )`$ depends on the size of the system $`R`$. It is interesting that even for the case of averaging over the absolute minima considered here, the function $`L_{var}(\eta )`$ grows monotonically with $`\eta `$ only for the systems that are large enough ($`R7.5`$). The reason is that for a small system the number of potential wells of the effective random potential is too small to yield the average over absolute minimum that would be essentially smaller than the average value calculated over all minima. Indeed, Figs. 3 and 4, in which we compare Eq. (46) to the results of the numerical calculations for $`R=15`$ and $`R=5`$, lend support to the validity of the approach leading to Eq. (46). One can see from these figures that the average kink width grows monotonically with $`\eta `$ for $`R=15`$ but is nonmonotonic (similar to the case depicted on Fig. 2) for small $`R=5`$. But in this latter case the boundary conditions become very important and most likely they are responsible for the difference between Figs. 2 and 4. ## IV Conclusion In the paper we consider the Klein-Gordon models with the $`\delta `$-correlated spatial disorder and investigate both analytically and numerically the width of immobile kinks as a function of the intensity of disorder. The analytical collective coordinates approach is based on Rice’s averaging theorem from the theory of random processes as well as on the maximum entropy inference proposed by Jaynes. We have shown that the properties of the kinks exhibit strong dependence on the assumptions as to their statistical distribution over the minima of the effective random potential. Namely, there exists a crossover from monotonically increasing (when a kink occupies the deepest potential well) to the nonmonotonic (at equiprobable distribution of kinks over the potential minima) dependence of the average kink width as a function of the disorder intensity. We have shown also that the same crossover may take place with the changing size of the system: the average kink width monotonically increases for the systems of big size but is nonmonotonic for the small ones. It is interesting to compare the effects of the disorder in the KG model with the effects in the nonlinear Schrödinger (NLS) model. As it was recently shown in Refs. , the $`\delta `$-correlated spatial disorder in the NLS systems creates an additional factor contributing to the decrease of the excitation width. This effect, being insensitive to the manner of the statistical distribution of kinks over the minima of effective random potential, favors the stabilization of excitations in highly nonlinear or multidimensional systems, which would either disperse or collapse otherwise. The stabilizing function of disorder is of no doubt important for practical applications and to elucidate the extent to which it is universal seems to be an intriguing question. The considered example of the KG systems demonstrates that there exists a class of systems, for which, in contrast to the NLS system, the effects of disorder can lead in different cases to diametrically opposed behavior. In the case of the SG model we can consider the term $`\eta ϵ(x)`$ as a change of the Josephson current density due to fluctuations of the thickness of an insulating layer. Quite recently Mints studied such a model to account for a self-generated magnetic flux observed by Mannhart et al. He has shown that in the case of $`\eta <1`$ a state with a self-generated flux exists and can be studied experimentally in the presence of Josephson vortices. However, as is shown in the present paper, the Josephson energy \[equal to $`4L_{cos}`$ in Eq. (41)\] and the magnetic energy \[equal to $`4/L_{der}`$ in Eq. (42)\] of the Josephson vortices are functions of the intensity of fluctuations of the insulating layer thickness. And their contribution into the experimentally observable magnetic flux will strongly (up to the change of a sign) depend on the statistical distribution of the vortices along the Josephson junction. Thus, for the proper description of the problem one must develop a thermodynamic model. ## Acknowledgments We (S.M., Yu.G., and S.Sh.) thank the Department of Theoretical Physics of the Palacký University in Olomouc for the hospitality. S.M. and Yu.G. acknowledge support from the Fund for Development of the MŠMT ČR No. 155/1997 and from the Ukrainian Fundamental Research Fund (Grant No. 2.4/355). E.M. and S.Sh. acknowledge support from Grant No. 202/97/0166 of the GAČR. Partial support from Grant No. 2/4109/97 of the VEGA Grant Agency is also acknowledged.
no-problem/9906/cond-mat9906160.html
ar5iv
text
# Spin Hall effect \[ ## Abstract It is proposed that when a charge current circulates in a paramagnetic metal a transverse spin imbalance will be generated, giving rise to a ’spin Hall voltage’. Similarly, that when a spin current circulates a transverse charge imbalance will be generated, hence a Hall voltage, in the absence of charge current and magnetic field. Based on these principles we propose an experiment to generate and detect a spin current in a paramagnetic metal. \] Consider the ’spontaneous’ or ’anomalous’ Hall effect. In ferromagnetic metals, the Hall resistivity (transverse electric field per unit longitudinal current density) is found to be empirically fitted by the formula $$\rho _H=R_oB+4\pi R_sM$$ (1) (in cgs units), with B the applied magnetic field and M the magnetization per unit volume. $`R_o`$ is the ordinary Hall coefficient and $`R_s`$ the ’anomalous’ Hall coefficient, experimentally found to be generally substantially larger than the ordinary Hall coefficient as well as strongly temperature-dependent. Within models that assume that the electrons giving rise to magnetism in ferromagnetic metals are itinerant, a variety of mechanisms have been proposed to explain the origin of the coefficient $`R_s`$. These include skew scattering by impurities and phonons, and the ’side jump’ mechanism. In early work it was also proposed that the effect will arise in the absence of periodicity-breaking perturbations, but this is generally believed not to be correct. In this paper we will not discuss the origin of the anomalous Hall effect. Rather, we take the existence of the effect in ferromagnetic metals as experimental proof that electrons carrying a spin and associated magnetic moment experience a transverse force when they are moving in a longitudinal electric field, for any of the reasons listed above or others. If there is a net magnetization in the system there will be a magnetization current associated with the flow of electric current, and the transverse force will give rise to a charge imbalance in direction perpendicular to the current flow and hence to an anomalous Hall effect. Consider then the situation where no magnetization exists, that is, a paramagnetic metal or doped semiconductor, or a ferromagnetic metal above its Curie point, carrying a charge current in the $`x`$ direction. The electrons still carry a spin, and the same scattering mechanism(s) that gave rise to the anomalous Hall effect in the magnetic case will scatter electrons with spin up preferentially in one direction perpendicular to the flow of current, and spin down electrons preferentially in the opposite direction. Here we have in mind a slab geometry as usually used in Hall effect experiments, and spin up and spin down directions are defined perpendicular to the plane of the slab. Because there is equal number of spin up and spin down electrons no charge imbalance will result, but we argue that a spin imbalance will: there will be an excess of up spins on one side of the sample and of down spins on the opposite side. The situation is depicted schematically in Figure 1. Although it may appear that if there is spin rotational invariance the spin up and down directions are not well defined, we argue that the slab geometry naturally defines such directions. The effect can be simply understood as arising from spin-orbit scattering. Consider a ’beam’ of unpolarized electrons incident on a spinless scatterer, with potential $$V=V_c(r)+V_s(r)\stackrel{}{\sigma }\stackrel{}{L}$$ (2) with $`\stackrel{}{\sigma }`$ and $`\stackrel{}{L}`$ the electron’s spin and orbital angular momentum. The term $`V_s(r)`$ is the usual spin-orbit scattering potential, proportional to the gradient of the scattering potential. The scattered beam will be spin polarized, with polarization vector $$\stackrel{}{P}_f=\frac{fg^{}+f^{}g}{|f|^2+|g|^2}\widehat{n}$$ (3) where $`\widehat{n}`$ is a unit vector perpendicular to the scattering plane, in direction $`\stackrel{}{k}_i\times \stackrel{}{k}_f`$, with $`\stackrel{}{k}_i,\stackrel{}{k}_f`$ incident and scattered wavevectors. $`f`$ and $`g`$ are spin-independent and spin-dependent parts of the scattering amplitude. $`\widehat{n}`$ has opposite signs for particles scattered to the right and left of the scatterer, hence there is a left-right asymmetry to the spin polarization of the scattered beam, whose sign depends on the sign of $`V_s(r)`$. In the geometry considered here the scattering plane is defined by the plane of the slab, since there is considerably more phase space for scattering in that plane than perpendicular to it. Furthermore, in a crystal prefered spin directions may arise from crystalline anisotropy, and it may be useful to consider a single crystal sample where one such direction is perpendicular to the slab. Finally, a prefered spin direction is also defined by the magnetic field generated by the current flow, which in the slab geometry will point predominantly in the $`+z`$ direction on half of the slab along the $`y`$ direction and in the $`z`$ direction on the other half. This magnetic field will contribute an additional spin imbalance, which may add or substract to the one discussed here depending on the sign of the skew scattering mechanism. We will not be interested in this component of the spin imbalance for reasons discussed below. In the case of the ordinary Hall effect, the charge imbalance results in a difference in the Fermi levels of both sides of the sample, and hence a voltage $`V_H`$ which can be measured with a voltmeter. In the case under discussion here, the Fermi levels for each spin electrons will also be different on both sides of the sample, but the difference will be of opposite sign for both spins. How can one detect this spin voltage $`V_{SH}`$, or equivalently the associated spin imbalance? One possible way would be to measure the difference in magnetization at both edges of the slab. This may perhaps be achieved by using a superconducting quantum interference device microscope with high spatial resolution that can measure local magnetic fields. However, it would be necessary to separate the contributions from the effect discussed here and the magnetic field generated by the current flow, which is likely to be difficult because the latter one should be much larger. A more interesting way follows from the analogy with the ordinary Hall effect. In that case, if the two edges of the sample are connected by a conductor, a charge current will circulate, since the electrons in the connecting conductor do not experience the Lorentz force felt by the electrons in the longitudinal current. Similarly, in our case we argue that when the edges of the sample are connected a spin current will circulate. This spin current will be driven solely by the spin imbalance generated by the skew scattering mechanism(s) affecting the longitudinal current and not by the component of the spin imbalance which is due to the magnetic field originating in the current flow. How does one detect such a spin current? We may use the same principle that allowed the spin imbalance to be created in the first place. When the two edges of the sample are connected and a spin current circulates, a transverse voltage will be generated that can be measured by a voltmeter. The situation is schematically depicted in Figure 2. Let us consider some experimental parameters. First, the width of the sample $`L`$ needs to be smaller than the spin diffusion length $`\delta _s`$. $`\delta _s`$ is the length over which spin coherence is lost due to scattering processes that do not conserve spin. We will rely heavily on the seminal work of Johnson and Silsbee (JS), who studied spin current flow between a ferromagnet and a paramagnet, aluminum. JS estimated $`\delta _s450\mu m`$ at $`T=4.3K`$ and $`\delta _s170\mu m`$ at $`T=36.6K`$ in their $`Al`$ sample, which had residual resistivity ratio (RRR) of about $`1000`$. We will assume for definiteness a sample of $`Al`$ as in the JS experiment, of width $`L=100\mu m`$, which should allow for transverse spin coherent transport at least over the range of temperatures given above. The resistivity of the sample for such RRR will be of order $`\rho =2.7\times 10^3\mu \mathrm{\Omega }cm`$ at low temperatures. The ’magnetization’ associated with the spin up electrons in the sample is $`M=n_{}\mu _B`$, with $`n_{}`$ the density of spin up electrons and $`\mu _B`$ the Bohr magneton. If only up electrons were present, when a longitudinal current density $`j_x`$ flows an anomalous Hall voltage $$V_H=4\pi R_sLj_xn_{}\mu _B$$ (4) would be generated, with $`L`$ the width of the sample and $`R_s`$ the anomalous Hall coefficient. Equation (4) gives also the ’spin Hall voltage’ for spin up electrons that will be generated, and an equal one with opposite sign will result for the spin down electrons in the paramagnetic case. Hence we obtain for the spin Hall voltage $$V_{SH}=2\pi R_sLj_xn\mu _B$$ (5) with $`n`$ the total conduction electron concentration. To obtain an estimate of the magnitude of the effect we will simply assume that $`R_s`$ is the same as the free electron ordinary Hall coefficient of $`Al`$, $`R_o=1/nec=3.45\times 10^{11}m^3/C`$ . As mentioned above, values of the anomalous Hall coefficient tend to be larger than those of the ordinary one. For a current density $`j_x=6\times 10^6A/m^2`$ as used in the JS experiment Eq. (5) yields a spin Hall voltage $`V_{SH}=22nV`$. When we connect the two edges of the sample by a transverse metal strip, a spin current will flow in that strip. Assuming that the resistivity for the spin current is the same as that for the charge current we have for the current for each spin $$j_\sigma =\frac{V_{SH}}{\rho L}$$ (6) which yields for the parameters under consideration $`j_\sigma =8.0\times 10^6A/m^2`$. Assuming the same skew scattering mechanism operating on the transverse sample, the resulting spin Hall voltage due to this spin current is $$V_{SH}^\sigma =4\pi R_slj_\sigma n_\sigma \mu _B$$ (7) with $`l`$ the width of the transverse strip. Now however because spin up and down currents circulate in opposite directions the spin voltages add, giving rise to a real voltage due to the spin current $`V_{SC}=2V_{SC}^\sigma `$, that can be detected by an ordinary voltmeter. The voltage due to the spin current is then, from Eqs. (5), (6) and (7) $$V_{SC}=8\pi ^2R_s^2l\frac{(n\mu _B)^2}{\rho }j_x.$$ (8) Note that the transverse width $`L`$ has dropped out in Eq. (8), because even though it gives larger spin voltage $`V_{SH}`$ it also increases the resistance to the spin current in the transverse direction. Still, a dependence on $`L`$ is implicit in Eq. (8) since when $`L`$ becomes comparable to or larger than the spin diffusion length $`\delta _s`$, $`V_{SC}`$ will decrease. Neither does the thickness of the transverse layer enter in Eq. (8): a thicker layer would increase the spin current but not the current density. For the parameters under consideration here, assuming for example $`l=100\mu m`$ , Eq. (8) yields $`V_{SC}=58nV`$, easily measurable. In the more general case where the transverse strip is of different composition and/or purity than the longitudinal strip Eq. (8) becomes $$V_{SC}=8\pi ^2R_{s1}R_{s2}l\frac{n_1n_2\mu _B^2}{\rho _2}j_x$$ (9) where indices $`1`$ and $`2`$ refer to longitudinal and transverse strips respectively. Figure 3 shows top and side views of the sample envisaged. A thin insulating layer should be deposited on top of the sample (longitudinal strip) of width $`L`$, and small contact areas should be etched to expose the sample surface and allow for metallic contact between the longitudinal and transverse strips. Then, a thin transverse strip of width $`l`$ should be deposited on the insulator such that it also covers the contact areas. The length of the contacts along the $`x`$ direction should be sufficiently small that no significant voltage drop should occur on them due to the longitudinal current, which would be transmitted to the transverse strip. The voltage drop along a contact of length $`l_c`$, $$V_d=l_c\rho j_x,$$ (10) should be substantially smaller than the signal $`V_{SC}`$. For the parameters used as example here, $`V_d=0.2nV`$ for a contact width $`l_c=1\mu m`$. Also, a spurious voltage may arise if the two contacts are not perfectly alligned. Again, if the contacts are offset by $`\mathrm{\Delta }x`$ the magnitude of the spurious voltage will be at most Eq. (10) with $`\mathrm{\Delta }x`$ replacing $`l_c`$, so for our parameters $`V_d1nV`$ if $`\mathrm{\Delta }_x=5\mu m`$. Note also that a smaller resistivity $`\rho `$ both increases the signal voltage Eq. (8) and decreases the spurious voltage Eq. (10). Finally, the resistance of the contacts should be much smaller than the resistance of the transverse strip in order for Eq. (8) to remain valid. This argues for a thin transverse strip (large resistance) and a thin insulating layer (smaller contact resistance along width of layer). It would appear to be simple to achieve a contact resistance at least two orders of magnitude smaller than the transverse strip resistance. Note also that the sign of the expected signal $`V_{SC}`$, as indicated in Fig. 2, is opposite to the voltage $`V_d`$ that would arise from voltage drop across the contacts. As long as the signs of the anomalous Hall coefficients $`R_{s1}`$ and $`R_{s2}`$ are the same, so in particular for $`R_{s1}=R_{s2}`$, the sign of the spin current voltage $`V_{SC}`$ will always be as indicated in Fig. 2, that is, $`V_{SC}`$ drives a current in direction opposite to the primary current $`j_x`$. Thus a measurement of $`V_{SC}`$ for the case where the longitudinal and transverse strips are of the same material provides no information on the sign of $`R_s`$. If the signs of $`R_{s1}`$ and $`R_{s2}`$ are opposite however the sign of the voltage $`V_{SC}`$ would be reversed. Application of a magnetic field in a direction parallel to the plane of the strips will lead to precession of the spins and destruction of the spin polarization for a characteristic field $`(\gamma T_2)^1`$, as discussed by Johnson and Silsbee, with $`\gamma `$ the gyromagnetic ratio and $`T_2`$ the spin relaxation time of the conduction electrons. Thus it would lead to suppression of the spin current voltage. The sensitivity of the signal to an applied magnetic field in the plane of the strips would provide direct evidence for the role of electron spin. For the case of the $`Al`$ sample of JS the signal would be entirely suppressed for magnetic fields in the range of $`20`$ to $`50`$ Gauss depending on temperature. The sign of the expected effect as indicated in Figure 2 is however the same as would be obtained from a ’drag effect’ of the current in the lower strip on the upper strip. Such effects, which may arise from electron-phonon or electron-electron interactions, have been seen in doped semiconductor structures , and they could contribute to the effect discussed here. However, the drag effect does not require contact between the lower and upper strips, should not vary with applied magnetic field in the plane, and should sensitively depend on the thickness of the insulating layer (in our case sensitivity to insulating layer thickness might only enter insofar as it could affect the contact resistance between lower and upper strips). These differences should make it possible to differentiate one effect from the other. Then it is also possible that the drag effect that has been observed occurred in the presence of some contact between the two layers or of tunneling that allowed spin current to flow and thus had a contribution from the effect discussed here. To our knowledge sensitivity to applied magnetic field in the plane was not checked in those cases. In the presence of a magnetic field $`B`$ in the $`z`$ direction an ordinary Hall voltage across the longitudinal strip will be generated, which will cause charge current to flow across the transverse strip and give another contribution to the voltage generated by the spin current Eq. (8). The total voltage across the transverse strip will be $$V_t(B)=(R_o^2B^2+R_s^2B_{eq}^2)\frac{l}{\rho }j_x,$$ (11) with $`B_{eq}=4\pi n\mu _B/\sqrt{2}`$. The sign of the ordinary contribution to the voltage across the transverse strip is the same as that of the spin Hall effect, independent of the sign of the applied magnetic field $`B`$ and of the sign of $`R_o`$. Even if experimental resolution impedes accurate measurement of $`V_t`$ for $`B=0`$, it may be possible to extract the effect discussed here from extrapolation of results for $`V_t(B)`$ to $`B=0`$. In conclusion, the experiment proposed here, if successful, would achieve the following: (1) It would provide a realization of spin current flow in the absence of charge current flow; (2) it would demonstrate that flow of a spin current results in the generation of a transverse electric field; (3) it would show the generation of spin imbalance in a paramagnetic metal when a charge current circulates; (4) it would establish the existence of a skew scattering mechanism in a paramagnetic metal; (5) it would provide information on the magnitude and temperature dependence of the anomalous Hall coefficient $`R_s`$ in a paramagnetic metal; (6) measurement of the dependence of the voltage $`V_{SC}`$ on strip width, temperature and magnetic field would provide information on processes that lead to loss of spin coherence; (7) measurement of dependence of the magnitude of $`R_s`$ on sample purity and temperature would provide information on the scattering mechanism(s) responsible for $`R_s`$, and in particular on whether a periodic potential by itself can give rise to an anomalous Hall effect; (8) assuming a known sign for $`R_s`$ of the longitudinal strip, it would allow determination of the sign of $`R_s`$ of the transverse strip. These and other findings resulting from this experiment could have practical applications in the field of spin electronics. Even though we discussed the effect here assuming a metallic sample it is possible that semiconducting samples may allow for easier detection of this effect. Furthermore, it would be of interest to study this effect in the limit where the strips in figure 3 are two-dimensional, as in the electron bilayer systems in GaAs double quantum well structures extensively used in studies of the quantum Hall effect.
no-problem/9906/astro-ph9906252.html
ar5iv
text
# Large scale flows in the solar interior: Effect of asymmetry in peak profiles ## 1 Introduction Ring diagram analysis has been extensively used to infer horizontal flows in outer part of the solar convection zone (Hill 1988; Patrón et al. 1997; Basu, Antia & Tripathy 1999). This technique is based on the study of three-dimensional (henceforth 3d) power spectra of solar p-modes on a part of the solar surface. In all the studies so far the power spectra have been fitted to symmetric Lorentzian peak profiles to calculate the frequency shifts due to velocity field. It has been demonstrated that in general, the peaks in solar oscillation power spectra are not symmetric (Duvall et al. 1993; Toutain 1993; Nigam & Kosovichev 1998; Toutain et al. 1998) and the use of symmetric profiles may cause the fitted frequency to be shifted away from the true value. The resulting frequency shift in the high degree modes using ring diagram analysis has been found to be significant for f-modes and low order p-modes (Antia & Basu 1999). Thus, it would be interesting to study how the use of asymmetric profiles affects the measurement of flow velocities. In this work we use the data obtained by the Michelson Doppler Imager (MDI) on board the Solar and Heliospheric Observatory (SOHO) to measure the flow velocities in the outer part of the solar convection zone. We fit the power spectra using asymmetric peak profiles and compare the results with our earlier results obtained using symmetric profiles (Basu et al. 1999). Although, it is possible to study the variation in horizontal velocity with both longitude and latitude, in this work, like in Basu et al. (1999), we have only considered the longitudinal averages, which contains information about the latitudinal variation in these flows. For this purpose, at each latitude we have summed the spectra obtained for different longitudes to get an average spectrum which has information on the average flow velocity at each latitude. We have studied both the rotational and meridional components of flow velocities. The rest of the paper is organized as follows: the basic technique used to calculate the horizontal flow velocities using ring diagrams is described in § 2. The results are discussed in § 3, while the conclusions from our study are summarized in § 4. ## 2 The technique We adopt the ring diagram technique (Hill 1988; Patrón et al. 1997) to obtain the 3d power spectra of solar oscillations. We have used data from full-disk Dopplergrams obtained by the MDI instrument of the Solar Oscillations Investigation (SOI) on board SOHO. We have used the same spectra that were obtained by Basu et al. (1999). These spectra are based on regions covering about $`15^{}\times 15^{}`$ with $`128\times 128`$ pixels in heliographic longitude and latitude and centered at latitudes ranging from $`60^{}`$N to $`60^{}`$S. Each region was tracked for 4096 minutes and we have taken averages over 12 spectra covering an entire Carrington rotation in longitude. These spectra are spaced by $`30^{}`$ in longitude of the central meridian, which would correspond to a time interval of about 3273 minutes. Thus there is some overlap between different spectra. Similarly, we have taken spectra centered at different latitudes separated by $`5^{}`$ and hence there would be considerable overlap in region covered by spectra centered at neighboring latitudes. The data cover the period from about May 24 to June 21, 1996. Thus summing gives us a time averaged spectrum for the period covered. To extract the flow velocities and other mode parameters from the 3d power spectra we fit a model of the form $$P(k_x,k_y,\nu )=\frac{\mathrm{exp}(A_0+(kk_0)A_1+A_2(\frac{k_x}{k})^2+A_3\frac{k_xk_y}{k^2})(S^2+(1+Sx)^2)}{x^2+1}+\frac{e^{B_1}}{k^3}+\frac{e^{B_2}}{k^4},$$ (1) where $$x=\frac{\nu ck^pU_xk_xU_yk_y}{w_0+w_1(kk_0)},$$ (2) $`k^2=k_x^2+k_y^2`$, $`k`$ being the total wave number, and the 13 parameters $`A_0,A_1,A_2,A_3,c,p,U_x,U_y,w_0,w_1,S,B_1`$ and $`B_2`$ are determined by fitting the spectra using a maximum likelihood approach (Anderson, Duvall & Jefferies 1990). Here, $`k_0`$ is the central value of $`k`$ in the fitting interval and $`\mathrm{exp}(A_0)`$ is the mean power in the ring. The coefficient $`A_1`$ accounts for the variation in power with $`k`$ in the fitting interval, while $`A_2`$ and $`A_3`$ terms account for the variation of power along the ring. The term $`ck^p`$ is the mean frequency, while $`U_xk_x`$ and $`U_yk_y`$ represent the shift in frequency due to large scale flows and the fitted values of $`U_x`$ and $`U_y`$ give the average flow velocity over the region covered by the power spectrum and the depth range where the corresponding mode is trapped. The mean half-width is given by $`w_0`$, while $`w_1`$ takes care of the variation in half-width with $`k`$ in the fitting interval. The terms involving $`B_1,B_2`$ define the background power, which is assumed to be of the same form as Patrón et al. (1997). $`S`$ is a parameter that controls the asymmetry, and the form of asymmetry is the same as that prescribed by Nigam & Kosovichev (1998). This parameter is positive for positive asymmetry, i.e., more power on the higher frequency side of the peak, and negative for negative asymmetry. By setting $`S=0`$ we can fit symmetric Lorentzian profiles. The fits are obtained by maximizing the likelihood function $`L`$ or minimizing the function $`F`$ given by $$F=\mathrm{ln}L=\underset{i}{}\left(\mathrm{ln}M_i+\frac{O_i}{M_i}\right),$$ (3) where the summation is taken over each pixel in the fitting interval. The term $`M_i`$ is the result of evaluating the model given by Eq. (1) at $`i^{\mathrm{th}}`$ pixel defined by $`k_x,k_y,\nu `$ in the 3d power spectrum, and $`O_i`$ is the observed power at the same pixel. To evaluate the quality of the fit we use the merit function (cf., Anderson et al. 1990) $$F_m=\underset{i}{}\left(\frac{O_iM_i}{M_i}\right)^2$$ (4) where the summation is over all pixels in the fitting interval. Ideally, the merit function per degree of freedom should be close to unity. The number of degrees of freedom can be defined as the difference between the number of pixels in the fitting interval and the number of free parameters in the model. We fit each ring separately by using the portion of power spectrum extending halfway to the adjoining rings. For each fit a region extending about $`\pm 100\mu `$Hz from the chosen central frequency is used. We choose the central frequency for fit in the range of 2–5 mHz. Power outside this range is not significant. The rings corresponding to $`0n6`$ have been fitted. In this work we express $`k`$ in units of $`R_{}^1`$, which enables us to identify it with the degree $`\mathrm{}`$ of the spherical harmonic of the corresponding global mode. The fitted $`U_x`$ and $`U_y`$ for each mode represents an average of the velocities in the $`x`$ and $`y`$ directions over the entire region in horizontal extent and over the vertical region where the mode is trapped. We can invert the fitted $`U_x`$ (or $`U_y`$) for a set of modes to infer the variation in horizontal flow velocity $`u_x`$ (or $`u_y`$) with depth. We use the Regularized Least Squares (RLS) as well as the Optimally Localized Averages (OLA) techniques for inversion as outlined by Basu et al. (1999). The results obtained by these two independent inversion techniques are compared to test the reliability of inversion results. For the purpose of inversion, the fitted values of $`U_x`$ and $`U_y`$ are interpolated to the nearest integral value of $`k`$ (in units of $`R_{}^1`$) and then the kernels computed from a full solar model with corresponding value of degree $`\mathrm{}`$ are used for inversion. Since the fitted modes are trapped in outer region of the Sun, inversions are carried out for $`r>0.97R_{}`$ only. ## 3 Results Following the procedure outlined in Section 2 we fit the form given by Eq. (1) to suitable regions of the 3d spectra. We use approximately 900 modes covering $`200\mathrm{}1100`$ and $`2000\nu 5000`$ $`\mu `$Hz for each spectrum. Fig. 1 shows some of the fitted quantities for the averaged spectrum obtained from the region centered at the equator. This figure can be compared with Fig. 2 in Basu et al. (1999). The asymmetry parameter $`S`$ is found to be significant and has a negative value for all the modes. The magnitude of $`S`$ appears to increase with frequency. However, the use of asymmetric profiles does not appear to affect the fitted half-width or the horizontal velocities $`U_x`$ and $`U_y`$ significantly. As with symmetric profiles, the fitted half-width $`w_0`$ appears to increase at low frequencies. This increase is probably artificial since the actual width is less than the resolution limit of the spectra. We have verified that keeping the width fixed during the fit for these modes does not affect the values of $`U_x`$ and $`U_y`$ obtained from the fits. In order to test whether the fit with additional parameter $`S`$ is indeed better we show in Fig. 2 the merit function \[cf., Eq. (4)\] per degree of freedom for fits to both symmetric and asymmetric profiles. It is clear that the merit function has reduced significantly when the additional parameter $`S`$ is fitted. The value is close to unity for all modes at low $`n`$ which can be identified by the ridges in the low frequency end of Fig. 2. The improvement in the fit is not very clear for higher order modes, as the merit function does not appear to reduce significantly. Similar results were found for fits to 2d spectra obtained by averaging over the azimuthal direction (Antia & Basu 1999). In the 2d spectra too the f-modes were found to be distinctly asymmetric (Antia & Basu 1999), while the difference between fits to symmetric and asymmetric profiles was progressively less clear for higher order modes, even though the asymmetry parameter $`S`$ had similar magnitude. This is mainly because the estimated errors increase with $`n`$ as the corresponding rings in the 3d power spectra get smaller. It may be noted that for symmetric profiles we have done a number of experiments by including more terms denoting variations in parameters not included in Eq. (1), but these do not improve the fits significantly and merit function does not reduce perceptibly (Basu et al. 1999). Thus it appears that some asymmetry is indeed required to get good fits. This asymmetry can be clearly seen in azimuthally averaged spectra shown by Antia & Basu (1999). The asymmetry parameter $`S`$ shown in Fig. 1 is similar to what was found for the 2d fits in Antia & Basu (1999). Hence, it is clear that although it is difficult to visualize asymmetry in 3d fits, the peak profiles are indeed asymmetric. There does not appear to be significant difference between the $`U_x`$ or $`U_y`$ obtained by fitting symmetric or asymmetric profiles to spectra obtained for the equatorial regions. However, at high latitudes the situation becomes somewhat different as can be seen from Figs. 3 and 4 which compare the fits to spectra from regions centered at a latitude of $`50^{}`$N. It may be noted that even for these latitudes the fitted values for the asymmetry parameter $`S`$ are similar to what is found for equatorial spectra. However, the fitted values of $`U_x`$ and $`U_y`$ are found to be more sensitive to asymmetry at higher latitudes, probably due to effects of foreshortening. The fits with the asymmetric profile give better results in general, as can be seen from the fact that the ridges in $`U_x,U_y`$ for different values of $`n`$ tend to merge better. The difference is particularly significant in $`U_y`$. The steep trend in the different ridges corresponding to low $`n`$ in $`U_y`$ obtained by fitting symmetric profiles results in large number of outliers during the inversions. As a result, we have to weed out a large number of modes to get reasonable inversion results. We weed out all modes which have residuals larger than $`4\sigma `$ for either $`U_x`$ or $`U_y`$ in an RLS inversion with low smoothing. This does not happen when asymmetric profiles are used. However, the fits are still not perfect at these latitudes even with use of asymmetric profiles and the merit function per degree of freedom is also somewhat larger than unity for low order modes. Clearly, more terms or a different form is required to improve the fits at high latitudes. To study the effect of asymmetry on inverted profiles for horizontal velocity components $`u_x`$ and $`u_y`$, we perform inversions using both the fits and the results are shown in Figs. 5 and 6. It can be seen that except in deeper layers ($`r0.98R_{}`$), the inverted results are similar regardless of whether we fit a symmetric or asymmetric profile to the spectra. From the inversion results it is not possible to decide which fit is better, but looking at the residuals and the fitted $`U_x,U_y`$ in Figs. 3,4 it appears that asymmetric profiles would give more reliable results. The rotation velocity at each latitude can be decomposed into the symmetric part \[$`(u_N+u_S)/2`$\] and an antisymmetric part \[$`(u_Nu_S)/2`$\]. The symmetric part can be compared with the rotation velocity as inferred from the splittings of global modes (Schou et al. 1998) which sample just the symmetric part of the flow. Since this component is not significantly affected by the asymmetry in peak profile we do not show the detailed results. However, the north-south antisymmetric component of rotation velocity is small and its significance is not well established from earlier studies and hence we reexamine these results with fits to asymmetric peak profiles. The results for near surface layers are shown in Fig. 7, which can be compared with Fig. 7 of Basu et al. (1999). It can be seen that the results are not significantly different from our earlier results. In deeper layers the uncertainties are larger and it is difficult to say anything about this component. Since there is a distinct improvement in fits to $`U_y`$ at high latitudes when asymmetric peak profiles are used, we try to find its effect on meridional flow inferred from ring diagram analysis. Following Hathaway et al. (1996) we try to write the meridional component as $$u_y(r,\varphi )=\underset{i}{}a_i(r)P_i^1(\mathrm{cos}(\varphi ))$$ (5) where $`\varphi `$ is the colatitude, and $`P_i^1(x)`$ are associated Legendre polynomials. The first six terms in this expansion are found to be significant and their amplitudes are shown in Fig. 8, which can be compared with Fig. 12 of Basu et al. (1999). It can be seen that use of asymmetric profiles introduces small changes in amplitudes mainly in deeper layers. The amplitudes of all even components increases slightly in deeper layers when asymmetric profiles are used for fitting. In particular, the $`P_4^1(\varphi )`$ component suggested by Durney (1993) is found to have an amplitude of about 3–4 m/s for $`0.97R_{}<r<0.99R_{}`$. There is no indication of any sign change in the meridional flow up to the depth of $`0.03R_{}`$, which is consistent with results of Braun & Fan (1998). However, the magnitude of meridional flow velocity that we find (30 m/s) is much larger than that obtained by Braun & Fan, who find a mean meridional velocity of 10–15 m/s averaged over latitudes $`20^{}`$$`60^{}`$. Clearly, more work is required to understand these differences. We can also find the latitude at which $`u_y=0`$, which is the latitude near the equator where the meridional flow diverges towards the two poles. The results shown in Fig. 9 can be compared with those obtained by González Hernández et al. (1999). In general we find this point to be much closer to the equator as compared to what they have found. The difference may be mostly due to the fact that we have averaged over all longitudes in one Carrington rotation and as a result our error estimates are much lower. For all depths we find that this point is within $`4^{}`$ of the equator and we do not find any chaotic behavior near the surface. Of course, the resolution of inversions is limited to a depth below $`r=0.999R_{}`$ since we have only used modes with $`\mathrm{}<1100`$, which have lower turning points below this depth. Thus the thin layer near the surface is not resolved by inversions and we cannot expect reliable results there. It is not clear if the small departure of this point from the equator is significant as the difference in latitude estimated from the two inversion techniques is comparable to the value. A part of the effect may also be due to systematic errors arising from misalignment in the MDI instrument (Giles et al. 1997) and hence, no particular significance may be ascribed to the location of this point. ## 4 Conclusions Using the ring diagram technique applied to MDI data we have determined horizontal velocities in the outer part of the solar convection zone ($`r>0.97R_{}`$). We find that the use of asymmetric peak profiles improve the fits to the 3d spectra —- as measured by the merit function — significantly. The asymmetry parameter is found to be negative for all the modes, i.e., there is more power on the lower frequency side of the peak. This is the same as what has been found earlier at the low-degree end of the power spectrum (Duvall et al. 1993; Toutain et al. 1998). However, the use of asymmetric profile does not affect the fitted velocities substantially. The inferred meridional flow is dominated by the $`\mathrm{sin}(2\theta )`$ component which has an amplitude of about 30 m/s in most of region covered in this study. The $`P_4^1(\varphi )`$ component suggested by Durney (1993) also has significant amplitude of about 3–4 m/s in deeper layers. The north-south symmetric component of the meridional flow is generally small and the meridional velocity (as a function of latitude) changes sign at a point very close to equator at all depths. This point marks the region where flow diverges towards the two poles on two sides. The small departure of this point from equator may not be significant. There is no change in sign of meridional velocity with depth up to about 21 Mm. This work utilizes data from the Solar Oscillations Investigation / Michelson Doppler Imager (SOI/MDI) on the Solar and Heliospheric Observatory (SOHO). SOHO is a project of international cooperation between ESA and NASA. The authors would like to thank the SOI Science Support Center and the SOI Ring Diagrams Team for assistance in data processing. The data-processing modules used were developed by Luiz A. Discher de Sa and Rick Bogart, with contributions from Irene González Hernández and Peter Giles.
no-problem/9906/cs9906003.html
ar5iv
text
# THE SYNTACTIC PROCESSING OF PARTICLES IN JAPANESE SPOKEN LANGUAGE ## Abstract Particles fullfill several distinct central roles in the Japanese language. They can mark arguments as well as adjuncts, can be functional or have semantic funtions. There is, however, no straightforward matching from particles to functions, as, e.g., ga can mark the subject, the object or an adjunct of a sentence. Particles can cooccur. Verbal arguments that could be identified by particles can be eliminated in the Japanese sentence. And finally, in spoken language particles are often omitted. A proper treatment of particles is thus necessary to make an analysis of Japanese sentences possible. Our treatment is based on an empirical investigation of 800 dialogues. We set up a type hierarchy of particles motivated by their subcategorizational and modificational behaviour. This type hierarchy is part of the Japanese syntax in VERBMOBIL. ## 1 Introduction The treatment of particles is essential for the processing of the Japanese language for two reasons. The first reason is that these are the words that occur most frequently. The second reason is that particles have various central functions in the Japanese syntax: case particles mark subcategorized verbal arguments, postpositions mark adjuncts and have semantic attributes, topic particles mark topicalized phrases and no marks an attributive nominal adjunct. Their treatment is difficult for three reasons: 1) despite their central position in Japanese syntax the omission of particles occurs quite often in spoken language. 2) One particle can fulfill more than one function. 3) Particles can cooccur, but not in an arbitrary way. In order to set up a grammar that accounts for a larger amount of spoken language, a comprehensive investigation of Japanese particles is thus necessary. Such a comprehensive investigation of Japanese particles was missing up to now. Two kinds of solutions have previously been proposed: (1) the particles are divided into case particles and postpositions. The latter build the heads of their phrases, while the former do not (cf. , ). (2) All kinds of particles build the head of their phrases and have the same lexical structure (cf. ). Both kinds of analyses lead to problems: if postpositions are heads, while case particles are nonheads, a sufficient treatment of those cases where two or three particles occur sequentially is not possible, as we will show. If on the other hand there is no distinction of particles, it is not possible to encode their different behaviour in subcategorization and modification. We carried out an empirical investigation of cooccurrences of particles in Japanese spoken language. As a result, we could set up restrictions for 25 particles. We show that the problem is essentially based at the lexical level. Instead of assuming different phrase structure rules we state a type hierarchy of Japanese particles. This makes a uniform treatment of phrase structure as well as a differentiation of subcategorization patterns possible. We therefore adopt the ‘all-head’ analysis, but extend it by a type hierarchy in order to be able to differentiate between the particles. Our analysis is based on 800 Japanese dialogues of the VERBMOBIL data concerning appointment scheduling. ## 2 The Type Hierarchy of Japanese Particles Japanese noun phrases can be modified by more than one particle at a time. There are many examples in our data where two or three particles occur sequentially. On the one hand, this phenomenon must be accounted for in order to attain a correct processing of the data. On the other hand, the discrimination of particles is motivated by their modificational and subcategorizational behaviour. We carried out an empirical analysis, based on our dialogue data. Table 1 shows the frequency of cooccurrence of two particles in the dialogue data. There is a tendency to avoid the cooccurrence of particles with the same phonology, even if it is possible in principal in some cases. The reason is obvious: such sentences are difficult to understand. treats wa, ga, wo, ni, de, to, made, kara and ya as ‘particles’. They are divided into those that are in the deep structure and those that are introduced through transformations. An example for the former is kara, examples for the latter are ga(SBJ), wo(OBJ), ga(OBJ) and ni(OBJ2). assigns all particles the part-of-speech P. Examples are ga, wo, ni, no, de, e, kara and made. All particles are heads of their phrases. Verbal arguments get a grammatical relation \[GR OBJ/SBJ\]. In the part-of-speech class P contains only ga, wo and ni. defines postpositions and case particles such that postpositions are the Japanese counterpart of prepositions in English and cannot stand independently, while case particles assign case and can follow postpositions. Her case particles include ga, wo, ni, no and wa. divides case markers (ga, wo, ni and wa) from copula forms (ni, de, na and no). He argues that ni, de, na and no are the infinitive, gerund and adnominal forms of the copula. In the class of particles, we include case particles, complementizers, modifying particles and conjunctional particles. We thus assume a common class of the several kinds of particles introduced by the other authors. But they are further divided into subclasses, as can be seen in figure 1. We assume not only a differentiation between case particles and postpositions, but a finer graded distinction that includes different kinds of particles not mentioned by the other authors. de is assumed to be a particle and not a copula, as proposes. It belongs to the class of adverbial particles. One major motivation for the type hierarchy is the observation we made of the cooccurrence of particles. Case particles (ga, wo, ni) are those that attach to verbal arguments. A complementizer marks complement sentences. Modifying particles attach to adjuncts. They are further divided into noun-modifying particles and verb-modifying particles. Verb modifying particles can be topic particles, adverbial particles, or postpositions. Some particles can have more than one function, as for example ni has the function of a case particle and an adverbial particle. Figure 1 shows the type hierarchy of Japanese particles. The next sections examine the individual types of particles. ### 2.1 Case Particles There is no number nor gender agreement between noun phrase and verb. The verbs assign case to the noun phrases. This is marked by the case particles. Therefore these have a syntactic function, but not a semantic one. Unlike in English, the grammatical functions cannot be assigned through positions in the sentence or c-command-relations, since Japanese exhibits no fixed word position for verbal arguments. The assignment of the grammatical function is not expressed by the case particle alone but only in connection with the verbal valency. There are verbs that require ga-marked objects, while in most cases the ga-marked argument is the subject: * | nantoka | | --- | | somehow | | yotei | | --- | | time | | ga | | --- | | GA | | toreru | | --- | | can take | | N desu | | --- | | COP | | ga | | --- | | SAP | (Somehow (I) can find some time.) Japanese is described as a head-final language. therefore assumes only one phrase structure rule: Mother $``$ Daughter Head. However, research literature questions whether this also applies to nominal phrases and their case particles. :45 assume Japanese case particles to be markers. On the one hand, there are several reasons to distinguish case particles and modifying particles. On the other hand, I doubt whether it is reasonable to assume different phrase structures for NP+case particle and NP+modifying particle. The phrase-structural distinction of case particles and postpositions leads to problems, when more than one particle occur. The following example comes from the Verbmobil corpus: * | naNji | | --- | | what time | | kara | | --- | | from | | ga | | --- | | GA | | yoroshii | | --- | | good | | desu | | --- | | COP | | ka | | --- | | QUE | (At what time would you like to start?) If one now assumes that the modifying particle kara is head of naNji as well as of the case particle ga, the result for naNji kara ga with the head-marker structure described in <sup>1</sup><sup>1</sup>1The Marking Principle says: In a headed phrase, the MARKING value is token-identical with that of the MARKER-DAUGHTER if any, and with that of the HEAD-DAUGHTER otherwise. would be as shown in figure 2. The case particle ga would have to allow nouns and modifying particles in SPEC. The latter are however normally adjuncts that modify verbal projections. Therefore the head of kara entails the information that it can modify a verb. This information is inherited to the head of the whole phrase by the Head-Feature Principle as is to be seen in the tree above. As a result, this is also admitted as an adjunct to a verb, which leads to wrong analyses for sentences like the following one: * | \*naNji | | --- | | what time | | kara | | --- | | from | | ga | | --- | | GA | | sochira | | --- | | you | | ga | | --- | | GA | | jikaN | | --- | | time | | ga | | --- | | GA | | toremasu | | --- | | can take | | ka | | --- | | QUE | If, on the other hand, case particles and topic markers are heads, one receives a consistent and correct processing of this kind of example too. This is because the head information \[MOD none\] is given from the particle ga to the head of the phrase naNji kara ga. Thus this phrase is not admitted as an adjunct. Instead of assuming different phrase structure rules, a distinction of the kinds of particles can be based on lexical types. HPSG offers the possibility to define a common type and to set up specifications for the different types of particles. We assume Japanese to be head-final in this respect. All kinds of particles are analysed as heads of their phrases. The relation between case particle and nominal phrase is a ‘Complement-Head’ relation. The complement is obligatory and adjacent<sup>2</sup><sup>2</sup>2Obligatory Japanese arguments are always adjacent, and vice versa.. Normally the case particle ga marks the subject, the case particle wo the direct object and the case particle ni the indirect object. There are, however, many exceptions. We therefore use predicate-argument-structures instead of a direct assignment of grammatical functions by the particles (and possibly transformations). The valency information of the Japanese verbs does not only contain the syntactic category and the semantic restrictions of the subcategorized arguments, but also the case particles they must be annotated with<sup>3</sup><sup>3</sup>3 investigates the particles ni, ga and wo and also states that grammatical functions must be clearly distinguished from surface cases. In most cases the ga-marked noun phrase is the subject of the sentence. However, this is not always the case. Notably stative verbs subcategorize for ga-marked objects. An example is the stative verb dekimasu<sup>4</sup><sup>4</sup>4see for a semantic classification of verbs that take ga-objects: * | kanojo | | --- | | she | | ga | | --- | | GA | | oyogi | | --- | | swimming | | ga | | --- | | GA | | dekimasu | | --- | | can | (She can swim.) These and other cases are sometimes called ‘double-subject constructions’ in the literature. But these ga-marked noun phrases do not behave like subjects. They are neither subject to restrictions on subject honorification nor subject to reflexive binding by the subject. This can be shown by the following example: * | gogo | | --- | | afternoon | | no | | --- | | NO | | hou | | --- | | side | | ga | | --- | | GA | | yukkuri | | --- | | at ease | | hanashi | | --- | | talking | | ga | | --- | | GA | | dekimasu | | --- | | can | | ne | | --- | | SAP | (We can talk at ease in the afternoon.) hanashi does not meet the semantic restriction \[+animate\] stated by the verb dekimasu for its subject. There are even ga-marked adjuncts. assumes these ‘double-subject constructions’ to be derived from genitive relations. But this analysis seems not to be true for example (5)), because the following sentence is wrong: * | \*gogo | | --- | | afternoon | | no | | --- | | NO | | hou | | --- | | side | | no | | --- | | NO | | yukkuri | | --- | | at ease | | hanashi | | --- | | talk | | ga | | --- | | GA | | dekimasu | | --- | | can | | ne | | --- | | SAP | The case particle wo normally marks the direct object of the sentence. In contrast to ga, no two phrases in one clause may be marked by wo. This restriction is called ‘double-wo constraint’ in research literature (see, for example, :249ff.). Object positions with wo-marking as well as subject positions with ga-marking can be saturated only once. There are neither double subjects nor double objects. This restriction is also valid for indirect objects. Arguments found must be assigned a saturated status in the subcategorization frame, so that they cannot be saturated again (as in English). The verbs subcategorize for at most one subject, object and indirect object. Only one of these arguments may be marked by wo, while a subject and an object may both be marked by ga. These attributes are determined by the verbal valency. The wo-marked argument is not required to be adjacent to the verb. It is possible to reverse NP-ga and NP-wo as well as to insert adjuncts between the arguments and the verb. The particle ni can have the function of a case particle as well as that of an adjunct particle modifying the predicate. also identify homophoneous ni that can mark adjuncts or complements. They use the notion of ‘affectedness’ to distinguish them. This is however not useful in our domain. suggest testing the possibility of passivization. Some verbs subcategorize for a ni-marked object, as for example naru: * | raigetsu | | --- | | next month | | ni | | --- | | NI | | naru | | --- | | become | | N desu | | --- | | COP | | ga | | --- | | SAP | (It will be next month.) ni-marked objects cannot occur twice in the same clause, just as ga-marked subjects and wo-marked objects. The ‘double-wo constraint’ is neither a specific Japanese restriction nor a specific peculiarity of the Japanese direct object. It is based on the wrong assumption that grammatical functions are assigned by case particles. There are a lot of examples with double NP-ni, but these are adjuncts. The lexical entries of case particles get a case entry in the HEAD. Possible values are ga, wo, ni and to. They are neither adjuncts nor specifiers and thus get the entries \[MOD none\] and \[SPEC none\]. They subcategorize for an adjacent object. This can be a noun, a postposition or an adverbial particle<sup>5</sup><sup>5</sup>5A fundamental difference between Japanese grammar and English grammar is the fact that verbal arguments can be optional. For example, subjects and objects that refer to the speaker are omitted in most cases in spoken language. The verbal arguments can freely scramble. Additionally, there exist adjacent verbal arguments. To account for this, our subcategorization contains the attributes SAT and VAL. In SAT it is noted, whether a verbal argument is already saturated (such that it cannot be saturated again), optional or adjacent. VAL contains the agreement information for the verbal argument. Adjacency must be checked in every rule that combines heads and arguments or adjuncts.. ### 2.2 The Complementizer to to marks adjacent complement sentences that are subcategorized for by verbs like omou, iu or kaku. * | sochira | | --- | | you | | ni | | --- | | NI | | ukagaitai | | --- | | visit | | to | | --- | | TO | | omoimasu | | --- | | think | | node | | --- | | SAP | (I would like to visit you.) Some verbs subcategorize for a to marked object. This object can be optional or obligatory with verbs like kuraberu. * | kono | | --- | | that | | hi | | --- | | day | | mo | | --- | | too | | chotto | | --- | | somewhat | | hito | | --- | | people | | to | | --- | | TO | | au | | --- | | meet | | yotei | | --- | | plan | | ga | | --- | | GA | | gozaimasu | | --- | | exist | (That day too, there is a plan to meet some people.) to in these cases is categorized as a complementizer. Another possibility is that to marks an adjunct to a predicate, which qualifies to as a verb modifying particle: * | shimizu | | --- | | Shimizu | | seNsei | | --- | | Prof. | | to | | --- | | TO | | teNjikai | | --- | | exhibition | | wo | | --- | | WO | | go-issho | | --- | | together | | sasete | | --- | | do | | itadaku | | --- | | HON | (I would like to organize an exhibition with Prof. Shimizu.) Finally, the complementizer to can be an NP conjunction (which will not be considered at the moment, see ). The complementizer gets a case entry, because its head is a subtype of case-particle-head. It subcategorizes for a noun, a verb, an utterance, an adverbial particle or a postposition. ### 2.3 Modifying Particles An essential problem is to find criteria for the distinction of case particles and modifying particles. On the semantic level they can be distinguished in that modifying particles introduce semantics, while case particles have a functional meaning. According to this, the particle no is a modifying one, because it introduces attributive meaning, as opposed to (:134), who classifies it as a case particle. Another distinctive criterion that is introduced by :135 says that modifying particles<sup>6</sup><sup>6</sup>6He calls them ‘postpositions’. are obligatory in spoken language, while case particles can be omitted. Case particles are indeed suppressed more often, but there are also cases of suppressed modifying particles. These occur mainly in temporal expressions in our dialogue data: * | soredewa | | --- | | then | | juuyokka | | --- | | 14th | | no | | --- | | NO | | gogo | | --- | | afternoon | | Ø | | --- | | Ø | | niji | | --- | | 2 o’clock | | Ø | | --- | | Ø | | robii | | --- | | lobby | | no | | --- | | NO | | hou | | --- | | side | | de | | --- | | DE | | o machi | | --- | | HON-wait | | shite | | --- | | do | | orimasu | | --- | | AUX-HON | (I will then wait in the lobby at 2 o’clock on the 14th.) Finally gives the criterion that case particles can follow modifying particles while modifying particles cannot follow case particles. This criterion in particular implies that a finer distinction is necessary, as we have shown that it is not that easy. This can be realized with HPSG types. According to this criterion, no behaves like a modifying particle, while according to the criterion on meaning, it behaves like a case particle. Our first distinction is thus a functional one: modifying particles differ from case particles in that their marked entities are not subcategorized for by the verb. Case particles get the head information \[CASE case\] that controls agreement between verbs and their arguments. Modifying particles do not get this entry. They get the information in MOD that they can become adjuncts to verbs (verb modifying particles) or nouns (the noun modifying particle no) and semantic information. They subcategorize for a noun, as all particles do. The modifying particles share the following features in their lexical entries. #### 2.3.1 Verb Modifying Particles The verb modifying particles specify the modification of the verb in MOD. The postpositions modify a (nonauxiliary) verb as an adjunct and subcategorize for a nominal object. treats ni and de as the infinitive and the gerund form of the copula. ni is similar to the infinitive form to the extend that it can take an adverb as its argument (gogo wa furii ni nat-te i-masu – afternoon - WA - free - become). But the infinitive is clearly distinct from the characteristics of ni, that cannot be used with N desu, cannot mark a relative sentence (\*John ga furii ni koto) and cannot be marked with the complementizer to (\*John ga furii ni to omou). The adjunctive form ‘de’ has both qualities of a gerundive copula and qualities of a particle. But there is some data that shows different behaviour of de and other gerundives. Firstly, it concerns the cooccurrence possibilities of de and other particles, compared to gerundive forms and particles: | de wa - V-te wa | de mo - V-te mo | de no - V-te no | de ni - \*V-te ni | | --- | --- | --- | --- | | de ga - \*V-te ga | de wo - \*V-te wo | de de - \*V-te de | | Secondly, a gerund may modify auxiliaries, e.g. shite kudasai, shite orimasu, but de may not. Additionally there is something which distinguishes de of a copula: it may not subcategorize for a subject. A word that is an adjunct to verbs, subcategorizes for an unmarked noun or a postpositional phrase and is subcategorized for by several particles (see above) fits well into our description of a verb modifying particle. The adverbial particles ni, de and to subcategorize for a noun or a postposition. As already described, to behaves like an adverbial particle, too. #### 2.3.2 The Noun Modifying Particle NO no is a particle that modifies nominal phrases. This is an attributive modification and has a wide range of meanings.<sup>7</sup><sup>7</sup>7See also :134ff. assigns no to the class of case particles. However, the criteria she sets up to distinguish between case particles and postpositions do not apply to this classification of no: firstly, Tsujimura’s postpositions have their own semantic meaning. Case particles have a functional meaning. no however has a semantic, namely attributive meaning. Secondly, Tsujimura’s postpositions are obligatory in spoken language, case particles are optional. no is as obligatory as kara and made. Finally, Case particles can - as Tsujimura states - follow postpositions, but postpositions cannot follow case particles. According to this criterion, no behaves like a case particle. no combines qualities of case particles with those of modifying particles (which Tsujimura calls ‘postpositions’). This means that a special treatment of this particle is necessary. The particle no subcategorizes for a noun, as the other particles do. It also modifies a noun. This separates it from the other modifying particles. The particle no modifies a noun phrase and occurs after a noun or a verb modifying particle. #### 2.3.3 Particles of Topicalization The topic particle wa can mark arguments as well as adjuncts. In the case of argument marking it replaces the case particle. In the case of adjunct marking it can replace the verb modifying particle or it can occur after it. On the syntactic level, it has to be decided, whether the topic particle marks an argument or an adjunct, when it occurs without a verb modifying particle. This is difficult because of the optionality of verbal arguments in Japanese. If it marks an argument, it has to be decided which grammatical function this argument has. This problem can often not be solved on the purely syntactic level. Semantic restrictions for verbal arguments are necessary: * | basho | | --- | | place | | no | | --- | | NO | | hou | | --- | | side | | wa | | --- | | WA | | dou | | --- | | how | | shimashou | | --- | | shall do | | ka | | --- | | QUE | (How shall we resolve the problem of the place?) Subject and object of the verb shimashou are suppressed in this example. The sentence can be interpreted as having a topic adjunct, but no surface subject and object, when using semantic restrictions for the subject (agentive) and the object (situation). analyses Japanese topicalization with a trace that introduces a value in SLASH and the ‘Binding Feature Principle’ that unifies the value of SLASH with a wa-marked element<sup>8</sup><sup>8</sup>8The Binding Feature Principle says: The value of a binding feature of the mother is identical to the union of the values of the binding feature of the daughters minus the category bound in the branching. . This treatment is similar to the one introduced by for the treatment of English topicalization. However, Japanese topicalization is fundamentally different from English one. Firstly, it occurs more frequently. Up to 50% of the sentences are concerned (). Secondly, there are examples where the topic occurs in the middle of the sentence, unlike the English topics that occur sentence-initially. Thirdly, suppressing of verbal arguments in Japanese could be called more a rule than an exception in spoken language. The SLASH approach would introduce traces in almost every sentence. This, in connection with scrambling and suppressed particles, could not be restricted in a reasonable way. If one follows Gunji’s interpretation of those cases, where the topic-NP can be interpreted as a noun modifying phrase, a genitive gap has to be assumed. But this leads to assuming a genitive gap for every NP that is not modified. Further, genitive modification can be iterated. Finally, two or three occurences of NP-wa are possible in one utterance. Thus, we decided to assign topicalized sentences the same syntactic structure as non-topicalized sentences and to resolve the problem on the lexical level. The topic particle is, on the syntactic level, interpreted as a verbal adjunct. The binding to verbal arguments is left to the semantic interpretation module in VERBMOBIL, see figure 5. mo is similar to wa in some aspects. It can mark a predicative adjunct and can follow de and ni. But it can also follow wa, an adjective and a sentence with question mark: * | dekiru | | --- | | can | | ka | | --- | | QUE | | mo | | --- | | MO | | shiremaseN | | --- | | do not know | (I don’t know if I can) mo is a particle that has the head of a topic-adverbial particle, but a different subcategorization frame than wa. koso is another topic particle that can occur after nouns, postpositions or adverbial particles. ### 2.4 Omitted Particles Some particles can be omitted in Japanese spoken language. Here is an example from the Verbmobil corpus: * | rokugatsu | | --- | | June | | Ø | | --- | | Ø | | juusaNnichi | | --- | | 13th | | no | | --- | | NO | | kayoubi | | --- | | Tuesday | | Ø | | --- | | Ø | | gogo | | --- | | afternoon | | kara | | --- | | KARA | | wa | | --- | | WA | | ikaga | | --- | | good | | deshou | | --- | | COP | | ka | | --- | | QUE | (Would the 13th of June suit you?) This phenomenon can be found frequently in connection with pronouns and temporal expressions in the domain of appointment scheduling. assumes that exclusively wa can be suppressed. however shows that there are contexts, where ga, wo or even e can be omitted. He assigns it as ‘phonological deletion’. analyses omitted wo particles and explains these with linearization: a particle wo can only be omitted, when it occurs directly before a verb. however gives examples to prove the opposite. It can be observed that NPs without particles can fulfill the functions of a verbal argument or of a verbal adjunct (ex. (14)). We decided to interpret these NPs as verbal adjuncts and to leave the binding to argument positions to the semantic interpretation. NPs thus get a MOD value that allows them to modify nonauxiliary verbs. ### 2.5 ga-Adjuncts One can find several examples with ga marked adjuncts in the Verbmobil data. On the level of information structure it is said that ga marks neutral descriptions or exhaustive descriptions (c.f. , ). Gunji analyses these exhaustive descriptions syntactically in the same way as he analyses his ‘type-I topicalization’. They build adjuncts that control gaps or reflexives in the sentence. He views ga marked adjuncts without control relations as relying on a very specialized context. However, his treatment leads to problems. Firstly, in all cases, where ga marks a constituent that is subcategorized as ga-marked by the verb, a second reading is analysed that contains a ga marked adjunct controlling a gap. This is not reasonable. The treatment of the different meaning of ga marking and wa marking belongs to the semantics and not into the phrase structure. Secondly, this treatment assumes gaps. We already criticized this in connection with topicalization. Therefore, we do not need reflexive control at the moment. However, it contains mostly examples with ga marked adjuncts without syntactic control relation to the rest of the sentences. At the level of syntax, we do not decide whether a ga-marked subject or object is a neutral description or an exhaustive listing. This decision must be based on context information, where it can be ascertained whether the noun phrase is generic, anaphoric or new. We distinguish occurrences of NP+ga that are verbal arguments from those that are adjuncts. The examples for ga-marked adjuncts in the Verbmobil dialogues either describe a temporal entity or a human. All cases found are predicate modifying. To further restrict exhaustive interpretations, we introduced selectional restrictions for the marked NP, based on observations in the data. ## 3 Conclusion The syntactic behaviour of Japanese particles has been analysed based on the Verbmobil dialogue data. We observed 25 different particles in 800 dialogues on appointment scheduling. It has been possible to set up a type hierarchy of Japanese particles. We have therefore adopted a lexical treatment instead of a syntactic treatment based on phrase structure. This is based on the different kinds of modification and subcategorization that occur with the particles. We analysed the Japanese particles according to their cooccurrence potential, their modificational behaviour and their occurrence in verbal arguments. We clarified the question which common characteristics and differences between the individual particles exist. A classification in categories was carried out. After that a model hierarchy could be set up for an HPSG grammar. The simple distinction into case particles and postpositions was proved to be insufficient. The assignment of the grammatical function is done by the verbal valency and not directly by the case particles. The topic particle is ambiguous. Its binding is done by ambiguity and underspecification in the lexicon and not by the Head-Filler Rule as in the HPSG for English (). The approach presented here is part of the syntactic analysis of Japanese in the Verbmobil machine translation system. It is implemented in the PAGE parsing system . It has been proved to be essential for the processing of a large amount of Japanese dialogue data. Further research concerning coordinating particles (to, ya, toka, yara, ka etc.) and sentence end particles (ka, node, yo, ne etc.) is necessary.
no-problem/9906/cond-mat9906113.html
ar5iv
text
# Symmetric Diblock Copolymers in Thin Films (II): Comparison of Profiles between Self-Consistent Field Calculations and Monte Carlo Simulations. ## I Introduction. Confinement of spatially structured fluids into thin films gives rise to a rich and interesting interplay between the intrinsic length scale of their structure in the bulk and the geometry of the film. One the one hand, confinement generally modifies the phase diagram of fluids. For spatially structured phases the confinement leads to transitions between phases with identical symmetry but different orientation with respect to the confining walls. On the other hand the surfaces alter the local structure of the fluid in their vicinity. This gives rise to the enrichment of one component at the surface and the formation of wetting layers, but it also entails changes in the local fluid structure (e.g., packing and alignment effects). Controlling the properties of confined fluids is important for various practical applications. The orientation of the morphology is, for instance, important for connecting the self-assembled material to other devices (e.g., to form a molecular sieve). Studying effects of confinement on spatially structured polymeric fluids is particularly rewarding, because the large length scale of the phenomena, set by the chain extension, leads to a rather universal behavior, i.e., many features are independent from the details of the chemical architecture. The equilibrium properties of polymeric systems are usually well describable by self-consistent field calculations. Moreover, the large length scales facilitate the application of experimental techniques. The detailed composition profiles in the vicinity of surfaces and even profiles of individual segments are experimentally accessible. However, the large lengths scale of the phenomena goes along with protracted long time scales to reach equilibrium. There are quantitative comparisons between experiments and theory: The profiles of individual segments in films of diblock copolymers and their blends have been measured by small angle neutron scattering, and the results have been compared to self-consistent field calculations. The studies found that the self-consistent field calculations describe the qualitative features, but that the profiles were broadened by long wavelength fluctuations of the local position of the interfaces. In the previous paper, hereafter denoted as paper I, we have used a self-consistent field technique similar to Matsen for calculating the phase diagram of a symmetric $`AB`$ diblock copolymer melt confined into a thin film as a function of the film thickness and temperature. The calculations have been compared to Monte Carlo simulations in the framework of the bond fluctuation model at intermediate segregation ($`\chi N=30`$). Depending on the film thickness the systems assembled into parallel oriented lamellar phases $`L_2`$ and $`L_4`$, in which $`2`$ or $`4`$ $`AB`$ interfaces run parallel to the surfaces of the film, or a perpendicular lamellar phase $`L_{}`$ in which the interfaces run perpendicular to the confining surfaces. The dependence on the film thickness exhibited by the Monte Carlo simulation is in agreement with the self-consistent field calculations. The aim of the present paper is to investigate the local structure of these films in more detail. The detailed profiles of the composition and individual segments provide a sensitive testing bed for comparing Monte Carlo simulations and self-consistent field theory. While the comparisons between experiments and theory have provided many valuable insights, Monte Carlo simulations might contribute to our understanding by investigating model systems in which many different quantities are simultaneously accessible. In the simulations we can analyze profiles of different quantities on various lateral length scales, and we can compare the $`AB`$ diblock copolymer melt with a blend of $`A`$ and $`B`$ homopolymers of identical architecture. There are several potential sources of deviations between the Monte Carlo simulations/experiments and the self-consistent field calculations: (i) At low incompatibility, i.e., around the onset of spatial ordering, fluctuations become important. These shift the transition temperature from its mean field value $`\chi N=10.5`$ to higher values of the incompatibility, and the transition becomes first order. Shifts of the transition temperature of similar magnitude are predicted in the P-RISM framework by Schweizer and co–workers. In our Monte Carlo simulations we observe the order-disorder transition in the regime $`13.5\chi N17`$ (cf. Fig.7 in paper I). At higher segregation, however, fluctuation effects become less important. (ii) The self-consistent field theory assumes the interface between the $`A`$ and $`B`$ domains to be ideally flat in the lamellar phase. However, there are long wavelength fluctuations of the local interfacial position (i.e., capillary waves). Their free energy cost vanishes as their wavelength diverges, and, hence, they are important at all temperatures. These fluctuations lead to a broadening of the laterally averaged profiles measured in experiments or Monte Carlo simulations. (iii) Moreover, computer simulations and experiments reveal that the copolymers assume a dumbbell–like shape already above the order-disorder transition temperature, as to avoid energetically unfavorable contacts between their different moieties. This leads to an increase of the chain extension. The interplay between the inter- and intramolecular energy has been studied in the P-RISM framework by Schweizer and co-workers and field theoretical approaches (iv) At high incompatibility ($`\chi 𝒪(1)`$), the smallest length scale of the spatial structure (e.g., the width of the interface between the $`A`$ and $`B`$ domains) decreases rapidly. If this length scale becomes comparable to the length scale of the underlying microscopic structure (e.g., persistence length of the polymer or the range of packing effects of the monomers) a description in the framework of the Gaussian chain model breaks down. In fact, even for simple systems like interfaces between polymers of different persistence length the Gaussian chain model gives qualitatively erroneous predictions at high incompatibility. Density functional or P-RISM calculations of more realistic models might account for the changes in the local fluid structure. Comparing self-consistent field calculations in the framework of the Gaussian chain model and density functional theory for a model with local fluid structure, Nath and co–workers observed rather pronounced deviations between both approaches. (v) The confining surfaces create a sharp density gradient. In this region the local structure of the polymeric fluid is important at all incompatibilities. The monomer density exhibits pronounced oscillations in the vicinity of the walls (cf. Fig.8 in paper I). Moreover, polymers align parallel to the wall. The latter effect is partially captured in the self-consistent field calculations and favors a perpendicular arrangement of the lamellae. The aim of our paper to present a quantitative comparison between the Monte Carlo simulations and self-consistent field calculations and to quantify the effects discussed above in the framework of a well-studied model. The outline of our paper is as follows: In the next section we give a brief synopsis of the model and the computational technique. The reader is refered to paper I for further technical details. Then we discuss the strength of interfacial fluctuations and compare the laterally averaged profiles of the Monte Carlo simulations with those of the self consistent field calculations. Taking account of interfacial fluctuations we find good agreement, but the strength of the fluctuation effect is smaller than expected. We close with a discussion of our findings. ## II Model and techniques In the following we consider a melt of symmetric $`AB`$ diblock copolymers confined into a thin film. The chain length is denoted by $`N`$. One half of the diblock consists of $`A`$ monomers, while the other half consists $`B`$ monomers. There is a short range repulsion $`\chi `$ between $`A`$ and $`B`$ monomers, which drives the microphase separation. The lateral extension of the film is denoted as $`L`$, while $`\mathrm{\Delta }_0`$ is the distance between the parallel, hard confining surfaces. The monomer density in the middle of the film is denoted by $`\rho `$. We employ the bond fluctuation model (BFM) for the Monte Carlo (MC) simulations. In this coarse grained lattice model an effective monomer blocks all eight corners of a unit cube from further occupancy. We work at a monomer number density of $`\rho =1/16`$; a value which corresponds to a concentrated solution or melt. Monomers along the chain are connected via bond vectors from the set , , , , , including all permutations and sign combinations. The large number of bond vectors allows 87 distinct bond angles and gives a rather good approximation of continuous space properties. Monomers interact via a square well potential which is extended over the nearest 54 lattice sites, which constitute the first neighbor shell in the monomer density pair correlation function. Monomers of the same species attract each other with strength $`ϵ`$ while a contact in the range of the square well potential between unlike species increases the energy by an amount $`ϵ`$. The surfaces are parallel and impenetrable. An $`A`$ monomer in the two layers adjacent to the surfaces decreases the energy by $`ϵ_w`$, while a $`B`$ monomer in this region increases the energy by the same amount. In the MC simulations we use chain length $`N=32`$, and energy parameters $`ϵ=0.1769k_BT`$ and $`ϵ_w=0.1k_BT`$. This corresponds to intermediate segregation $`\chi N=30`$. The monomer wall interaction is rather weak; the $`A`$ component does not wet the surface. The configurations of the polymers are updated via local hopping attempts of individual monomers, slithering snake-like movements and exchanges of the identity $`AB`$ of the two blocks. This allows an efficient relaxation of the chain conformation. One Monte Carlo step consists of 3 slithering snake attempts per chain, 1 local hopping attempt per monomer, and 1 $`AB`$ flip per diblock. The self consistent field (SCF) calculations employ the Gaussian chain model. In order to map the bond fluctuation model onto the Gaussian chain model we choose the statistical segment length $`b`$ such that $`R^2=b^2(N1)289`$ equals the end-to-end distance of the chains in the Monte Carlo (MC) simulations in the athermal limit ($`ϵ=0`$). All distances are measured in units of the lattice spacing. In accord with previous studies we obtain $`b=3.05`$. Monomers of different species repel each other via a contact interaction of strength $`\chi `$. This Flory Huggins parameter $`\chi `$ is related to the potential well depth in the MC simulations via $`\chi =2z_cϵ/k_BT`$, where $`z_c`$ denotes the effective coordination number. The latter quantity has been extracted from the intermolecular paircorrelation function to be $`z_c=2.65`$. Hence, we estimate that $`ϵ=0.1769k_BT`$ corresponds to $`\chi N=30`$ in the Gaussian chain model. This correspondence between the parameters of the bond fluctuation model and the Gaussian chain model has proven useful for studying the bulk and interfacial properties of binary homopolymer blends and ternary mixtures of two homopolymers and a diblock copolymer. The effect of the confining surfaces is twofold. On the one hand the monomer density decays continuously from its value $`\rho `$ in the center of the film to zero in a narrow region of width $`\mathrm{\Delta }_w=0.15R_e`$ close to the surface. On the other hand, $`A`$ monomers in this surface region are attracted while $`B`$ monomers are repelled. The strength of this monomer wall interaction is chosen such that the surface energy in the MC simulations coincides with the value of the surface free energy in the SCF calculations in the limit that one species covers the surface completely and packing effects are neglected (cf. paper I). Using a SCF technique of Matsen we have numerically solved the model in mean field approximation. Further details of the MC simulations and the SCF calculations can be found in paper I. The phase diagram in the SCF calculations is presented in Fig.1. At high segregation, we observe an alternation of parallel oriented lamellae if the film thickness $`\mathrm{\Delta }`$ matches a multiple of the bulk period $`D_b`$, and perpendicular aligned lamellae at intermediate values of the film thickness. This behavior corresponds to the parameters of the simulation $`\chi N=30`$ and $`1.71\mathrm{\Delta }/R_e3.2`$. At lower segregation or larger film thickness, the frustration of the parallel lamellar phase due to a mismatch between the intrinsic length scale of the lamellar order and the confinement decreases, and one encounters direct transitions between parallel lamellar phases. In this case the perpendicular lamellar phase $`L_{}`$ and the adjacent parallel lamellar phase form a triple point. For very small film thickness $`\mathrm{\Delta }<3D_b`$ we find a second order transition between the perpendicular lamellar phase and the disordered phase. Note that there is no transition between the disordered state and the parallel ordered morphologies. Upon increasing the incompatibility the order propagates gradually from the surfaces into the bulk (cf. also Fig.7 in paper I). ## III Results. Several independent systems are quenched from $`ϵ=0`$ to $`ϵ=0.1769k_BT`$. For film thickness $`\mathrm{\Delta }_0=30`$ and $`\mathrm{\Delta }_0=56`$ they assemble into parallel oriented lamellar phases $`L_2`$ and $`L_4`$ with two or four $`AB`$ interfaces, while we find a perpendicular oriented lamellar phase for film thickness $`\mathrm{\Delta }_0=46`$. The Monte Carlo runs consists of at least $`3.810^6`$ Monte Carlo steps. We do not observe transitions between different morphologies within a simulation run and cannot rule out non-equilibrium effects completely. But the morphologies which are observed agree for all but one system of thickness $`\mathrm{\Delta }_0=56`$ with the predictions of the self-consistent field (SCF) theory. Moreover, the structures obtained are free of defects and the laterally averaged composition profiles of independent quenches agree to a high accuracy. This is illustrated in Fig.2 for $`\mathrm{\Delta }_0=46`$, where 6 systems have assembled into a perpendicular lamellar phase with a repeat distance $`D=1.936R_e`$. This value is about $`6\%`$ larger than the prediction of the SCF theory. The quantitative agreement of the composition profiles of independent systems shows that each system has sampled the fluctuations of the internal interfaces appropriately. ### A Interfacial fluctuations Fig.3 compares the two dimensional composition profiles in the perpendicular oriented lamellar phase at $`\chi N=30`$ and $`\mathrm{\Delta }/R_e=1.83`$. The walls ($`z`$-axis) run horizontally on the top and on the bottom of the figure. The $`x`$ axis across the film is vertical. In each configuration of the MC simulations we have located the instantaneous position of the lamellae in the $`z`$ direction, and shifted the $`z`$ coordinate such that the center of an $`A`$ lamella is located at $`z=0`$. These shifted profiles have been averaged in the $`y`$ direction and over all configurations. Hence, the profiles in the MC simulations are broadened by interfacial fluctuations with wavevectors in the $`xy`$ plane. The effect of these capillary waves is readily observed: the interfacial width in the MC simulations is significantly broader than in the SCF calculations. Similar deviations have also been observed in other studies. Apart from this broadening of the profiles, the MC simulations and the SCF calculations share many subtle details. In both cases the $`AB`$ interface bends in the vicinity of the wall as to increase the surface fraction covered with the component of the lower surface free energy. The bending of the interface in the vicinity of the surfaces interferes with the distortion which originates from the opposite surface and leads to an oscillation of the interfacial position. This has been observed in SCF calculation and is also born out in the MC simulations. In both cases the interfacial width increases in the vicinity of the surfaces. In the SCF calculations this is due to a reduction of the density at the surface, while in the MC simulation the effect stems from the finite interaction range; monomers at the surface have less neighbors to interact with. This gives rise to a negative line tension when the $`AB`$ interface approaches the surface. In binary blends the interfacial structure can be characterized by the local interfacial position $`u(x,y)`$, which depends on the lateral coordinates, and the “intrinsic” profiles of the order parameter across the interface. Neglecting the coupling between long wavelength fluctuations of the local interfacial position and the “intrinsic” profile, the latter can be calculated as a profile across an ideally flat interface. The SCF technique calculates these “intrinsic” profiles. The fluctuations of the local interfacial position in binary blends are well describable by the capillary wave Hamiltonian: $$[u(𝐫_{})]=\frac{\sigma _{\mathrm{eff}}}{2}\mathrm{d}^2𝐫_{}\left[u\right]^2$$ (1) where $`\sigma _{\mathrm{eff}}`$ denotes the effective interfacial tension between the unmixed phases and $`u`$ denotes the deviation of the local interfacial position from its lateral average. For unconfined interfaces in blends the value of the effective interfacial tension agrees well with independent measurements of the free energy costs of an $`AB`$ interface or SCF calculations. For interfaces in binary blends near a wall the effective interfacial tension extracted from the fluctuation spectrum depends on the distance between the interface and the wall, and the interaction between the interface and the wall imparts a long wavelength cut–off to the spectrum of interfacial fluctuations. We investigate the fluctuations of the local interfacial position in more detail for the parallel lamellar phases $`L_2`$ and $`L_4`$. In the MC simulations we have measured the local interfacial position via a block analysis. This is illustrated schematically in Fig.4(a). We divide the simulation box into lateral columns of size $`B\times B`$, and determine the local position of the interfaces in each column. The positions are Gaussian distributed around their lateral average across the system and $`s^2`$ denotes the variance of the distribution. Using the capillary wave Hamiltonian one finds: $$s^2=\frac{k_BT}{2\pi \sigma _{\mathrm{eff}}}\mathrm{ln}\left(\frac{L_{\mathrm{max}}}{B}\right)$$ (2) where $`L_{\mathrm{max}}`$ is the large length scale cut–off for interfacial fluctuations. If the interfaces were unconstrained this cut–off would be set by the lateral system extension $`L_{\mathrm{max}}=L`$. In the copolymer system neighboring interfaces interact and this might lead to a noticeable modification of the fluctuation spectrum. The results of this block analysis is presented in Fig.4(b), which displays the variance of the interfacial position as a function of the lateral coarse graining size $`B`$. Circles refer to the $`L_2`$ phase while squares correspond to the inner (in the middle of the film) and outer (close to the surfaces) interfaces of the $`L_4`$ phase. The fluctuations of the outer interfaces in the $`L_4`$ phase and the interfaces in the $`L_2`$ phase are similar. The spectrum of fluctuations is cut off at about $`L_{\mathrm{max}}40`$, indicating a rather strong interaction between the interfaces and the wall. The fluctuations for the inner interfaces of the $`L_4`$ phase are well describable by Eq.(2). The spectrum is cut off by the lateral system size $`L_{\mathrm{max}}=L=96`$. From the slope of the straight line we estimate the effective interfacial tension to be: $`\sigma _{\mathrm{eff}}=0.157k_BT`$. This value can be compared to the interfacial tension of an $`AB`$ interface in a binary blend $`\sigma _{AB}=\rho b\sqrt{\chi /6}(14\mathrm{ln}2/\chi N)0.0684k_BT`$. Using a similar fluctuation analysis, the effective interfacial tension of a single unconfined copolymer bilayer has been measured in the framework of the BFM at $`ϵ=0.15k_BT`$ or $`\chi N=25`$. The analysis found that the effective interfacial tension $`\sigma _{\mathrm{eff}}=0.03k_BT`$ of a copolymer bilayer with respect to undulations is much smaller than twice the interfacial tension in a binary blend $`2\sigma _{AB}=2\times 0.066k_BT`$ at that temperature. The latter finding is in accord with the expectation that the free energy cost of the interface in a copolymer melt is smaller than the interfacial tension in a blend. The interfacial tension in a blend is the limiting value at high segregation. Therefore, our finding for the confined copolymer melt suggests that interfacial fluctuations are strongly suppressed due to the confined geometry and mutual interactions between the interfaces. Estimating the additional free energy costs of interfacial fluctuations due to chain stretching, Semenov argued that equation (2) (with $`\sigma _{\mathrm{eff}}=\sigma _{AB}`$) is appropriate when the spectrum is cut off at length scales larger than the distance between lamellae. Using $`\sigma _{AB}`$, the width $`\pi w`$ of the $`AB`$ interface as a lower cut–off and the periodicity $`D`$ as an upper cut–off we obtain $`s=0.16R_e`$. This value is larger than our MC results (cf. next section). Our findings suggest that the effective interfacial tension differs from the interfacial tension in a binary polymer blend. Other treatments of waves on the surface of brushes result in wavevector dependent contributions to the free energy of fluctuations of the brush height. Laradji et al. have investigated fluctuations around the lamellar solution of the SCF theory in the bulk. Unlike the situation at an interface between partially miscible homopolymers, the eigenmodes of the fluctuations in the lamellar bulk phase are not capillary waves of individual interfaces but coherent modulations of the whole stack of interfaces. This points to a strong coupling between neighboring lamellae. Lacking a simple quantitative prediction for the fluctuation spectrum of an ensemble of strongly interacting interfaces in a confined geometry, however, we treat the strength of the interfacial fluctuations as an adjustable parameter in the following comparison. ### B Profiles The comparison between the composition profiles in the $`L_2`$ phase in the MC simulations and the SCF calculations is displayed in Fig.5(a). As discussed above, we find qualitative agreement, while there are quantitative deviations due to interfacial fluctuations. To account for these effects, we assume that the intrinsic profiles averaged over a small lateral patch of the interface are describable by the SCF calculations while the local position of the interface is Gaussian distributed. To supplement the SCF results with this additional broadening we convolute the profiles $`p`$ of any quantity with a Gaussian distribution: $$p_{\mathrm{cap}}(x)dx^{}p_{\mathrm{SCF}}(x^{})\frac{1}{\sqrt{2\pi s^2}}\mathrm{exp}([xx^{}]^2/2s^2)$$ (3) We employ the composition profile to adjust the width $`s`$ of the distribution as to achieve best agreement between the MC simulations and the SCF calculations. We then use the same value of $`s`$ to convolute profiles of all other quantities. Note that we assume the strength of the fluctuations to be uniform across the film. Certainly, this is only a rough approximation. As we have observed in Fig.4(b) the fluctuations of the inner lamellae are larger than the fluctuations of those close to the film surfaces in the $`L_4`$ phase. Hence, we expect $`s`$ to be larger in the middle of the film and smaller at the surfaces. Therefore, by convoluting the SCF profiles we overestimate the fluctuation effects at the film surfaces. Fig.5(b) shows the comparison between the MC simulations and the convoluted profiles of the SCF calculations with $`s/R_e=0.1`$. We achieve excellent agreement in the middle of the film and for the interfacial widths. The minor deviations at the surfaces are due to a reduced strength of interfacial fluctuations and the different structure at the surfaces. The panels (c) and (d) present the comparison between the MC results and the unconvoluted and convoluted SCF profiles for individual segment densities. Again, the convolution improves the agreement between the MC results and the SCF calculations. The $`A`$ ends are enriched at both surfaces, while the $`B`$ ends are located mostly in the middle of the film. The maximum at the center is, however, not very pronounced. This indicates that the strong stretching limit is not yet reached for $`\chi N=30`$ and the two brushes formed by the copolymers largely interdigitate. The profiles of the $`A`$ middle segment (monomer number 16) and the $`B`$ middle segment (monomer number 17) exhibit a maximum at the interfaces. In accord with experimental observations, the distribution of the middle segments is narrower than the distribution of the chain ends. A similar analysis has been performed for the $`L_4`$ phase and the $`L_{}`$ phase. The results are presented in Figs.6 and 7, respectively. The profiles in the $`L_4`$ phase are qualitatively similar to those of the $`L_2`$ phase. In the perpendicular phase $`L_{}`$ the profiles run perpendicular to the direction of the $`AB`$ interfaces of the lamellae and parallel to the walls. Since the $`A`$ component is slightly enriched at both surfaces also in the $`B`$ lamellae, the profile averaged across the film is less segregated in the $`B`$ lamellae than in the $`A`$ lamellae. This effect is observed in the MC simulations and the SCF profiles. For each phase we adjusted the strength of the interfacial fluctuations as to match the SCF profiles onto the MC data. For the profiles of all other quantities we used the same value of $`s`$, and thereby improved the agreement between the MC simulations and the SCF calculations. This procedure results in $`s/R_e=0.1`$, $`s/R_e=0.12`$, and $`s/R_e=0.127`$ for the $`L_2`$,$`L_4`$ and $`L_{}`$ phase, respectively. The increase of the strength of fluctuations is compatible with the intuition that the confined geometry reduces the interfacial fluctuations more in the $`L_2`$ phase than in the $`L_4`$ or $`L_{}`$ phase. Using these values of $`s`$ we can estimate the lateral block size $`B_{\mathrm{min}}`$ on which the MC data agree with the “intrinsic” profiles of the SCF calculations according to Fig.4(b). This yields the rough estimate $`B_{\mathrm{min}}10`$. Semenov suggested this cut–off to be $`B_{\mathrm{min}}=\pi w`$ for binary polymer blends and a recent MC study in the framework of the BFM found $`B_{\mathrm{min}}=1.2\pi w(13.1/\chi N)`$. Using $`w=b/\sqrt{6\chi }`$, we obtain $`B_{\mathrm{min}}4.5`$. This is again an indication that the interfacial fluctuations are not well describable with the capillary wave Hamiltonian in the confined copolymer system. One possible explanation for a larger value of the cut–off $`B_{\mathrm{min}}`$ is, e.g., a bending rigidity of the lamellae. ## IV Summary. We have presented SCF calculations and MC simulations for symmetric diblock copolymers confined into a thin film. Both surfaces attract the same component of the diblock via a short range potential. We have quenched several independent systems from the athermal state to $`\chi N=30`$, and we have compared the results of the MC simulations in the framework of the BFM with SCF calculations in the Gaussian chain model. We find qualitative agreement between the MC simulations and the SCF calculations. In particular, we observed the $`L_2`$, $`L_{}`$ and $`L_4`$ phases as predicted by the SCF calculations (cf. paper I). We have used these configurations to investigate the detailed structure in the thin film geometry. In the simulations we find evidence for a broadening of the profiles due to interfacial fluctuations. However, the spectrum of interfacial fluctuations is not well describable by the capillary wave Hamiltonian. The effective interfacial tension is higher than the interfacial tension in a binary blend of $`A`$ and $`B`$ homopolymers, and the spectrum is cut—off at large wave lengths. A similar increase of the effective interfacial tension of an interface in a binary polymer blend in the vicinity of a wall has been observed, however, the effect is more pronounced in the copolymer system. The confinement and the mutual interactions between neighboring lamellae strongly suppress interfacial fluctuations. To mimic the effect of interfacial fluctuations we convolute the SCF profiles with a Gaussian. The strength of the fluctuations is treated as a free parameter for each film thickness, however, we use the same value for profiles of different quantities. Taking account of interfacial fluctuations, we find almost quantitative agreement between the MC results and the SCF profiles for the composition, the density of chain ends and middle segments. In agreement with experiments both MC simulations and SCF calculations show that the middle segements of the copolymer are stronger localized at the $`AB`$ interface than the ends of the diblock in the middle of the domains. The good agreement between MC simulations and SCF calculations suggests that composition fluctuations play only a minor role at intermediate segregation. Moreover, the local fluid structure of the bond fluctuation model does not have a large influence on the profiles. The latter finding is rather unexpected, because the interactions on the monomer scale $`\chi =30/32`$ are large and the interfacial width between the $`A`$ and $`B`$ domains is not much larger than the microscopic length of the model. ### Acknowledgment It is a great pleasure to thank P.K. Janert, F. Schmid, and M.W. Matsen for valuable discussions/correspondence. We acknowledge generous access to the CRAY T3E at the HLR Stuttgart and HLRZ Jülich, as well as access to the CONVEX SPP at the computing center in Mainz. Financial support was provided by the DFG under grant Bi314/17.
no-problem/9906/cond-mat9906237.html
ar5iv
text
# Thermodynamic properties of the 𝑆⁢𝑂⁢(5) theory for the antiferromagnetism and 𝑑-wave superconductivity: a Monte Carlo study ## I Introduction High-$`T_c`$ superconductivity (SC) in cuprates is achieved by hole doping from the insulating state of antiferromagnetism (AF). The AF and SC phases are proximate each other in temperature vs. hole-doping-rate phase diagrams. Enhancement of AF correlations is observed above the SC transition temperature in the underdoped region . A clear signal from these experimental facts is the importance of the inter-relationship between these two very different and even expelling, at a first glance, properties. To explain theoretically the complex phase diagrams of the high-$`T_c`$ SC is still very challenging for the condensed matter physics. This problem has been approached using microscopic models, such as Hubbard hamiltonian and the simplified $`tJ`$ hamiltonian . It has been tried to derive microscopically the attractive force necessary for Cooper-pair formation from the magnetic interactions, and to construct the phase diagram with the AF and SC phases side by side. Up to date, however, there is no well accepted microscopic theory which can count for the most important features of the high-$`T_c`$ SC both qualitatively and quantitatively. In the $`SO(5)`$ theory this problem is approached in another way : The long-range SC and AF orders are presumed as the two possible long-range orders in pure systems. The three components of the AF order parameter and the real and imaginary parts of the SC order parameter compose a five-component superspin of $`SO(5)`$ symmetry. At low temperatures the $`SO(5)`$ symmetry is broken into two subspaces, the $`SO(3)`$ one associated with AF, and the $`U(1)`$ one associated with SC. The destination of the broken symmetry is controlled by the doping rate, or the chemical potential of holes. Therefore, the doping rate plays the role of $`SO(5)`$-symmetry breaking field. Much interest has been stimulated by the proposal of the $`SO(5)`$ theory, and considerable progresses in exploring this theory have been achieved . Since the superspin vector in the $`SO(5)`$ theory is of five dimensions, three for AF and two for SC, entropy effects on the competition between these two long-range orders are highly nontrivial. Thermal fluctuations are very crucial in determining phase diagrams . Therefore, investigation of the $`SO(5)`$ theory at finite temperatures is essential for a comprehensive understanding of the theory. Ultimately one should compare the predictions by the theory with phase diagrams observed experimentally. The $`SO(5)`$ theory also raises interests in the point of view of phase transitions and critical phenomena. To reveal the thermodynamic properties of the $`SO(5)`$ theory is the objective of the present study. Another important issue is the competition between the long-range SC order in the presence of an external magnetic field, realized in the flux-line lattice (FLL), and the long-range AF order. Since many interesting magnetic-field responses have been clarified in the long-range AF and SC orders separately, the proximity of them in high-$`T_c`$ cuprates is very likely to produce more sophisticated phenomena. Actually, it is suggested that vortices induced by an external magnetic field in high-$`T_c`$ superconductors may possess AF cores . To explore the vortex states in high-$`T_c`$ cuprates in the scheme of SO(5) theory is also very important. In order to achieve the above purposes, Monte Carlo simulations on a classical model hamiltonian in three-dimensional (3D) space are performed . The present approach takes into account thermal fluctuations both in the rotation of $`SO(5)`$ superspins between the AF and SC subspaces, and in the phase variables of SC order parameters. The remainder of this paper is organized as follows: The hamiltonian is presented in Sec. II, with descriptions on technical details of simulation. In Sec. III, simulation results for the null external magnetic field are presented. There are the AF, AF and SC phase-separation, and SC phases in the phase diagrams. The spin-gap phenomenon is also addressed. Section IV is devoted to reveal the effects of an external magnetic field in the $`SO(5)`$ theory. Coexistence between the long-range AF order and the FLL of long-range SC order is observed. Vortex cores are found of larger AF components than elsewhere. Summary is given in Sec. V. ## II Hamiltonian and simulation techniques The hamiltonian in the present study is given by $$=\underset{i,j}{}J_{i,j}^{SC}𝐭_i𝐭_j+\underset{i,j}{}J_{i,j}^{AF}𝐬_i𝐬_j+g\underset{i}{}𝐬_i^2,$$ (1) defined on the simple cubic lattice. The vector t, of two components and coupling ferromagnetically with nearest neighbors, is for the $`d`$-wave SC order parameter; the vector s, of three components and with AF coupling between nearest neighbors, is for the AF order parameter. The interplay between the SC and AF order parameters is introduced by the $`SO(5)`$ constraint on the superspin: $$𝐬_i^2+𝐭_i^2=1.$$ (2) The $`g`$ factor is a field breaking the $`SO(5)`$ symmetry into the $`U(1)`$ and $`SO(3)`$ subgroups, and is proportional to the doping rate in a loose sense . The following notes on the above hamiltonian seem appropriate at this stage. First, the above hamiltonian can be considered as the Ginzburg-Landau description of the $`SO(5)`$ theory. Both of the AF and SC order parameters, s and t, are defined in a scale larger than the atomic one, but much smaller than the macroscopic one. In this sense they should be called as the local order parameters. The constraint (2) does not imply the existence of long-range order in the macroscopic scale. The long-range order parameter for the AF component is the staggered magnetization, and that for the SC component is the helicity modulus . Second, although no quantum effect is included explicitly in hamiltonian (1), the competition between the two different long-range orders is taken into account sufficiently. Therefore, the profound, nontrivial thermodynamic properties of the $`SO(5)`$ theory can be captured. Third, thermal fluctuations in phase variables of SC order parameters, which are especially important for underdoped high-$`T_c`$ cuprates , are taken into account by the first term in the above hamiltonian, and treated using the Monte Carlo technique. Furthermore, this hamiltonian is easily developed so as to incorporate an external magnetic field for the study of vortex states. Fourth, the superspin amplitude is fixed to unity in the above hamiltonian. The onset of superspin amplitude itself upon cooling can also be taken into account in the mean-field fashion, and is expected to correspond to the so-called pseudo-gap phenomenon . Finally, only the simplest symmetry-breaking field $`g`$ associated with the quadratic terms of order parameters is included in the hamiltonian. Other symmetry-breaking fields appear when high orders of the order parameters are considered . Although an argument on magnitudes of these fields is absent right now, it is reasonably expected that the most important features of the breaking of $`SO(5)`$ symmetry into the AF and SC subspaces are captured by the $`g`$ field in (1). A typical simulation process starts from a random configuration of superspins at a sufficiently high temperature. The system is then cooled gradually. The equilibrium state at a given temperature is generated using typically 50,000 MC sweeps of update from the state of a slightly higher temperature. In each sweep of update, candidate vectors are generated randomly on the five-dimensional unit sphere for superspins on all sites in the system, and are subject to the standard Metropolis algorithm to determine if they are accepted for the next configuration . After this equilibriation process, statistics on physical quantities is performed over 100,000 MC sweeps. Around transition temperatures, more than $`10^6`$ MC sweeps are spent in order to make sure of sufficient equilibriation and statistics. The system size for simulations in null external magnetic field is $`L^3=40^3`$, with periodic boundary conditions in all crystal directions. As the $`SO(5)`$ superspins are continuous in five dimensions, and the system is of three dimensions in crystal space, a thorough analysis of finite-size effects on simulation results, which is important for determining the relevant critical and bicritical exponents in high precisions, is extremely time consuming. Only for several chosen parameter sets, larger systems have been simulated in order to make sure that the main properties derived from the present simulations do not suffer from finite-size effects. Systematic errors (finite-size effects) are therefore not estimated for data presented in this paper. Statistical errors are comparable to sizes of marks in figures as far as not specified. The AF coupling in the $`ab`$ plane $`J_{ab}^{AF}J`$ is taken as the energy unit, and temperature is measured by $`J/k_B`$ throughout the present paper. ## III Phase diagrams and correlation functions for $`H=0`$ ### A Isotropic system: $`J_{ab}^{SC}=J_c^{SC}=J_c^{AF}=J_{ab}^{AF}J`$ Figure 1 is the temperature vs. $`g`$-field phase diagram of the system with the same AF and SC coupling in all crystal directions. Both the N/AF and N/SC phase transitions are of second order, in the 3D Heisenberg and XY universality class, respectively. The two phase boundaries merge tangentially at the bicritical point $`[g_b,T_b]=[0,0.85J/k_B]`$ . For $`g=g_b`$ and $`T>T_b`$, the AF and SC correlation lengths for the two-point correlation functions are equal to each other, and isotropic in all crystal directions; the weights of AF and SC components are $`3/5`$ and $`2/5`$, proportional to the number of degrees of freedom. Away from the $`SO(5)`$-symmetric line, positive (negative) $`g`$ fields suppress AF (SC) correlations at all temperatures. ### B Anisotropic system: $`J_{ab}^{SC}=10J_c^{SC}=J_c^{AF}=J`$ The temperature vs. $`g`$-field phase diagram of the system of couplings $`J_{ab}^{SC}=J_c^{AF}=J`$ and $`J_c^{SC}=0.1J`$ is presented in Fig. 2. The bicritical point is at $`[g_b,T_b]=[1.18J,0.64J/k_B]`$. The equal-weight partition of the superspin at $`g=g_b`$ observed in the isotropic system is broken. Nevertheless, as indicated in the inset of Fig. 2, in the $`ab`$ plane the AF correlation length is equal to the SC correlation length when the $`g`$ field is fixed at the bicritical value. This agreement is not trivial in contrast with the isotropic system. The SC correlation length in the $`c`$ axis is much smaller than the other correlation lengths. ### C Strongly anisotropic system: $`10J_{ab}^{SC}=100J_c^{SC}=100J_c^{AF}=J`$ In order to simulate real high-$`T_c`$ cuprates, the AF exchange coupling should be taken much stronger than the effective SC coupling, and both AF and SC couplings are much weaker in the $`c`$ axis. The temperature dependence of the AF staggered magnetization and the helicity modulus of the SC components are shown in Fig. 3 for the system of couplings $`J_{ab}^{SC}=0.1J`$ and $`J_c^{SC}=J_c^{AF}=0.01J`$ at the symmetry breaking field $`g=1.96J`$. Since the helicity modulus is proportional to the superfluid density , it is clear that the long-range SC order is established below the critical temperature $`T_c0.115J/k_B`$. As shown in the same figure, the AF correlation length in the $`ab`$ plane, $`\xi _{ab}^{AF}`$, increases at first as temperature is reduced, and then is suppressed as temperature approaches $`T_c`$. The maximal $`\xi _{ab}^{SC}`$ is taken at the temperature $`T_{sg}0.15J/k_B`$. The weight of AF components, $`𝐬^2`$, decreases monotonically in the whole cooling process and shows a sharp decline among $`T_{sg}`$ and $`T_c`$. Therefore, the enhancement of $`\xi _{ab}^{AF}`$ above $`T_{sg}`$ is clearly the result of reduction of thermal fluctuations; the suppression of $`\xi _{ab}^{AF}`$ below $`T_{sg}`$ is because of the loss of the AF order in its competition with the SC order. This peculiar behavior occurs because the AF coupling in the $`ab`$ plane overwhelms over the SC one, while the SC groundstate is established by the large $`g`$ field. Temperature dependence of the internal energy and the specific heat for this system are depicted in Fig. 4. No feature can be found around $`T_{sg}`$ in these two thermodynamic quantities. Therefore, the only phase transition takes place at $`T_c`$, and $`T_{sg}`$ corresponds merely to a crossover. The SC correlation length in the $`ab`$ plane, $`\xi _{ab}^{SC}`$, diverges when temperature approaches $`T_c`$ in Fig. 3, as usually in a thermodynamic second-order phase transition. It is found experimentally that the spin-lattice relaxation rate assumes its maximum at a temperature well above the SC critical point . The present simulation results indicate that this spin-gap phenomenon can be explained by the competition among the long-range SC and AF orders, and thermal fluctuations. Since the enhancement of AF correlations above the SC critical point is observed in the strongly anisotropic system of Fig. 3, but not in isotropic and slightly anisotropic systems of Figs. 1, and 2, it becomes clear that in order to observe the spin-gap behavior, the system should have SC couplings much weaker than AF ones, as in real high-$`T_c`$ cuprates. Figure 5 is the temperature vs. $`g`$-field phase diagram of the same couplings for Figs. 3 and 4. The bicritical point is at $`[g_b,T_b]=[1.93J,0.12J/k_B]`$. The latent heat associated with the first-order transition between the AF and SC phases is approximately $`Q0.05J`$, and decreases to zero as the bicritical point is approached. The spin-gap like phenomenon is observed in the region $`g_b<g<2.2J`$. For $`g>2.2J`$, AF correlations are suppressed by SC components at all temperatures. The experimental fact that spin-gap behaviors are observed only in the underdoped region of high-$`T_c`$ cuprates may be explained by the present simulation result. The ratio between the spin-gap temperature and the SC critical point is $`T_{sg}/T_c1.6`$ at $`g=2J`$, which counts well the experimental observation . In Fig. 5, the spin-gap temperature $`T_{sg}`$ decreases as the bicritical point is approached. This might seem curious at a first glance, since it is clear from hamiltonian (1) that the larger the $`g`$ field the smaller the AF components. Shown in Figs. 6 (A) and (B) are the temperature dependence of the AF correlation length and the staggered susceptibility at several $`g`$ fields. Although both of them are monotonically suppressed by increasing $`g`$ field when temperature is fixed, the temperature where they take maxima, $`T_{sg}`$, increases with the $`g`$ field, as clearly seen in Figs. 6. It is noted that the spin-gap temperature increases with decreasing doping rate in experiments. The present theory therefore conflicts with experimental observations in this aspect. The SC correlations are suppressed in the normal state above the AF phase boundary in the present system. In this sense, there is no counterpart of the spin-gap temperature above Néel points. However, it is interesting to observe in Fig. 7 that for the $`g`$ field in a certain region below the bicritical value, the SC weight, $`𝐭^2`$, takes maximum at a temperature above the corresponding Néel point. The temperature associated with the maximal SC weight, denoted by $`T_p`$ in Fig. 5, may be identified with the pairing temperature . There is no feature in the internal energy and the specific heat around this crossover temperature. ## IV Phase diagram and vortex states for $`H>H_{c1}`$ ### A Model hamiltonian and phase diagram An external magnetic field penetrates into a type-II superconductor via thin flux lines associated with flux quanta for $`H`$ larger than the lower critical field $`H_{c1}`$. SC is broken along the flux lines. High-$`T_c`$ superconductors are extremely type-II with very large Ginzburg-Landau numbers $`\kappa 100`$. Research of the vortex states in high-$`T_c`$ SC has been growing into a vivid field of condensed matter physics and statistics. The most important feature of the vortex states is that the Abrikosov FLL melts into FL liquid via a first-order phase transition . In the scheme of the $`SO(5)`$ theory, the free energy of a vortex state can be reduced by rotating the superspins from the SC subspace into the AF subspace at the flux-line cores . The possibility of AF cores of flux lines in the $`SO(5)`$ theory was first addressed by Arovas et al. . Recently, Alama et al. discussed the $`\kappa `$ dependence of the core state . In these studies, the Abrikosov mean-field theory was developed so as to incorporate the AF components. However, the Abrikosov mean-field theory for the vortex states is not appropriate for the high-$`T_c`$ SC since it only takes into account the amplitude of SC order parameter, and cannot treat thermal fluctuations in the phase variables, which are essentially important for determining the phase diagram of the vortex states in high-$`T_c`$ SC . The hamiltonian for the $`SO(5)`$ theory in the presence of an external magnetic field may be given as following : $$=\underset{i,j}{}J_{ij}^{SC}|𝐭_i||𝐭_j|\mathrm{cos}\left(\phi _i\phi _jA_{ij}\right)+\underset{i,j}{}J_{ij}^{AF}𝐬_i𝐬_j\underset{i}{}𝐇𝐬_i+g\underset{i}{}𝐬_i^2,A_{ij}=\frac{2\pi }{\varphi _0}_i^j𝐀𝑑𝐫,$$ (3) where $`|𝐭|`$ and $`\phi `$ are the amplitude and phase of the SC order parameter. The same constraint (2) is applied. Fluctuations of the magnetic induction are neglected. This approximation is justified when the separation between vortices is larger than the SC correlation length and much smaller than the penetration depth in the $`ab`$ plane, a condition satisfied in large portion of $`HT`$ phase diagrams of high-$`T_c`$ superconductors. The Josephson coupling should also be dominant over the electromagnetic coupling. The first term in the above hamiltonian, known as the fully frustrated 3D XY model, has been used successfully for explaining many important thermodynamic properties of the vortex states in high-$`T_c`$ SC . In order to simplify the situation, the case of an external magnetic field parallel to the $`c`$ axis is addressed in the present paper . The vector potential is given by $`𝐀=(yB/2,xB/2,0)`$ with $`B=f\varphi _0/l_{ab}^2`$. Here, $`\varphi _0`$ is the flux quantum, $`l_{ab}`$ the unit length in the $`ab`$ plane, and $`f`$ the average number of flux in each square unit cell in the $`ab`$ plane. The data shown in the following are for $`f=1/25`$, corresponding to the inter-vortex distance of $`d_v=\sqrt{2/\sqrt{3}}l_{ab}/\sqrt{f}5.37l_{ab}`$ in the triangular FLL. The system size is chosen as $`L_a\times L_b\times L_c=50\times 50\times 40`$ with periodic boundary conditions in all crystal directions . Although the relation between the magnetic induction $`B`$ and the Zeeman field $`H`$ is not clear, the value of the Zeeman field is not much relevant to the following discussions, and thus is fixed to $`H=0.1J`$. In the present approach, vortices are defined by topological singularities in the configuration of phase variables of SC order parameters: $`_{\mathrm{cell}}(\phi _i\phi _jA_{ij})=(nf)2\pi `$, where $`n`$ is the vorticity. The temperature vs. $`g`$-field phase diagram of the system with couplings $`J_{ab}^{SC}=J_c^{AF}=J`$ and $`J_c^{SC}=0.1J`$ is depicted in Fig. 8. There are three ordered phases, namely the AF phase, AF and FLL coexistence phase, and FLL phase. The onset of long-range SC order is a first-order phase transition, same as those in systems of no AF components : At the melting temperature $`T_m`$ the FL liquid is frozen into the triangular FLL; the helicity modulus along the $`c`$ axis jumps sharply from zero to a finite value; there is a $`\delta `$-function peak in the specific heat, associated with a small latent heat. The onset of the long-range AF order is always a second-order phase transition . From the comparison between the two phase diagrams in Figs. 2 and 8 of same couplings, it is clear that suppression of the transition temperature of the long-range SC order by the external magnetic field is much more significant than that of the long-range AF order. This difference is understood easily considering the structures of these two long-range orders: For the long-range AF order, the magnetic spins are aligned antiferromagnetically, almost within the $`ab`$ plane. This configuration reduces the influence of the magnetic field on the onset of long-range AF order. On the other hand, the external magnetic field induces flux lines in the SC state, and produces strong fluctuations in phase variables of SC order parameters. The long-range SC order is established only when the flux lines are frozen into FLL. The above difference in the magnetic-field responses results also in the expansion of the AF phase into the SC territory, as can be seen in Figs. 2 and 8. In the region $`1.0Jg1.32J`$, the long-range AF and SC orders coexist at low temperatures. The temperature dependence of the helicity modulus along the $`c`$ axis, the staggered magnetization, and the specific heat are shown in Fig. 9 for $`g=1.1J`$. The N/AF transition at $`T_N`$ is a thermodynamic second-order phase transition, above the first-order onset of the long-range SC order at $`T_m`$. Although suppressed by the SC order in certain degree, the staggered magnetization survives to groundstate. The AF phase boundary drops sharply from $`[g,T]=[1.32J,0.59J/k_B]`$ in Fig. 8. The phase transition at this almost vertical part of AF phase boundary is investigated by tuning the $`g`$ field at a fixed temperature, in addition to the cooling process mentioned in Sec. II. The turning point on the N/AF phase boundary in Fig. 8 and the tricritical point in Ref. are at the same temperature. ### B AF vortex cores The vortex cores in the SC phase of the $`SO(5)`$ theory are different from those without AF competition studied up to date. The structure factors $`S(𝐪_{ab},z=0)`$ for vortices, $`s^2`$, $`s_{ab}`$, and $`s_c`$ for $`g=1.5J`$ in the FLL phase are displayed in Figs. 10. The Bragg peaks in the structure factor Fig. 10(A) for the vortex correlations are from the triangular FLL. One also finds Bragg peaks in structure factor Fig. 10(B) for the AF amplitudes at the same wave numbers of Fig. 10(A). This coincidence indicates clearly that cores of flux lines are of larger AF weights than elsewhere. The $`g`$-field dependence of the Bragg-spot height for AF weights in the FLL phase, such as those in Fig. 10(B), is investigated when temperature is fixed. As shown in Fig. 11 for $`T=0.3J/k_B`$, $`S(𝐪_{ab}=𝐪_{max},z=0)`$ decays with increasing $`g`$ field in a power law $`Sp/g^q`$ with $`p=0.8\pm 0.05`$ and $`q=3\pm 0.1`$. The haloes at the wave numbers $`Q=(\pm \pi ,\pm \pi )`$ in the structure factors for $`s_{ab}`$ and $`s_c`$ in Figs. 10(C) and (D) correspond to short-range AF fluctuations. The weak spot at $`Q=(0,0)`$ in Fig. 10(D) is from the small ferromagnetic component $`s_c`$ induced by the external magnetic field. The structure factors $`S(𝐪_{ab},z=0)`$ in the AF and FLL coexistence phase for vortices, $`s^2`$, $`s_{ab}`$, and $`s_c`$ are displayed in Figs. 12. From the structure factors Figs. 12(A) and (B) for vortices and AF amplitudes, it is clear that AF components are enhanced in cores of flux lines, as in the FLL phase. In structure factors Figs. 12(C) and (D) for $`s_{ab}`$ and $`s_c`$, there are strong Bragg peaks at $`Q=(\pm \pi ,\pm \pi )`$ associated with the long-range AF order. Satellite spots are observed around the main Bragg peaks in Figs. 12(C) and (D). These satellite spots are easily identified with those in Figs. 12(A) and (B). Therefore, in the AF and FLL coexistence phase, the phase of the long-range AF order is preserved in cores of flux lines. ## V Summary Thermodynamic properties of the $`SO(5)`$ theory are investigated using Monte Carlo simulations on a model hamiltonian which counts thermal fluctuations both in the rotations of superspins between the SC and AF subspaces, and in the phase variables of the SC order parameters. The latter factor is essentially important for explaining thermodynamic phase transitions associated with the onset of long-range SC order in high-$`T_c`$ cuprates in null and finite external magnetic fields. Therefore, the present approach is superior to Abrikosov-type mean-field treatments of the $`SO(5)`$ theory, in which thermal fluctuations in the phases of SC order parameters are neglected. For null external magnetic field, there is a bicritical point in the temperature vs. $`g`$-field phase diagram, at which the second-order N/AF and N/SC phase boundaries merge tangentially into the first-order AF/SC phase boundary. Hysteresis phenomenon is observed at the first-order AF/SC phase transition, which may suggest a phase-separation region in the phase diagram. In systems with much stronger AF couplings than SC ones while the SC groundstate is achieved by $`g`$ fields larger than the bicritical value, AF correlations are enhanced at a crossover temperature above the SC critical point. The origin of this enhancement in AF correlations is clarified to be the competition among the long-range AF and SC orders and thermal fluctuations. When the $`g`$ field becomes too large, this crossover fades away since AF correlations are suppressed at all temperatures by the large SC component. These results are consistent with the spin-gap phenomenon observed experimentally in the following aspects: First, real cuprates are very anisotropic in the AF and SC couplings $`J^{AF}0.1`$ eV and $`J^{SC}0.01`$ eV; Second, the spin-gap phenomenon has been observed experimentally only in the underdoped region. In contrast with experimental observations, however, the spin-gap temperature decreases as the bicritical point is approached from the SC side. Near the bicritical point, there is a crossover temperature above the Néel temperature where the weight of SC components takes maximum. The $`SO(5)`$ theory in an external magnetic field is also investigated. The long-range SC order is established through a first-order freezing transition from the FL liquid into the FLL, while the onset of the long-range AF order is associated with a second-order phase transition. These two phase boundaries cross each other, and thus produce a region in the phase diagram where the two long-range orders coexist. In the FLL phase, only short-range AF fluctuations are enhanced at cores of flux lines. 1D long-range AF order along the flux line cannot be realized because of strong thermal fluctuations. In the coexistence phase, superlattice spots surrounding the strong AF Bragg peaks at $`Q=(\pm \pi ,\pm \pi )`$ are observed in the simulated structure factor, and are identified with the modulation by the triangular flux-line lattice of SC. This simulation result can be checked by the neutron scattering technique. Acknowledgements The author would like to thank S.-C. Zhang, T. Koyama, and Y. K. Bang for stimulating conversations on the $`SO(5)`$ theory. M. Tachiki is very grateful for drawing author’s attention to vortex states in high-$`T_c`$ superconductivity and continuous encouragement. He appreciates S. Miyashita, N. Akaiwa, M. Itakura, and Y. Nonomura for helpful discussions on technical points of MC simulation. The present simulations are performed on the Numerical Materials Simulator (SX-4) of National Research Institute for Metals (NRIM), Japan. Figure Captions Fig. 1: Temperature vs. $`g`$-field phase diagram of the system with isotropic AF and SC couplings. The bicritical point is at $`[g_b,T_b]=[0,0.85J/k_B]`$. Fig. 2: Temperature vs. $`g`$-field phase diagram of the system with the couplings $`J_{ab}^{SC}=J_c^{AF}=J`$, and $`J_c^{SC}=0.1J`$. The bicritical point is at $`[g_b,T_b]=[1.18J,0.64J/k_B]`$. Inset: temperature dependence of the AF and SC correlation lengths in the $`ab`$ plane at $`g=g_b`$. Fig. 3: Temperature dependence of the AF and SC order parameters, correlation lengths, and the weight of AF components at $`g=1.96J`$ in the system with couplings $`J_{ab}^{SC}=0.1J`$ and $`J_c^{SC}=J_c^{AF}=0.01J`$. Here, $`T_c`$ is the SC transition point, and $`T_{sg}`$ is the spin-gap temperature. Fig. 4: Temperature dependence of the internal energy and the specific heat per site for the same system in Fig. 3. Fig. 5: Temperature vs. $`g`$-field phase diagram for the same couplings in Fig. 3. The bicritical point is at $`[g_b,T_b]=[1.93J,0.12J/k_B]`$. The spin-gap temperature $`T_{sg}`$ and pairing temperature $`T_p`$ fade away around $`g=2.2J`$ and $`g=1.5J`$ respectively. Fig. 6: Temperature dependence of the AF correlation length in the $`ab`$ plane (A) and the staggered susceptibility (B) at several typical $`g`$ fields. Maxima are assumed at the spin-gap temperatures $`T_{sg}`$ for $`g>g_b=1.93J`$. Fig. 7: Temperature dependence of the SC weight at a series of $`g`$ fields. Maxima are assumed at the pairing temperatures $`T_p`$ for $`1.5J<g<g_b`$. Fig. 8: Temperature vs. $`g`$-field phase diagram of the system with the couplings $`J_{ab}^{SC}=J_c^{AF}=J`$ and $`J_c^{SC}=0.1J`$. The flux density is given by $`f=1/25`$, and the Zeeman field is $`H=0.1J`$. Fig. 9: Temperature dependence of the helicity modulus along the $`c`$ axis, the staggered magnetization, and the specific heat per site in the system of the same couplings of Fig. 8 at $`g=1.1J`$. Fig. 10: Structure factors $`S(𝐪_{ab},z=0)`$ for vortices (A), $`s^2`$ (B), $`s_{ab}`$ (C), and $`s_c`$ (D) in the FLL phase in Fig. 8. Fig. 11: $`g`$-field dependence of the Bragg-spot height for the AF weights at $`T=0.3J/k_B`$. Fig. 12: Structure factors $`S(𝐪_{ab},z=0)`$ for vortices (A), $`s^2`$ (B), $`s_{ab}`$ (C), and $`s_c`$ (D) in the AF and FLL coexistence phase in Fig. 8.
no-problem/9906/hep-ph9906527.html
ar5iv
text
# 1 Introduction ## 1 Introduction In string theory, there are usually many flat directions that are expected to acquire a mass from supersymmetry (SUSY) breaking. The mass is of order the supersymmetry breaking mass; the fields are very light but have no collider implications because their interactions are suppressed by the Planck scale. However, moduli fields can be very important (and dangerous) from a cosmological vantage point . The moduli fields are expected to have a Planck scale amplitude in the early universe, and will therefore dominate the energy density of the universe as soon as they start to oscillate. However, the modulus lifetime for a Planck-coupled modulus field is so long that standard cosmological scenarios are adversely affected . Most important for generic moduli with mass of order the electroweak scale, it can be shown that the moduli decay occurs after big-bang nucleosynthesis (BBN), destroying the successful predictions. This problem is referred to as the “cosmological moduli problem.” One potential resolution of this problem is that the gravitino mass (i.e., the moduli mass) is larger than generically assumed; this requires the mass to be 10 $``$ 100 TeV . With this mass, the modulus lifetime can be shorter than 1 sec, so that decay occurs before BBN and the standard BBN is unaffected. In “standard” gravity-mediated scenarios, such a large modulus mass is difficult to understand. Furthermore, raising the modulus mass permitting a more rapid decay does not automatically lead to a successful cosmology. This is because the decay of a modulus field can produce a sizable number of superparticles which cascade down to the lightest superparticle (LSP). Assuming $`R`$-parity conservation, it has been claimed that the relic density of the LSP is likely to overclose the Universe . Recently, however, a novel framework for supersymmetry breaking has been proposed in which SUSY breaking parameters are generated by the super-Weyl anomaly effects and are therefore loop-suppressed relative to the standard “hidden sector” predictions. In particular, this scenario predicts the gaugino masses as $`m_{\mathrm{G}i}={\displaystyle \frac{b_ig_i^2}{16\pi ^2}}M,`$ (1.1) where $`g_i`$ are the gauge coupling constants with $`i=1,2,3`$ identifying the gauge group, and $`b_i`$ are the $`\beta `$-function coefficients of the gauge coupling constant. Furthermore, $`M`$ is the auxiliary field in the supergravity multiplet whose vacuum expectation value (VEV) is expected to be of order the gravitino mass $`m_{3/2}`$. The above relation tells us two important consequence of the anomaly-mediated mass spectrum. The first is that the Wino becomes the LSP, instead of the Bino which is the conventional candidate for the LSP. Indeed, substituting the weak scale values of $`g_i`$, the anomaly-mediated model predicts $`m_{\mathrm{G1}}:m_{\mathrm{G2}}:m_{\mathrm{G3}}3:1:10`$. Second, an important feature is that the gravitino is extremely heavy in this framework. Since the gaugino masses are one-loop suppressed relative to the gravitino mass, the gaugino masses at the electroweak scale require the gravitino mass to be 10 $``$ 100 TeV. In the scenario of Ref. , “sequestered sector SUSY breaking,” a consistent theory is presented in which the scalar masses are sufficiently light, despite the large gravitino mass. Therefore, in this framework, namely sequestered sector SUSY breaking with an anomaly-mediated mass spectrum for gauginos, it is quite natural to expect the large gravitino mass that solves the cosmological moduli problem. Some alternative possibilities for solving the moduli problem with a light modulus mass, such as using an enhanced symmetry point or late time inflation , have also been suggested. However, the heavy modulus mass is probably the simplest possibility. Furthermore, the problem of too large a residual LSP mass density (even in the presence of a large gravitino mass) assumed a Bino-like neutralino whose pair annihilation cross section is $`p`$-wave suppressed. However, if the LSP is not Bino-like, the interaction of the LSP changes and a larger pair annihilation cross section may be realized. This suppresses the mass density of the LSP. With the anomaly-mediated spectrum, the Wino-like neutralino is the LSP. Since the Wino has a larger pair annihilation cross section than the Bino, the mass density of the LSP can be sufficiently suppressed. It should be also noted that the number of the LSP produced by the decay of one modulus field is usually assumed to be $`O(1)`$. However, the number of produced LSP is model-dependent, so a much smaller number density of the LSP could be produced, as we will discuss in Appendix. Based on these observations, we will demonstrate that not only can the cosmological moduli problem be solved in the anomaly-mediated SUSY breaking (AMSB) scenarios, but, furthermore, the LSP relic density from modulus decay can be reduced to an acceptable level. In fact, if the parameters are right, the Wino is a perfect dark matter candidate. This has an important advantage from the point of view of detecting SUSY dark matter. Ordinarily, the thermal relic density and the anti-matter fluxes are both determined by the strength of the dark matter candidate’s coupling. A large detection efficiency requires a large coupling, while a large relic density requires a relatively small coupling in order to impede annihilation. This in general implies relatively low efficiency for detecting supersymmetric dark matter. In our scenario however, because the Winos are produced from moduli decay, there can be sufficiently many to comprise dark matter, despite the large cross section. This is good both from the vantage point of standard detection, and also for the new searches for anti-matter, particularly by the Alpha Magnetic Spectrometer (AMS). The outline of this paper is as follows. We calculate the mass density of the Wino-like LSP produced by the decay of the moduli fields in Section 2. We will see that the density parameter $`\mathrm{\Omega }_\chi `$ of the LSP can be in a reasonable range (0.1 $``$ 1) in some regions of parameter space, and hence the old problem of the overproduction of the LSP due to the modulus decay may be solved in AMSB scenario. Furthermore, this fact gives us a motivation to consider the relic Wino as the dominant component of cold dark matter (CDM). In Section 3, we discuss possible signals from Wino CDM. We will see that the larger Wino cross sections permit a much more optimistic scenario for the possibility of detecting SUSY dark matter than the more conventional type. In section 4, we conclude. In Appendix A, the properties of the moduli fields are discussed. ## 2 Mass Density of the Wino LSP We first discuss the cosmological evolution of the modulus field and the density of the LSP. In the very early Universe, the modulus field has a large amplitude, expected to be as large as the Planck scale. It begins to oscillate when the expansion rate $`H`$ of the Universe becomes comparable to $`m_\varphi `$.<sup>#1</sup><sup>#1</sup>#1Even if the initial amplitude of the modulus field is smaller than $`O(M_{})`$, the energy density of the Universe is dominated by that of the modulus field when $`\varphi `$ decays if the initial amplitude is larger than $`10^{12}`$ GeV. In this paper, we assume this is the case. After this period, the energy density of the Universe is dominated by that of the modulus field. Then, when $`H\mathrm{\Gamma }_\varphi `$, the modulus field decays. The decay products are quickly thermalized and the Universe is reheated. Furthermore, the decay of the modulus field produces LSP’s. Produced LSP’s also lose their energy by scattering off background particles and become non-relativistic. The evolution of the number density $`n_\chi `$ of the LSP is obtained by solving the following coupled Boltzmann equations: $`{\displaystyle \frac{dn_\chi }{dt}}+3Hn_\chi `$ $`=`$ $`\overline{N}_{\mathrm{LSP}}\mathrm{\Gamma }_\varphi n_\varphi v_{\mathrm{rel}}\sigma n_\chi ^2,`$ (2.1) $`{\displaystyle \frac{dn_\varphi }{dt}}+3Hn_\varphi `$ $`=`$ $`\mathrm{\Gamma }_\varphi n_\varphi ,`$ (2.2) $`{\displaystyle \frac{d\rho _{\mathrm{rad}}}{dt}}+4H\rho _{\mathrm{rad}}`$ $`=`$ $`(m_\varphi \overline{N}_{\mathrm{LSP}}m_\chi )\mathrm{\Gamma }_\varphi n_\varphi +2m_\chi v_{\mathrm{rel}}\sigma n_\chi ^2,`$ (2.3) where $`m_\chi `$ is the mass of the LSP, and $`\overline{N}_{\mathrm{LSP}}`$ is the averaged number of the LSP produced in the decay of one modulus field. Here, $`n_\varphi `$ is the number density of the modulus field which is related to the energy density of the modulus field $`\rho _\varphi `$ as $`\rho _\varphi =m_\varphi n_\varphi `$. The quantity $`\rho _{\mathrm{rad}}`$ is the energy density of the radiation which is related to the background temperature $`T`$ as $`\rho _{\mathrm{rad}}=\frac{\pi ^2}{30}g_{}T^4`$, where $`g_{}`$ is the effective number of the massless degrees of freedom. In our calculation, we use $`g_{}=10.75`$, since we consider a situation with a reheating temperature of $`T_\mathrm{R}O(110\text{MeV})`$. One important quantity in solving these Boltzmann equations is the thermally averaged annihilation cross section $`v_{\mathrm{rel}}\sigma `$. If the LSP is Bino-like, the annihilation is through $`p`$-wave processes and $`v_{\mathrm{rel}}\sigma `$ is suppressed. As a result, the decay of the modulus field overproduces LSP’s for a reasonable reheating temperature of $`T_\mathrm{R}O(110\text{MeV})`$ . However, in the case of the Wino LSP, the pair annihilation proceeds through an $`s`$-wave process. In particular, by exchanging a charged Wino, the neutral Wino (i.e., LSP) can annihilate into a $`W`$-boson pair. In the non-relativistic limit, the annihilation cross section is given by $`v_{\mathrm{rel}}\sigma _{\stackrel{~}{W}^0\stackrel{~}{W}^0W^+W^{}}={\displaystyle \frac{g_2^4}{2\pi }}{\displaystyle \frac{1}{m_\chi ^2}}{\displaystyle \frac{(1x_W)^{3/2}}{(2x_W)^2}},`$ (2.4) where $`x_Wm_W^2/m_\chi ^2`$, and $`g_2`$ is the gauge coupling constant of SU(2)<sub>L</sub>. In the following calculation, we use this formula for the annihilation cross section of the LSP.<sup>#2</sup><sup>#2</sup>#2We neglect the possible co-annihilation of charged and neutral Winos. If the Wino is in kinematic equilibrium, the number density of the charged Wino is extremely suppressed. This is because the mass splitting between charged and neutral Winos is of order 100 MeV $``$ 1 GeV , which is much larger than the temperature we are considering. In this case, our approximation is extremely well justified. Another important parameter is $`\mathrm{\Gamma }_\varphi `$, the decay width of the modulus field. Since the interaction of the modulus field is proportional to inverse powers of $`M_{}`$, where $`M_{}2.4\times 10^{18}\text{GeV}`$ is the reduced Planck scale, $`\mathrm{\Gamma }_\varphi `$ is extremely suppressed as seen in Appendix A. In order to discuss this in a model-independent way, we parameterize the decay width as $`\mathrm{\Gamma }_\varphi ={\displaystyle \frac{1}{2\pi }}{\displaystyle \frac{m_\varphi ^3}{\mathrm{\Lambda }_{\mathrm{eff}}^2}}.`$ (2.5) Here, $`\mathrm{\Lambda }_{\mathrm{eff}}`$ is the effective suppression scale for the interaction of the modulus field, which can be determined as a function of the couplings given in the previous section. For an unsuppressed two-body decay process, we expect $`\mathrm{\Lambda }_{\mathrm{eff}}M_{}`$. It is instructive to discuss the qualitative behavior of the solution to the Boltzmann equations (2.1) $``$ (2.3). Since the modulus field decays when the expansion rate becomes comparable to $`\mathrm{\Gamma }_\varphi `$, the reheating temperature is estimated as $`T_\mathrm{R}\left({\displaystyle \frac{\pi ^2}{90}}g_{}\right)^{1/4}\sqrt{\mathrm{\Gamma }_\varphi M_{}}7.7\text{MeV}\times \left({\displaystyle \frac{m_\varphi }{100\text{TeV}}}\right)^{3/2}\left({\displaystyle \frac{\mathrm{\Lambda }_{\mathrm{eff}}}{M_{}}}\right)^1\left({\displaystyle \frac{g_{}}{10.75}}\right)^{1/4},`$ (2.6) where we used the instantaneous decay approximation . This reheating temperature has to be larger than $`1\mathrm{MeV}`$ in order not to affect the success of standard big-bang nucleosynthesis . This is guaranteed for a modulus mass of $`100\mathrm{TeV}`$ with a naive two-body decay rate $`\mathrm{\Gamma }_\varphi \frac{1}{4\pi }\frac{m_\varphi ^3}{M_{}^2}`$. So, from the vantage point of the standard cosmological moduli problem, the sequestered sector scenario is very advantageous. In order to achieve this estimated two-body decay rate, the moduli should decay into gauge boson pairs, Higgs pairs, or gravitino pairs through the interactions given in Eqs. (A.1), (A.4), or (A.8), respectively, with unsuppressed coupling constant (i.e., $`\lambda 1`$). With the decays of the moduli fields, the LSP is produced. However, the evolution of the cosmological density of the LSP is different from the usual case. In the standard scenario, the produced LSP’s reach thermal equilibrium. Therefore, when the temperature $`T`$ is higher than $`m_\chi `$, the number density of the LSP is comparable to those of massless particles, while once $`T`$ becomes lower than $`m_\chi `$, $`n_\chi `$ is Boltzmann suppressed. Then, at some temperature $`T_{\mathrm{dec}}`$, the pair annihilation rate of the LSP becomes smaller than the expansion rate of the Universe, and the LSP decouples from the thermal bath; after this stage, the number of the LSP in a comoving volume is fixed. In this case, the thermal relic density of the Wino-like LSP is estimated as $`\mathrm{\Omega }_\chi ^{(\mathrm{thermal})}h^25\times 10^4\times \left({\displaystyle \frac{m_\chi }{100\mathrm{GeV}}}\right)^2,`$ (2.7) where $`h`$ is the present Hubble constant in units of 100 km/sec/Mpc. Obviously, in the standard scenario, the Wino LSP cannot be the dominant component of the CDM since the mass density given above is too small. If $`T_\mathrm{R}`$ is higher than $`T_{\mathrm{dec}}`$, the relic density of the Wino-like LSP is given by Eq. (2.7). However, the typical decoupling temperature is given by $`T_{\mathrm{dec}}\frac{1}{30}m_\chi `$ , and hence Eq. (2.6) tells us that $`T_\mathrm{R}`$ is much lower than $`T_{\mathrm{dec}}`$. In this case, the LSP from the modulus decay is never in chemical equilibrium, and its number density just decreases because of pair annihilation. However, the pair annihilation rate eventually becomes smaller than the expansion rate of the Universe, since the annihilation rate is proportional to the number density of the LSP. Once this happens, the pair annihilation is no longer effective, and the LSP freezes out. This happens when the annihilation term in Eq. (2.1) (i.e., the second term in RHS) becomes less significant than the dilution term (i.e., the second term in LHS). The number density is estimated as $`n_\chi ^{(\mathrm{ann})}(T_\mathrm{R}){\displaystyle \frac{3H}{2v_{\mathrm{rel}}\sigma }}|_{T=T_\mathrm{R}}{\displaystyle \frac{3\mathrm{\Gamma }_\varphi }{2v_{\mathrm{rel}}\sigma }}.`$ (2.8) The pair annihilation proceeds as long as $`n_\chi `$ is larger than $`n_\chi ^{(\mathrm{ann})}`$ given above, and hence $`n_\chi ^{(\mathrm{ann})}`$ is the upper bound on the number density of the LSP for a given $`T_\mathrm{R}`$. It is notable that this upper bound is insensitive to the mechanism for LSP production. With this relation, we obtain the mass density to entropy ratio as $`{\displaystyle \frac{\rho _\chi ^{(\mathrm{ann})}}{s}}`$ $``$ $`{\displaystyle \frac{m_\chi n_\chi ^{(\mathrm{ann})}(T_\mathrm{R})}{(2\pi ^2/45)g_{}T_\mathrm{R}^3}}`$ (2.9) $``$ $`1.3\times 10^9\text{GeV}\times {\displaystyle \frac{(2x_W)^2}{(1x_W)^{3/2}}}`$ $`\times \left({\displaystyle \frac{m_\chi }{100\text{GeV}}}\right)^3\left({\displaystyle \frac{m_\varphi }{100\text{TeV}}}\right)^{3/2}\left({\displaystyle \frac{\mathrm{\Lambda }_{\mathrm{eff}}}{M_{}}}\right)\left({\displaystyle \frac{g_{}}{10.75}}\right)^{1/4}.`$ Since we expect no entropy production after the decay of $`\varphi `$, the above ratio should be preserved until today. One can easily see that the ratio $`\rho _\chi ^{(\mathrm{ann})}/s`$ is proportional to $`T_\mathrm{R}^1`$. For $`T_\mathrm{R}T_{\mathrm{dec}}`$, Eq. (2.9) approximately reproduces the standard result given in Eq. (2.7). However, since the reheating temperature is much lower than $`T_{\mathrm{dec}}`$, we expect a significantly larger mass density of the LSP as a result of moduli decay. One should note that the above ratio is proportional to $`\mathrm{\Lambda }_{\mathrm{eff}}`$, and that the mass density becomes smaller as the modulus field interacts more strongly. The result given in Eq. (2.9) is valid only if there is a sufficiently large number of LSP’s produced by the decay of the modulus field. If there is insufficient production, the pair annihilation is not effective, and all the produced LSP’s survive. In this case, the number density of LSP’s is estimated as $`n_\chi ^{(0)}(T_\mathrm{R})\overline{N}_{\mathrm{LSP}}n_\varphi (T_\mathrm{R}){\displaystyle \frac{3\overline{N}_{\mathrm{LSP}}\mathrm{\Gamma }_\varphi ^2M_{}^2}{m_\varphi }},`$ (2.10) and hence $`{\displaystyle \frac{\rho _\chi ^{(0)}}{s}}`$ $``$ $`{\displaystyle \frac{m_\chi n_\chi ^{(0)}(T_\mathrm{R})}{(2\pi ^2/45)g_{}T_\mathrm{R}^3}}`$ (2.11) $``$ $`5.8\times 10^6\text{GeV}`$ $`\times \overline{N}_{\mathrm{LSP}}\left({\displaystyle \frac{m_\chi }{100\text{GeV}}}\right)\left({\displaystyle \frac{m_\varphi }{100\text{TeV}}}\right)^{1/2}\left({\displaystyle \frac{\mathrm{\Lambda }_{\mathrm{eff}}}{M_{}}}\right)^1\left({\displaystyle \frac{g_{}}{10.75}}\right)^{1/4}.`$ Notice that the mass density is proportional to $`\overline{N}_{\mathrm{LSP}}`$, the average number of LSP’s produced by one modulus decay. Using Eqs. (2.9) and (2.11), the actual mass density is estimated as $`{\displaystyle \frac{\rho _\chi }{s}}\text{min}({\displaystyle \frac{\rho _\chi ^{(0)}}{s}},{\displaystyle \frac{\rho _\chi ^{(\mathrm{ann})}}{s}}).`$ (2.12) By comparing the above quantity with the current critical density $`{\displaystyle \frac{\rho _\mathrm{c}}{s}}3.6\times 10^9\mathrm{GeV}\times h^2,`$ (2.13) we obtain the density parameter $`\mathrm{\Omega }_\chi \rho _\chi /\rho _\mathrm{c}`$. A more accurate estimation of the density parameter is given by solving the Boltzmann equations. We solved the Boltzmann equations (2.1) $``$ (2.3) numerically and calculated the density parameter $`\mathrm{\Omega }_\chi `$ as a function of $`\overline{N}_{\mathrm{LSP}}`$, $`m_\varphi `$, $`m_\chi `$, and $`\mathrm{\Lambda }_{\mathrm{eff}}`$. In our calculation, we followed the evolution from the modulus-dominated era to the radiation-dominated Universe where the temperature is much lower than $`T_\mathrm{R}`$ and the LSP has already frozen out. In Figs. 2 and 2, we plot the constant $`\mathrm{\Omega }_\chi h^2`$ contour as a function of $`m_\chi `$ and $`\overline{N}_{\mathrm{LSP}}`$ for $`m_\varphi =100\text{TeV}`$ and 300 TeV respectively. As we can see, for $`\overline{N}_{\mathrm{LSP}}1`$, $`\mathrm{\Omega }_\chi h^2`$ is almost independent of $`\overline{N}_{\mathrm{LSP}}`$ since the pair annihilation is important in this region. (See Eq. (2.9).) However, for smaller $`\overline{N}_{\mathrm{LSP}}`$, the pair annihilation process becomes ineffective and $`\mathrm{\Omega }_\chi h^2`$ becomes sensitive to $`\overline{N}_{\mathrm{LSP}}`$. (See Eq. (2.11).) With the natural value of $`\mathrm{\Lambda }_{\mathrm{eff}}M_{}`$, $`\mathrm{\Omega }_\chi h^2`$ can be 0.1 $``$ 10, which is smaller than the result of the conventional calculation with the Bino LSP. In the AMSB, the Wino is the LSP, and hence the pair annihilation among the LSP’s is more enhanced than the Bino LSP case. Furthermore, the LSP density can be significantly suppressed if $`\overline{N}_{\mathrm{LSP}}1`$. (Notice that the second effect has not been considered before, and it can be important even in the Bino LSP case.) Because of these two reasons, the AMSB scenario realizes a smaller mass density of LSP’s from moduli decay. In particular, $`\mathrm{\Omega }_\chi h^2`$ can be 0.1 $``$ 1 almost irrespective of the mass of the LSP for $`\overline{N}_{\mathrm{LSP}}10^310^4`$. In this case, we can avoid the problem of overclosing the Universe. Furthermore, even with $`\overline{N}_{\mathrm{LSP}}1`$, $`\mathrm{\Omega }_\chi h^2O(1)`$ if the Wino mass is less than about 200 GeV. So not only do we avoid the overproduction of Binos which is part of the cosmological moduli problem, but we actually see the the Wino is an excellent dark matter candidate. The success of this scenario depends on the precise structure of the operators through which the moduli decay. The possible operators are listed in Appendix A. If the moduli are lighter than twice the gravitino mass, the operator (A.4) should be the most important. The operator (A.1) might also be important; however its coefficient is very likely loop-suppressed. In either case, we would expect $`\overline{N}_{\mathrm{LSP}}10^310^4`$,<sup>#3</sup><sup>#3</sup>#3If the decay is to Higgses, one would obtain the LSP’s from the suppressed three-body decay for example. which is optimal for obtaining critical density for a large range of Wino mass. However, if $`F_\varphi `$ is large, both operators (A.1) and (A.4) would have small coefficients in order to avoid too large a gaugino mass and $`\mu `$-parameter in the sequestered sector scenario. In this case, the moduli must decay into gravitino pairs through the operator (A.8). This would give a sufficiently high reheating temperature to avoid the first of the cosmological moduli problems, but requires a small Wino mass to obtain critical density (and not more) for the relic Winos. In this case, one requires slightly heavier moduli fields. Finally, let us briefly discuss gravitino cosmology in the AMSB scenario. In the gravity-mediated SUSY breaking scenario, it is well known that a gravitino much lighter than $`10\mathrm{TeV}`$ is problematic . In particular, the gravitinos are produced in the early Universe through scattering processes, and their decay can spoil the great success of the standard BBN scenario if $`m_{3/2}10\mathrm{TeV}`$ . Furthermore, the LSP may be overproduced via the decay of the gravitino. However, in the sequestered sector case, these problems may be avoided since the gravitino behaves almost like the modulus field. In the sequestered sector scenario, the gravitino mass can be so large that the decay happens before BBN starts. Furthermore, the LSP is produced by the decay of the gravitino with $`\overline{N}_{\mathrm{LSP}}1`$, but the density of the LSP may be reduced by pair annihilation. The only difference is that the primordial number density of the gravitino is determined by the reheating temperature just after the primordial inflation, which we call $`T_\mathrm{R}^{(\mathrm{inf})}`$.<sup>#4</sup><sup>#4</sup>#4$`T_\mathrm{R}^{(\mathrm{inf})}`$ should be distinguished from $`T_\mathrm{R}`$, the reheating temperature just after the decay of the modulus field. If we can neglect the pair annihilation of the produced LSP, the cosmological abundance of the LSP is proportional to the primordial abundance of the gravitino. In this case, if there is no effect from the modulus field, the mass density of the LSP is given by $`{\displaystyle \frac{\rho _\chi ^{(\mathrm{grav})}}{s}}1.4\times 10^9\times \left({\displaystyle \frac{m_\chi }{100\text{GeV}}}\right)\left({\displaystyle \frac{T_\mathrm{R}^{(\mathrm{inf})}}{10^{11}\text{GeV}}}\right),`$ (2.14) where we used the gravitino abundance given in Ref. . However, once the number density becomes large enough, the pair annihilation becomes effective and the above formula is not valid any more. In this case, it is relevant to use $`\rho _\chi ^{(\mathrm{ann})}/s`$ given in Eq. (2.9), since this is the maximally allowed mass density. (For the gravitino case, in Eq. (2.9), $`\mathrm{\Lambda }_{\mathrm{eff}}M_{}`$ and $`m_\varphi =m_{3/2}`$.) Therefore, the mass density of the LSP in this case is estimated as $`{\displaystyle \frac{\rho _\chi }{s}}\text{min}({\displaystyle \frac{\rho _\chi ^{(\mathrm{grav})}}{s}},{\displaystyle \frac{\rho _\chi ^{(\mathrm{ann})}}{s}}).`$ (2.15) When $`T_{\mathrm{R}}^{(\mathrm{inf})}{}_{}{}^{>}10^{11}10^{12}\mathrm{GeV}`$, the pair annihilation is effective and the mass density of the LSP becomes insensitive to $`T_\mathrm{R}^{(\mathrm{inf})}`$. In particular, if $`m_\chi 100\mathrm{GeV}`$, $`\mathrm{\Omega }_\chi `$ is always $`O(1)`$ for high enough $`T_\mathrm{R}^{(\mathrm{inf})}`$, and the Wino-like LSP can be a candidate for CDM. If $`m_\chi 100\mathrm{GeV}`$, $`T_\mathrm{R}^{(\mathrm{inf})}`$ should be tuned to be $`10^{11}\mathrm{GeV}`$ to have $`\mathrm{\Omega }_\chi 1`$. Of course, if both the modulus field and the gravitino exist in the early Universe, primordial gravitinos can be diluted by the decay of the modulus field. In this case, the mass density of the LSP produced by the gravitino decay is negligible. The gravitino cosmology is also discussed in Ref. , but the authors neglected the effect of the pair annihilation of the LSP’s. As a result, they claimed that $`T_\mathrm{R}^{(\mathrm{inf})}10^{11}\mathrm{GeV}`$ is necessary to realize the Wino CDM. Based on Eq. (2.14), they also derived an upper bound on the reheating temperature of $`T_{\mathrm{R}}^{(\mathrm{inf})}{}_{}{}^{<}10^{11}\mathrm{GeV}`$ in order not to overclose the Universe. However, these arguments are modified as above once we include the effect of pair annihilation. Since the decay of the modulus field produces a large entropy, one may worry about the dilution of the baryon asymmetry of the Universe in our scenario. It is true that the dilution factor due to the modulus decay can be as large as $`10^{13}`$, and hence large primordial baryon asymmetry is required. However, this is not necessarily a problem. For example, the Affleck-Dine baryogenesis can provide enough baryon number asymmetry even with such a large dilution factor . In summary, the mass density of the LSP produced by the decay of the moduli fields can be sufficiently suppressed in the Wino-like LSP case, and the AMSB scenario provides an interesting solution to the cosmological moduli problem. Furthermore, in this case, the Wino-like LSP is a natural candidate for CDM. In the next section, we will see this is a very advantageous situation from the point of view of dark matter detection. ## 3 Detecting Wino CDM As we have seen, the Wino LSP is a promising candidate for cold dark matter. It comes out quite naturally from the parameters of the sequestered sector scenario. In this section, we consider the possibility of dark matter detection with Wino dark matter. We first discuss the CDM search with Ge detectors. If the LSP is the dominant component of the mass density of the halo, we may observe the energy deposit due to the LSP-nucleus scattering in a Ge detector. In the sequestered sector scenario, squark masses are calculable and are generally quite heavy. Therefore, the scattering processes mediated by the squark exchange are suppressed. However, the LSP $`\chi `$ also couples to the Higgs bosons as $`_{h\chi \chi }=y_{h\chi \chi }h\overline{\chi }\chi +y_{H\chi \chi }H\overline{\chi }\chi ,`$ (3.1) where $`h`$ and $`H`$ are the light and heavy CP even Higgses, respectively. Therefore, the LSP interacts with nuclei by exchanging the Higgs bosons. As we will see, this effect is important, and the detection rate in the Ge detector can be as large as $`0.10.01`$ event/kg/day which should be within the reach of the future detection of the CDM . The above Yukawa coupling constants, $`y_{h\chi \chi }`$ and $`y_{H\chi \chi }`$, arise from the mixings between the Wino and Higgsinos. These mixings are from the off-diagonal elements in the neutralino mass matrix, which is given by $`=\left(\begin{array}{cccc}m_{\mathrm{G1}}& 0& m_W\mathrm{cos}\beta \mathrm{tan}\theta _\mathrm{W}& m_W\mathrm{sin}\beta \mathrm{tan}\theta _\mathrm{W}\\ 0& m_{\mathrm{G2}}& m_W\mathrm{cos}\beta & m_W\mathrm{sin}\beta \\ m_W\mathrm{cos}\beta \mathrm{tan}\theta _\mathrm{W}& m_W\mathrm{cos}\beta & 0& \mu \\ m_W\mathrm{sin}\beta \mathrm{tan}\theta _\mathrm{W}& m_W\mathrm{sin}\beta & \mu & 0\end{array}\right),`$ (3.6) in the basis $`(i\stackrel{~}{B},i\stackrel{~}{W}^0,\stackrel{~}{H}_1^0,\stackrel{~}{H}_2^0)`$. Here, $`m_{\mathrm{G1}}`$ and $`m_{\mathrm{G2}}`$ are the gaugino masses for the gauginos associated with the U(1)<sub>Y</sub> and SU(2)<sub>L</sub> gauge groups respectively, $`\mu `$ is the supersymmetric Higgs mass, $`\mathrm{tan}\beta `$ is the ratio of Higgs VEVs, and $`\theta _\mathrm{W}`$ is the Weinberg angle. In the AMSB scenario, $`m_{\mathrm{G1}}`$ and $`m_{\mathrm{G2}}`$ are related by $`m_{\mathrm{G1}}3m_{\mathrm{G2}}`$, which we now assume. The above mass matrix can be diagonalized by using a unitary matrix, which we call $`U`$.<sup>#5</sup><sup>#5</sup>#5In our notation, the mass eigenstates of the neutralino are given by $`\chi _i=U_{1i}(i\stackrel{~}{B})+U_{2i}(i\stackrel{~}{W}^0)+U_{3i}\stackrel{~}{H}_1^0+U_{4i}\stackrel{~}{H}_2^0`$. With this unitary matrix, the mass of the LSP is given by $`m_\chi =|(U^{}U)_{11}|`$, for example. The Yukawa coupling constants are given by $`y_{h\chi \chi }`$ $`=`$ $`{\displaystyle \frac{1}{2}}g_2U_{21}(U_{31}\mathrm{cos}\beta U_{41}\mathrm{sin}\beta ),`$ (3.7) $`y_{H\chi \chi }`$ $`=`$ $`{\displaystyle \frac{1}{2}}g_2U_{21}(U_{31}\mathrm{sin}\beta +U_{41}\mathrm{cos}\beta ).`$ (3.8) Here, we neglect the difference between the neutral CP even Higgs mixing angle and the angle $`\beta `$. In the decoupling limit (i.e., $`m_hm_H`$), this difference is sufficiently small. When $`|\mu |`$ and $`|m_{\mathrm{G2}}|`$ are much larger than $`m_W`$, the LSP is mostly the Wino and $`U_{21}1`$. Other smaller elements are given by $`U_{11}0,U_{31}{\displaystyle \frac{m_W(m_{\mathrm{G2}}\mathrm{cos}\beta \mu \mathrm{sin}\beta )}{\mu ^2m_{\mathrm{G2}}^2}},U_{41}{\displaystyle \frac{m_W(m_{\mathrm{G2}}\mathrm{sin}\beta +\mu \mathrm{cos}\beta )}{\mu ^2m_{\mathrm{G2}}^2}},`$ (3.9) and in the same limit, $`y_{h\chi \chi }`$ and $`y_{H\chi \chi }`$ are given by $`y_{h\chi \chi }`$ $``$ $`{\displaystyle \frac{g_2m_W(m_{\mathrm{G2}}+\mu \mathrm{sin}2\beta )}{2(\mu ^2m_{G2}^2)}},`$ (3.10) $`y_{H\chi \chi }`$ $``$ $`{\displaystyle \frac{g_2m_W\mu \mathrm{cos}2\beta }{2(\mu ^2m_{G2}^2)}}.`$ (3.11) As one can see, the coupling of the light Higgs is sensitive to the relative sign of $`m_{\mathrm{G2}}`$ and $`\mu `$; $`y_{h\chi \chi }`$ is enhanced if $`\mu /m_{\mathrm{G2}}`$ is positive. Furthermore, as the ratio $`|\mu /m_{\mathrm{G2}}|`$ becomes larger, the Higgsino component of the LSP becomes smaller. As a result, the $`h\chi \chi `$ and $`H\chi \chi `$ interactions are suppressed. In the Bino LSP case, $`y_{h\chi \chi }`$ and $`y_{H\chi \chi }`$ are given by similar expressions with $`g_2m_W`$ being replaced by $`g_1m_W\mathrm{tan}\theta _\mathrm{W}`$. Therefore, the Wino LSP has stronger couplings to the Higgs boson than the Bino, which enhances the detection cross section. The interaction between the Higgs bosons and a nucleus are discussed in Ref. . Since we are interested in a scattering process with small recoil energy, the nucleus can be approximately regarded as an elementary particle with mass $`m_N`$. Furthermore, since $`m_N`$ is dynamically generated through QCD effects, $`m_N`$ is proportional to the QCD scale $`\mathrm{\Lambda }_{\mathrm{QCD}}`$. On the other hand, $`\mathrm{\Lambda }_{\mathrm{QCD}}`$ is sensitive to the fluctuation of the Higgs fields through the heavy quark mass dependence of $`\mathrm{\Lambda }_{\mathrm{QCD}}`$. By using these facts, we can relate the Higgs VEV dependence of the QCD scale to the Higgs-nucleus coupling constants. The interaction terms of the Higgs bosons with the nucleus $`N`$ are derived as $`_{\mathrm{int}}={\displaystyle \frac{_{\mathrm{mass}}}{m_N}}(y_{hNN}h+y_{HNN}H),`$ (3.12) where $`_{\mathrm{mass}}`$ is the mass term for the nucleus $`N`$ in the effective theory. We consider the case with Ge detector, and we will use the formula for $`N={}_{}{}^{76}\mathrm{Ge}`$. The “effective” coupling constants for the Higgs interactions are given by<sup>#6</sup><sup>#6</sup>#6We neglect the effects of the nuclear form factors. $`y_{hNN}`$ $`=`$ $`{\displaystyle \frac{2}{9}}{\displaystyle \frac{m_N}{v}},`$ (3.13) $`y_{HNN}`$ $`=`$ $`{\displaystyle \frac{2}{27}}{\displaystyle \frac{m_N}{v}}(\mathrm{tan}\beta 2\mathrm{cot}\beta ),`$ (3.14) with $`v246\mathrm{GeV}`$. As one can see in Eq. (3.14), the interaction of the heavy Higgs becomes stronger in the large $`\mathrm{tan}\beta `$ case, and the heavy Higgs exchange diagram may enhance the detection rate. With the above coupling constants, the total cross section for the scattering process $`\chi N\chi N`$ is given by $`\sigma _{\mathrm{scatt}}={\displaystyle \frac{4}{\pi }}\left({\displaystyle \frac{y_{hNN}y_{h\chi \chi }}{m_h^2}}+{\displaystyle \frac{y_{HNN}y_{H\chi \chi }}{m_H^2}}\right)^2{\displaystyle \frac{m_N^2m_\chi ^2}{(m_N+m_\chi )^2}},`$ (3.15) where $`m_h`$ and $`m_H`$ are the masses of $`h`$ and $`H`$, respectively. Assuming that the CDMs are virialized in the halo, we obtain the detection rate $`R={\displaystyle \frac{2}{\sqrt{3\pi }}}{\displaystyle \frac{\rho _\chi ^{(\mathrm{halo})}\overline{v}_\chi \sigma _{\mathrm{scatt}}}{m_Nm_\chi }}\mathrm{exp}\left({\displaystyle \frac{3(m_N+m_\chi )^2E_{\mathrm{thr}}}{4m_Nm_\chi ^2\overline{v}_\chi ^2}}\right),`$ (3.16) where $`\rho _\chi ^{(\mathrm{halo})}`$ is the mass density of the LSP in the halo, $`E_{\mathrm{thr}}`$ is the threshold energy of the detector, and $`\overline{v}_\chi `$ is the averaged velocity of the LSP in the halo. In Figs. 4 and 4, we show the expected detection rate of the Wino CDM in a <sup>76</sup>Ge detector for $`\mu =2m_{\mathrm{G2}}`$ and $`\mu =3m_{\mathrm{G2}}`$, respectively. In our calculation, we used the Higgs masses $`m_h=100\mathrm{GeV}`$ and $`m_H=300\mathrm{GeV}`$. The other parameters are taken to be $`E_{\mathrm{thr}}=2\mathrm{keV}`$, $`\overline{v}_\chi =320\mathrm{km}/\mathrm{sec}`$, and $`\rho _\chi ^{(\mathrm{halo})}=0.3\mathrm{GeV}/\mathrm{cm}^3`$. The behavior of Figs. 4 and 4 can be understood as follows. In the small $`\mathrm{tan}\beta `$ region, the scattering is dominated by the light Higgs exchange diagram. In this case, the detection rate $`R`$ decreases as $`\mathrm{tan}\beta `$ increase since $`y_{h\chi \chi }`$ is more suppressed for large $`\mathrm{tan}\beta `$ (if $`\mu /m_{\mathrm{G2}}>0`$). On the other hand, for the large $`\mathrm{tan}\beta `$ region, heavy Higgs exchange is the dominant contribution. Since the $`HNN`$ coupling $`y_{HNN}`$ is proportional to $`\mathrm{tan}\beta `$, $`R`$ is more enhanced for larger $`\mathrm{tan}\beta `$. Furthermore, $`R`$ decreases as $`m_\chi |m_{\mathrm{G2}}|`$ or $`|\mu |`$ increases, since the Wino-Higgsino mixing is more suppressed in this limit. We can see that the detection rate of order 0.1 $``$ 0.01 event/kg/day is possible in a Ge detector, which is within the reach of the on-going CDM searches . The detection rate is considerably larger than the conventional Bino CDM case. This is because the couplings $`y_{h\chi \chi }`$ and $`y_{H\chi \chi }`$ are (approximately) proportional to $`g_2^2`$ for the Wino LSP, instead of $`g_1^2`$ for the Bino LSP. Indeed, in the minimal supergravity model with the Bino LSP, the detection rate is typically $`O(10^3\mathrm{event}/\mathrm{kg}/\mathrm{day})`$ or less , which is an order of magnitude smaller than what is detectable.<sup>#7</sup><sup>#7</sup>#7In gravity-mediated SUSY breaking with the GUT relation among the gaugino mass parameters, a large Higgsino component in the LSP and/or a non-universal boundary condition for the scalar masses would be necessary for a detectable SUSY dark matter, unless $`\mathrm{tan}\beta `$ is large. It is very unlikely that conventional SUSY dark matter will be detected. Therefore, the Wino CDM has more chance to be detected in a Ge detector. Of course, the detection rate given in Eq. (3.16) is sensitive to the model parameters. For example, the discovery of the signal will be difficult if $`|\mu ||m_{\mathrm{G2}}|`$. Furthermore, in the large $`\mathrm{tan}\beta `$ region, $`Rm_H^4`$ and the detection rate decreases as $`m_H`$ increases. However, too large $`\mu `$ and too heavy Higgs are not preferred from the naturalness point of view. Therefore, in a significant fraction of the parameter space of the AMSB scenario, the Wino CDM should be detectable. For $`\mu /m_{\mathrm{G2}}<0`$, $`y_{h\chi \chi }`$ is suppressed because of the cancellation, as shown in Eq. (3.10). Therefore, in this case, the detection of the Wino CDM is difficult for small $`\mathrm{tan}\beta `$. On the other hand, for large $`\mathrm{tan}\beta `$, the detection rate is less sensitive to the relative sign between $`\mu `$ and $`m_{\mathrm{G2}}`$, and detection might be possible. There is an alternative means to search for dark matter, which is to look for the energetic neutrinos produced by the annihilation of the Wino LSP’s captured in the center of the sun and/or earth . Such a high energy neutrino can be converted to a high energy muon which can be observed as an upward going muon in Čerenkov detectors. Using the conversion factors given in Ref. , we find that an LSP with $`R0.10.01\mathrm{event}/\mathrm{kg}/\mathrm{day}`$ has the event rate for indirect detection in the Čerenkov detectors of $`\mathrm{\Gamma }_{\mathrm{ID}}101\mathrm{event}/\mathrm{m}^2/\mathrm{yr}`$. This is less sensitive than the direct detection in a Ge detector . It might be that the most promising method for searching for dark matter is to look for anti-matter, either anti-protons or positrons , produced by the pair annihilation of the LSP in our galaxy. The pair annihilation rate $`v_{\mathrm{rel}}\sigma `$ for the Wino LSP is given in Eq. (2.4), and numerically is given by $`3.8\times 10^{24}\mathrm{cm}^3/\mathrm{sec}`$ (for $`m_\chi =100\mathrm{GeV}`$) $``$ $`0.9\times 10^{24}\mathrm{cm}^3/\mathrm{sec}`$ (for $`m_\chi =300\mathrm{GeV}`$). Unlike standard SUSY dark matter, this rate is very large, and we expect a high flux of anti-particles. Furthermore, there are several on-going projects for measuring the anti-matter flux in the cosmic ray. In particular, a very accurate measurement is expected by the AMS experiment , which is a search for anti-matters with “Alpha Magnetic Spectrometer” on the space shuttle and on the international space station. Since the experiment is not affected by the atmosphere, AMS will greatly improve the measurements of the anti-matter fluxes in the cosmic ray. A recent calculation of the anti-proton flux $`\mathrm{\Phi }_{\overline{p}}`$ can be found in Ref. where the flux from an LSP which dominantly annihilates into $`W`$-boson pair is presented (see Example No. 4 in Ref. ). Since $`\mathrm{\Phi }_{\overline{p}}`$ from the LSP annihilation is proportional to $`v_{\mathrm{rel}}\sigma \times (\rho _\chi ^{(\mathrm{halo})}/m_\chi )^2`$, we estimate the anti-proton flux by rescaling the result given in Ref. . Adopting the canonical astrophysical parameters of Ref. , we find $`\mathrm{\Phi }_{\overline{p}}`$ to be $`0.46\mathrm{m}^2\mathrm{sec}^1\mathrm{sr}^1\mathrm{GeV}^1`$ (for $`m_\chi =100\mathrm{GeV}`$) $``$ $`0.012\mathrm{m}^2\mathrm{sec}^1\mathrm{sr}^1\mathrm{GeV}^1`$ (for $`m_\chi =300\mathrm{GeV}`$). Experimentally, the anti-proton flux is already measured by the BESS experiment, and is given by $`(1.36_{0.61}^{+0.86})\times 10^2\mathrm{m}^2\mathrm{sec}^1\mathrm{sr}^1\mathrm{GeV}^1`$ . As a result, a naive comparison of our estimate with the measured flux would already imply the constraint $`m_{\chi }^{}{}_{}{}^{>}250\mathrm{GeV}`$. One should note, however, that the flux is proportional to $`\rho _{\chi }^{(\mathrm{halo})}{}_{}{}^{2}`$. Furthermore, if we change the model of the halo, it affects the propagation of the anti-proton and the flux can be reduced significantly . Therefore, the theoretical result is very sensitive to astrophysics parameters, and we do not draw any strong conclusion here. It is unlikely that the situation for anti-protons will improve with the AMS experiment, since the anti-proton signal falls with higher energy faster than the background. However, the situation for positrons is very promising as we now discuss. In Ref. , the positron flux, $`\mathrm{\Phi }_{e^+}`$, is also presented. Rescaling the given result, we found that $`\mathrm{\Phi }_{e^+}`$ from the Wino-like CDM is comparable to the currently measured flux by the HEAT experiment even for $`m_\chi 100\mathrm{GeV}`$. Therefore, we believe our scenario is not seriously constrained by the present data, but could give a very prominent signal in the future. One important point is that the positron spectrum from the Wino CDM is peaked at high energy ($`\frac{1}{2}m_\chi `$, which can be a distinctive signature of the Wino annihilation in the halo. In order to study the significance of this signal, we rescaled the results given in Ref. to obtain the positron fraction $`\mathrm{\Phi }_{e^+}/(\mathrm{\Phi }_e^{}+\mathrm{\Phi }_{e^+})`$ at the peak.<sup>#8</sup><sup>#8</sup>#8In Ref. , positron fraction is calculated for the Higgsino-like CDM which also dominantly annihilates into $`W`$\- and $`Z`$-bosons. We estimated the positron fraction for the Wino-like CDM by rescaling with Eq. (2.4). We used the halo mass density to be $`\rho _\chi ^{(\mathrm{halo})}=0.3\mathrm{GeV}/\mathrm{cm}^3`$. The results are given in Table 1. As one can see, with the Wino CDM, the positron fraction can be about 5 $``$ 2 times larger than the background for $`m_\chi =100300\mathrm{GeV}`$. A measurement of the positron fraction has recently performed at lower energies by AMS who will also do very precise measurement at higher energies in the future. Since the signal can be peaked at higher energy, this should be an excellent way to search for Wino CDM.<sup>#9</sup><sup>#9</sup>#9In fact, Ref. suggests a distortion in the positron spectrum measured by the HEAT experiment . Positrons from the Wino CDM may be the source of this distortion. The signal is much stronger for Wino CDM than the more frequently studied Bino dark matter, which would be undetectable. The distortion suggested here should provide strong evidence for the Wino CDM scenario. ## 4 Conclusion In this paper, we discussed the cosmological moduli problem in the sequestered sector/AMSB scenario. In this scheme, the gravitino mass (corresponding to the moduli mass) is naturally 10 $``$ 100 TeV. As a result, cosmological moduli fields can decay before BBN starts. Furthermore, since the LSP is likely to be the Wino-like neutralino, the production of the LSP through moduli decay can be suppressed and the Universe should not be overclosed, contrary to the case of the Bino-like LSP. Moreover, the mass density of the Wino-like LSP can be naturally close to the critical density of the Universe, and hence the Wino is an interesting candidate for CDM. Furthermore, we have seen that if the halo density is dominated by the Wino CDM, the detection rate of the Wino CDM in Ge detector can be as large as $`0.10.01`$ event/kg/day, which is within the reach of the CDM search experiment, and furthermore, the positron signal in AMS should be measurable. ## Acknowledgements We would like to thank Ann Nelson, Paul Schechter, Yael Shadmi, and Frank Wilczek for useful discussions. This work was supported in part by the National Science Foundation under contract NSF-PHY-9513835, and in part by the Department of Energy under contracts DE-FG02-90ER40542, DE-FG-02-91ER40671, and cooperative agreement DF-FC02-94ER40818. The work of TM was also supported by the Marvin L. Goldberger Membership. ## Appendix A Moduli Field Couplings In this Appendix, we discuss the mass and couplings of the moduli fields. We assume that a modulus field $`\varphi `$ acquires its mass from SUSY breaking effects. Therefore, its mass is expected to be of the order of the gravitino mass $`m_{3/2}10100\mathrm{TeV}`$ . Some non-perturbative effects may be able to give much larger masses to the moduli fields. Such a heavy modulus field is cosmologically safe since its lifetime can be much shorter than 1 sec. It is also important to understand the relevant operators which contribute to the decay of $`\varphi `$. By taking the minimal SUSY standard model (MSSM) as the low energy effective theory, the following operators can exist. The modulus field can decay into a gauge field (and gaugino) through the operators: $`_\mathrm{G}={\displaystyle d^2\theta \frac{\lambda _\mathrm{G}}{M_{}}\varphi W^\alpha W_\alpha }+\text{h.c.},`$ (A.1) where $`\lambda _\mathrm{G}`$ is a constant to parameterize the strength of this interaction. When the gaugino mass $`m_{\stackrel{~}{G}}`$ is much smaller than $`m_\varphi `$, the branching ratio for the decay into the gaugino pair receives chirality suppression . As a result, the (partial) decay width and $`\overline{N}_{\mathrm{LSP}}`$ is given by $`\mathrm{\Gamma }_\mathrm{G}`$ $`=`$ $`{\displaystyle \frac{N_\mathrm{f}\lambda _\mathrm{G}^2}{8\pi }}{\displaystyle \frac{m_\varphi ^3}{M_{}^2}},`$ (A.2) $`\overline{N}_{\mathrm{LSP}}`$ $``$ $`O(m_{\stackrel{~}{G}}^2/m_\varphi ^2),`$ (A.3) where $`N_\mathrm{f}`$ is the number of the possible final states. (For example, $`N_\mathrm{f}=N^21`$ for an SU($`N`$) gauge group.) One should note that this operator also contributes to the gaugino mass if the $`\varphi `$ field participates in SUSY breaking. So either $`\lambda _G`$ is suppressed or $`F_\varphi `$ is in order to maintain the anomaly-mediated predictions. In the MSSM, the following operator is also allowed: $`_\mathrm{H}={\displaystyle d^4\theta \frac{\lambda _\mathrm{H}}{M_{}}\varphi H_1^{}H_2^{}}+\text{h.c.},`$ (A.4) where $`H_1`$ and $`H_2`$ are the down-type and up-type Higgses, respectively. With this operator, the modulus field can decay into a Higgs boson pair, and the decay width for this process is $`\mathrm{\Gamma }_\mathrm{H}={\displaystyle \frac{\lambda _\mathrm{H}^2}{8\pi }}{\displaystyle \frac{m_\varphi ^3}{M_{}^2}}.`$ (A.5) Notice that the Higgsino cannot be produced from this operator, and hence $`\overline{N}_{\mathrm{LSP}}=0`$ for this process. This operator is also dangerous if $`F_\varphi `$ is maximal since it generates too large $`\mu `$-parameter. One can also write down the following operator: $`_{\mathrm{Q}^{}\mathrm{Q}}={\displaystyle d^4\theta \frac{\lambda _{\mathrm{Q}^{}\mathrm{Q}}}{M_{}}\varphi Q^{}Q}+\text{h.c.}={\displaystyle \frac{\lambda _{\mathrm{Q}^{}\mathrm{Q}}}{M_{}}}\varphi (^2\stackrel{~}{Q}^{})\stackrel{~}{Q}+\mathrm{},`$ (A.6) where $`Q`$ is general chiral superfields in the MSSM and $`\stackrel{~}{Q}`$ is its scalar component. As suggested from the structure of the operator, the decay rate is given by $`\mathrm{\Gamma }_{Q^{}Q}O\left({\displaystyle \frac{\lambda _{\mathrm{Q}^{}\mathrm{Q}}^2}{4\pi }}{\displaystyle \frac{m_{\stackrel{~}{Q}}^4}{m_\varphi ^4}}{\displaystyle \frac{m_\varphi ^3}{M_{}^2}}\right),`$ (A.7) where $`m_{\stackrel{~}{Q}}`$ is the soft SUSY breaking mass for $`\stackrel{~}{Q}`$. Therefore, the decay through this operator is suppressed by a factor of $`O(m_{\stackrel{~}{Q}}^4/m_\varphi ^4)`$. In supergravity, we also expect the following interaction: $`_\psi ={\displaystyle \frac{1}{M_{}^2}}e^{K/2}W\psi _\mu \sigma ^{\mu \nu }\psi _\nu +\text{h.c.},`$ (A.8) where $`K`$ and $`W`$ are the Kähler potential and superpotential, and $`\psi _\mu `$ is the gravitino field, respectively. If the decay $`\varphi \psi _\mu \psi _\mu `$ is kinematically allowed, the effect of this operator can be important. In this case, $`\varphi `$ decays into a gravitino pair; then the produced gravitino decays into the standard model particle and its superpartner through the supercurrent interaction. For this process, the effective decay width (i.e., the inverse of the time scale of this process) and $`\overline{N}_{\mathrm{LSP}}`$ are estimated as $`\mathrm{\Gamma }_\psi `$ $``$ $`O\left({\displaystyle \frac{1}{4\pi }}{\displaystyle \frac{m_{3/2}^3}{M_{}^2}}\right),`$ (A.9) $`\overline{N}_{\mathrm{LSP}}`$ $``$ $`O(1).`$ (A.10) There are also several operators that permit three-body decay. For example, a modulus field can couple to the Yukawa-type operator in the $`F`$-term like<sup>#10</sup><sup>#10</sup>#10We assume the coupling of the modulus field to the light quarks in $`F`$-term is more suppressed. Otherwise, large vacuum expectation value of $`\varphi M_{}`$ induces too large Yukawa coupling constants for light quarks. $`_\mathrm{Y}={\displaystyle d^2\theta \frac{\lambda _\mathrm{Y}}{M_{}}\varphi H_2t_\mathrm{R}^\mathrm{c}q_\mathrm{L}}+\text{h.c.},`$ (A.11) where $`H_2`$, $`t_\mathrm{R}^\mathrm{c}`$, and $`q_\mathrm{L}`$ are chiral superfields for up-type Higgs, right-handed top quark, and left-handed third generation quark doublet, respectively. With this operator, the decay rate is calculated as $`\mathrm{\Gamma }_\mathrm{Y}={\displaystyle \frac{3\lambda _\mathrm{Y}^2}{128\pi ^3}}{\displaystyle \frac{m_\varphi ^3}{M_{}^2}}.`$ (A.12) This is a three-body process, and hence the decay rate is more suppressed than those of the two-body processes. For this process, $`\overline{N}_{\mathrm{LSP}}O(1)`$. One may also write down the scalar interaction of the form $`\frac{m_{3/2}}{M_{}}\varphi H_2\stackrel{~}{t}_\mathrm{R}^\mathrm{c}\stackrel{~}{q}_\mathrm{L}`$ arising from the supergravity effect. If the coefficient of this operator is $`O(1)`$, however, too large an $`A`$-parameter is generated with $`\varphi M_{}`$. Therefore, we assume this operator is somehow suppressed if $`\varphi M_{}`$. So, to summarize, the moduli fields are expected to have a decay rate as large as that estimated on dimensional grounds. The operators that could lead to this decay rate are the operator given in Eq. (A.8) which leads to the decay to gravitinos (if kinematically permitted), the coupling to gauge pairs (A.1), and the coupling to Higgses (A.4). However, a large value for the last two couplings require a small value for $`F_\varphi `$. The decay rate with no suppression factors is consistent with the requirements for avoiding the cosmological moduli problem. The density of Wino dark matter depends on $`\overline{N}_{\mathrm{LSP}}`$. In general, $`\overline{N}_{\mathrm{LSP}}`$ is expected to be small. The precise value depends on the magnitude of the coupling of the associated operators, but values of order $`10^410^2`$ are expected. The exception to this small value is the operator that permits decay to gravitinos. If kinematically possible, this would permit large $`\overline{N}_{\mathrm{LSP}}1`$.
no-problem/9906/astro-ph9906238.html
ar5iv
text
# Massive Multi-species, Multi-level NLTE Model Atmospheres for Novae in Outburst ## 1 Introduction Novae in the optically thick wind phase of their outburst pose a special challenge for atmospheric modeling because of their steep temperature gradients, low densities, large geometric extents, and high velocity differential flows. For example, a model with a bolometrically defined effective temperature, $`T_{\mathrm{eff}}`$, of 15 000 K will have values of the kinetic temperature, $`T_{\mathrm{kin}}`$, that range from 4500 to 140 000 K, of the mass density, $`\rho `$, that range from $`3\times 10^{15}`$ to $`6\times 10^9`$ g cm<sup>-3</sup>, a geometric extent, $`R_{\mathrm{out}}/R_{\mathrm{in}}`$, equal to 120 (this paper), and a typical maximum expansion velocity, $`V_0`$, of $`2000`$ km s<sup>-1</sup>. Because of the low values of $`\rho `$, radiative rates exceed collisional rates for many important transitions throughout the atmosphere. The steep $`T_{\mathrm{kin}}`$ gradient and low densities allow local regions to be exposed through scattering processes to radiation from distant regions in the atmosphere where the radiation temperature, $`T_{\mathrm{rad}}`$, differs greatly from the local value of $`T_{\mathrm{kin}}`$. Moreover, massive ultraviolet ($`UV`$) and optical line blanketing cause the radiation field to deviate greatly from a Planck ($`B_\nu `$) distribution. As a result of these considerations, non-local thermodynamic equilibrium (NLTE) effects are severe in these atmospheres and must be accounted for to construct accurate models. The low values of $`\rho `$ cause the atmosphere to be translucent ($`\tau <1`$) at large physical depths, which, combined with the large absolute value of $`_\tau T`$, conspires to allow the line spectrum of many ionization stages of an atomic element to be present in the emergent flux spectrum. Moreover, the rapid expansion velocities and steep $`v(r)`$ gradients cause this rich line spectrum to be smeared, which complicates the radiative transfer by increasing the amount of overlap among different transitions. Also, the large velocities cause special relativistic effects to be marginally significant in the transfer of radiation, at least to first order in $`v/c`$. Finally, the large geometric extent allows sphericity effects to be important. ##### Hauschildt et al. (1995) and Hauschildt et al. 1997b (henceforth HSSBSA) investigated the effects of NLTE on atmospheric models of novae in the optically thick wind phase of their outburst. They concluded that an accurate treatment of NLTE effects for many transitions is critical for correctly calculating both the structure of the atmosphere and the emergent flux spectrum. Among their main results are that: 1) NLTE effects change the predicted concentration of minority ionization stages of elements that contribute significantly to the line opacity, such as Fe and CNO, by as much as three orders of magnitude at depths where the continuum optical depth at $`5000`$Å, $`\tau _{\mathrm{cont},5000}`$, is less than $`10^2`$, 2) NLTE effects brighten the predicted $`UV`$ pseudo-continuum by as much as an order of magnitude in the case of their $`\mathrm{25\hspace{0.17em}000}`$ K model in the range $`\lambda <1000`$Å, and 3) a minimal incorporation of NLTE physics by including a large single scattering albedo in the line source functions of an LTE synthetic spectrum gives rise to a spectrum with approximately the same overall level of line blanketing absorption as a more complete NLTE treatment, although any individual line profile may be inaccurate. HSSBSA concluded that their NLTE models provided a better fit to observed spectra of novae. ##### One of the limitations of the modeling of Hauschildt et al. (1995) and HSSBSA is the limited number of elements and ionization stages treated in NLTE (15 elements and 36 stages). Given the importance of NLTE effects, we have increased the NLTE treatment to include 19 elements and 87 ionization stages. In Section 2 we describe the NLTE treatment, in Section 3 we describe the nova models, in Section 4 we present the results of our calculations, and in Section 5 we present our conclusions. ## 2 NLTE Treatment Table 1 shows the complete set of chemical elements and ionization stages now treated in NLTE by PHOENIX along with the number of levels and primary bound-bound transitions included in the model atoms and ions. Primary transitions are those that connect states with observed energy levels and that have a $`\mathrm{log}gf`$ value greater than $`3.0`$. They are explicitly treated in the NLTE rate equations. All other transitions of a species treated in NLTE are considered secondary. Secondary transitions are not included in the rate equations, but are still included in the line opacity, with the departure co-efficients of any levels not included in the rate equations set equal to that of the ground state (see Hauschildt and Baron (1995) for a detailed description). New species are denoted in bold face type and species for which the treatment has been improved with enlarged model atoms are denoted with italics. We have added four new elements and 51 new ionization stages to the NLTE treatment with the result that the code can now treat a total of 87 ionization stages among 19 elements. This has increased the number of levels and lines included in the statistical equilibrium solution by 5415 and 42521, respectively, thereby approximately doubling the total numbers included. Aluminum, P, K, and Ni are now included in the NLTE treatment for the first time. The latter is important for supernova modeling because of the role of the radio-active decay of Ni in the energy balance of SNe envelopes, but is inconsequential for the nova modeling described here. Most elements now have at least the lowest six ionization stages treated in NLTE. For most elements this level of ionization corresponds to an energy in the range of 100 to 200 eV, which is well above the local thermal energy in the atmospheres of the hottest novae that we have modeled. In addition, we have improved the NLTE treatment of Na I and Mg II by enlarging the model atoms from three levels and two lines to 53 levels and 142 lines for Na I, and from 18 levels and 37 lines to 273 levels and 835 lines for Mg II. ##### We have also enlarged the H I model atom from 30 levels and 435 lines to 80 levels and 3160 lines. However, we restrict the models to 30 levels of H I because our treatment of dissolution effects among high lying levels is only approximate. These dissolution effects have been shown by Hubeny et al. (1994) to have an important effect on model structure and synthetic spectra for models of main sequence stars in the temperature range being investigated here. We note that because Novae atmospheres are expanding rapidly, the gas pressure in the line forming region is orders of magnitude less than that of a static White Dwarf (WD) star. Therefore, we expect dissolution effects to be negligible for states with principal quantum number, $`n`$, less than 30. ##### Atomic data for the energy levels and $`bb`$ transitions have been taken from Kurucz (Kurucz (1994), Kurucz and Bell (1995)). An accurate treatment of photo-ionization is important for the correct solution of the opacity and chemical equilibrium of a NLTE gas. We have used the resonance averaged Opacity Project (Seaton et al. (1994)) data of Bautista et al. (1998) for the ground state photo-ionization cross sections for Li I- II, C I- IV, N I- VI, O I- VI, Ne I, Na I- VI, I- VI, Al I- VI, Si I- VI, S I- VI, Ca I- VII, and Fe I- VI. These data also incorporate $`X`$-ray band opacity due to ionizations from the $`K`$ electron shell. For those species that were already included in the PHOENIX NLTE treatment, the use of this $`bf`$ opacity data is an improvement. For the ground states of all stages of P, Ti, Co, and Ni, and for the excited states of all species, we have used the cross sectional data previously incorporated into PHOENIX, which are those of Reilman & Manson (Reilman and Manson (1979)) or those compiled by Mathisen (Mathisen (1984)). We account for coupling among all bound levels by electronic collisions using cross-sections calculated with the formula of Allen (1973), except for those levels connected by radiative transitions, for which we use the formula of Van Regemorter (1962). The cross sections of ionizing collisions with electrons are calculated with the formula of Drawin (1961). ##### With the expanded NLTE treatment, PHOENIX now treats simultaneously $`10000`$ levels and $`\mathrm{100\hspace{0.17em}000}`$ primary transitions in NLTE. This massive NLTE treatment is made possible by the efficiency of the operator splitting method for solving the multi-level NLTE rate equations (Hauschildt (1993)) and of the ALI/OS method for solving the radiative transfer equation (Hauschildt (1992)), and by the parallel implementation of PHOENIX (Hauschildt et al. 1997a , Baron and Hauschildt (1998)). ## 3 Models We have calculated three models, designated $`M1`$, $`M2`$, and $`M3`$, that represent the state of a nova at three different times during the optically thick wind phase of the outburst. Table 2 presents some of the model parameters. During the optically thick wind phase, the radius at which $`\tau _{\mathrm{cont},5000}`$ equals unity is decreasing due to the decrease in $`\rho `$ as the atmosphere expands. The decreasing value of $`\rho `$ also allows energetic photons from the central engine to penetrate further out into the expanding atmosphere, thereby increasing $`T_{\mathrm{eff}}`$. For these spherically extended atmospheres the value of $`T_{\mathrm{eff}}`$ is defined as the model temperature that corresponds to the total frequency integrated luminosity ($`\sigma T_{\mathrm{eff}}^4=\frac{3}{4\pi R_{\tau =1}^2}_0^{\mathrm{}}L_\nu 𝑑\nu `$). During this phase, the photospheric radius, $`r(\tau _{\mathrm{cont},5000}=1)`$, and $`T_{\mathrm{eff}}`$ change such that the bolometric luminosity, $`L_{\mathrm{bol}}`$, remains constant. For our models, $`L_{\mathrm{bol}}`$ is equal to $`\mathrm{50\hspace{0.17em}000}L_{}`$. The optical depth grid contains 50 points and spans the range from a $`\mathrm{log}\tau _{\mathrm{cont},5000}`$ value of -10 to three. Because of the decline in $`\rho `$, a fixed grid in $`\tau `$ space corresponds to a contracting grid in physical distance space. However, note that the inner radius at the bottom of the model, $`R_{\mathrm{in}}`$, decreases faster than the outer radius at the top of the model, $`R_{\mathrm{out}}`$, so that the geometric extent $`R_{\mathrm{out}}/R_{\mathrm{in}}`$, and, therefore, the sphericity, of the model is increasing with time. The density law, $`\rho (r)=\rho _0(r/R_0)^n`$, is prescribed and has a value of $`n=3`$. The velocity law, $`v(r)=v_0(r/R_0)`$, corresponds to a constant mass loss rate ($`\dot{M}(r)=`$constant), and has a value of $`v_0`$ equal to $`2000`$ km s<sup>-1</sup>. ##### For each of these models we have calculated a converged solution for the atmospheric structure that is subject to the constraint of radiative equilibrium for three cases: 1) quasi-LTE treatment for all species in which the Boltzmann and Saha distributions are used for the level populations, and all lines are assumed to have an albedo for single scattering of 0.95, 2) NLTE treatment for only the HSSBSA species, which are denoted by normal typeface in Table 1, and 3) NLTE treatment of all relevant species presented in Table 1. The line albedo used for case 1 was found by HSSBSA to be necessary to produce synthetic LTE spectra for nova models in which the line strengths were approximately realistic. For case 3, relevant species are those whose levels have a numerically significant population anywhere in the atmosphere and whose line spectrum is already known to be an important opacity source for determining the atmospheric structure or the appearance of the emergent spectrum. This limit was placed on the number of species treated in NLTE for the sake of computational expediency. Table 3 lists those species treated in NLTE in each case for each model. Note that for no model are all the species in Table 1 treated in NLTE. In this regard we note that some of the species in Table 1 are not important for Nova modeling, and the facility to treat them in NLTE was added to PHOENIX for the sake of other applications. For cases 2 and 3, the NLTE problem is converged self-consistently with the atmospheric structure. ## 4 Results ### 4.1 NLTE populations Figs. 1 to 4 show a comparison of the partial pressures of various species computed in LTE and with both NLTE treatments for all model. The current NLTE treatment includes all those species indicated in Table 3. Note that changes to the NLTE treatment may affect the concentration of any particular species in three different ways: 1) through the effect of all NLTE treated species on the chemical equilibrium of the gas by way of the contribution of each species to the $`e^{}`$ reservoir, 2) through changes in the equilibrium structure of the atmosphere as a result of NLTE effects on the total opacity and on the $`e^{}`$ density (see Fig. 5), and 3) through changes in the radiation field in a transition of one species that overlaps an important transition in another species. Examples of the latter include line interlocking and the pumping of transitions, and are especially important for species that have a rich line spectrum such as Fe, Co, and Ni. #### 4.1.1 Hydrogen, Helium, and electrons The upper left panels of Fig. 1 is directly comparable to Fig. 1 of HSSBSA. We confirm the result of HSSBSA that for a model of $`T_{\mathrm{eff}}`$ equal to $`\mathrm{15\hspace{0.17em}000}`$ K, NLTE enhances the H I concentration above the LTE values in the outer atmosphere where $`\mathrm{log}\tau _{\mathrm{cont},5000}<6`$ and slightly reduces it in the range where $`6<\mathrm{log}\tau _{\mathrm{cont},5000}<2`$. The largest NLTE effect is the reduction of the H I concentration by as much as 1.5 dex deep in the atmosphere where $`1<\mathrm{log}\tau _{\mathrm{cont},5000}<2.5`$. We note that the H II concentration is very close to the $`e^{}`$ concentration throughout the atmosphere, which indicates that H ionization is the dominant $`e^{}`$ contributor, even in the $`\mathrm{log}\tau _{\mathrm{cont},5000}<4`$ range where H is mostly neutral. As a result, the NLTE reduction of the $`e^{}`$ concentration in the $`\mathrm{log}\tau _{\mathrm{cont},5000}<6`$ range and enhancement of it in the $`6<\mathrm{log}\tau _{\mathrm{cont},5000}<2`$ range is mostly due to NLTE effects on the H ionization equilibrium. We also confirm the HSSBSA result that NLTE slightly reduces the He I concentration around $`\mathrm{log}\tau _{\mathrm{cont},5000}=4`$. We also find that NLTE reduces the He II concentration for $`\mathrm{log}\tau _{\mathrm{cont},5000}<1`$. However, the He II concentration is declining rapidly in this range as $`\tau `$ decreases due to recombination to He I as $`T(\tau )`$ decreases. This NLTE effect is not apparent in Fig. 1 of HSSBSA because of the more limited scale of their figure. The inclusion of the new species in NLTE has a negligible effect on the concentration of H I and II, and He I for this model. The new treatment produces a slight effect on the $`e^{}`$ concentration, first reducing it, then enhancing it as $`\mathrm{log}\tau _{\mathrm{cont},5000}`$ decreases below $`5`$. We note that changes in the state of the gas at $`\mathrm{log}\tau _{\mathrm{cont},5000}<6`$ only affect the profiles of the strongest spectral lines. The He II concentration at $`\mathrm{log}\tau _{\mathrm{cont},5000}<1`$ is reduced in the new treatment by approximately as much as the HSSBSA treatment reduced it from the LTE values. ##### The upper left panels of Fig. 2 shows the same species for the $`M2`$ model. Here, the effects of NLTE, and of the different NLTE treatments are larger. For $`\mathrm{log}\tau _{\mathrm{cont},5000}<4`$ where H I becomes relatively abundant, the HSSBSA treatment reduces the H I concentration by as much as two and a half orders of magnitude. However, the current NLTE treatment gives rise to a smaller reduction, producing results that are closer to the original LTE concentration. The new treatment also predicts a small enhancement in the concentration around $`\mathrm{log}\tau _{\mathrm{cont},5000}=5`$ that is not present in the HSSBSA results. As with the $`M1`$ model, the largest NLTE effect is reduction in H I concentration by as much as $`1.5`$ dex at greater depth where $`3<\mathrm{log}\tau _{\mathrm{cont},5000}<2.5`$. Both NLTE treatments give rise to very similar He concentrations and produce a reduction in the He I and II pressures for $`\mathrm{log}\tau _{\mathrm{cont},5000}<1.5`$. This reduction is as much as half an order of magnitude in the case of He I. Unlike the $`M1`$ model, the $`e^{}`$ pressure throughout the outer atmosphere is not driven by either the H or He ionization equilibria. The two regions of reduction in the $`e^{}`$ concentration at $`5<\mathrm{log}\tau _{\mathrm{cont},5000}<1`$ and $`9<\mathrm{log}\tau _{\mathrm{cont},5000}<6`$ are matched by similar reductions in He I concentration and the outermost of these reductions is also matched by a reduction in the H I concentration. A NLTE reduction in the population of electrons available for recombinations is driving a reduction in the neutral H and He populations. ##### Model $`M3`$ is hotter than any of the models discussed in HSSBSA. From the upper left panel of Fig. 4 we see that at $`T_{\mathrm{eff}}=\mathrm{35\hspace{0.17em}000}`$ K H is almost completely ionized throughout the entire atmosphere. Neutral H is a very small minority stage and is very sensitive to the treatment of the equilibrium. The NLTE suppression of H I at depth that was noted for the $`M1`$ and $`M2`$ models is more pronounced here: the NLTE H I concentration drops below its LTE value by as much as two orders of magnitude from $`5<\mathrm{log}\tau _{\mathrm{cont},5000}<2.5`$. The $`e^{}`$ and H<sup>+</sup> populations, which dominate the gas pressure throughout the entire atmosphere, show slight NLTE effects around $`4<\mathrm{log}\tau _{\mathrm{cont},5000}<2`$ where there is a slight NLTE reduction by as much as $`0.3`$ dex. #### 4.1.2 Metals Carbon, nitrogen, and oxygen (CNO) are important in nova spectra as a measure of convective dredge up of nuclear processed material in the progenitor WD, and as a test of thermonuclear runaway (TNR) models of nova explosions (Starrfield et al. (1998), Gehrz et al. (1998)). The upper right panel of Fig. 2 is directly comparable to Fig. 2 of HSSBSA. We confirm their result that C II is the dominant ionization stage, and that the HSSBSA NLTE treatment reduces the C I and enhances the C III concentration compared to the LTE levels for $`\mathrm{log}\tau _{\mathrm{cont},5000}<2`$. The current NLTE treatment produces approximately the same results for C I and C II, but produces a slight NLTE reduction in C III throughout most of the range where the HSSBSA treatment produces an enhancement. From examination of Fig. 2 and of Fig. 3 in HSSBSA we see that the same relative behavior for the different NLTE treatments is seen for the lowest three stages of N. From inspection of Fig. 1 and Fig. 4 in HSSBSA we see that NLTE effects for O are much smaller in the cooler $`M1`$ model. We note that the details of the NLTE treatment of CNO are identical in for both HSSBSA and the present calculation; the number of ionization stages and the size of the model atoms are the same in both treatments. The large differences seen in the concentration of some CNO stages in the $`M2`$ model are entirely due to the indirect effects of other species treated in NLTE by way of either the chemical equilibrium or the atmospheric structure. ##### The rich blanketing of line opacity contributed by Fe plays a crucial role in the time development of the $`UV`$ and optical light curves and spectra during the outburst (Hauschildt et al. (1994), Shore et al. (1994)). From inspection of the lower right panel in Figs. 1 and 2 we see that Fe II is the dominant ionization stage throughout most of the outer atmosphere in the $`M1`$ model, whereas Fe III plays the same role in the $`M2`$ and $`M3`$ models. For the $`M1`$ model, the HSSBSA NLTE treatment leads to a significant enhancement of the Fe I concentration above LTE values for $`\mathrm{log}\tau _{\mathrm{cont},5000}<2`$, whereas the current treatment leads to a reduction in this range and to a concentration that is much closer to LTE. Both treatments lead to significant enhancements in the Fe III concentration throughout the outer atmosphere. For the $`M2`$ and $`M3`$ models, NLTE effects in both treatments lead to a large reduction in both the Fe I and Fe II concentrations for $`\mathrm{log}\tau _{\mathrm{cont},5000}<2`$. However, the current NLTE treatment gives rise to an Fe I reduction that is approximately twice as large as the that of the HSSBSA treatment. We note that Fe II in particular has a very rich $`UV`$ line spectrum that determines the $`UV`$ flux distribution and affects the atmospheric structure during the optically thick wind phase of the outburst (Hauschildt et al. (1994), Shore et al. (1994)). Therefore, the reduction in Fe II concentration by two orders of magnitude that is produced by both NLTE treatments in the $`M2`$ and $`M3`$ models will have a significant effect on the model structure. This is also discussed by HSSBSA. As was the case with CNO, differences between the two NLTE treatments may be due to the indirect effects of other NLTE species. In addition, the difference in the Fe I- III results may also be due to the presence of a detailed Fe IV model atom in the NLTE calculation. ##### In Figs. 2 and 3 we show concentrations for a variety of other species treated in NLTE for the $`M2`$ model. NLTE results for S and Si were shown in HSSBSA, but they only treated the second and third ionization stages in NLTE, whereas we treat the lowest six stages. These two elements provide an object lesson in the importance of adjacent ionization stages to the accurate treatment of a particular stage in NLTE. In their Fig. 5 HSSBSA show an increase in the concentration of S I with respect to LTE by as much as three orders of magnitude for $`\mathrm{log}\tau <3`$ when S II and III are treated in NLTE. Our results for the $`M2`$ model, shown in the bottom right panel of Fig. 3, are not exactly comparable because $`M2`$ has a $`T_{\mathrm{eff}}`$ that is $`5000`$ K cooler than the model for which HSSBSA show results. However, our treatment, in which S I is included in NLTE, produces a seven orders of magnitude decrease in the S I concentration in the outer atmosphere. Moreover, whereas, HSSBSA found an increase by $`1`$ order of magnitude in the S III concentration in the same $`\tau `$ range, our calculations, in which S IV is included in NLTE, show an increase by as much as nine orders of magnitude in the outer atmosphere. Comparison of Fig. 6 in HSSBSA to the bottom left panel of Fig. 2 shows similar effects for Si. For example, whereas the HSSBSA treatment, in which Si I is in LTE, yields an increase with respect to LTE in the Si I concentration in the outer atmosphere of a $`\mathrm{25\hspace{0.17em}000}`$ K model, we find that the $`M2`$ model with Si I to IV in NLTE shows a decrease by as much as five orders of magnitude. We note that S I and Si I are both minority stages in these models. In general, the population of a minority stage is very sensitive to the photoionization rate and to the population of the reservoir stage. ##### Aluminum and the higher stages of Na are treated in NLTE for the first time. Magnesium and Ca both have $`UV`$ resonance lines from the second ionization stage that are important spectral features. HSSBSA treated only the second stage in NLTE, whereas we treat the lowest six. Inspection of these figures shows that for all of these species, NLTE effects, and the particular treatment of NLTE, are significant for one or more stages. ### 4.2 Atmospheric structure Fig. 5 shows the atmospheric structure of our models. From the left panels we see that the HSSBSA NLTE treatment gives rise to surface heating with respect to LTE for $`\mathrm{log}\tau _{\mathrm{cont},5000}<6`$ in the $`M2`$ model. By contrast, the current NLTE treatment gives rise to a $`T_{\mathrm{kin}}`$ structure that is cooler in the upper atmosphere. The HSSBSA treatment yields a surface cooling with respect to LTE in the $`M3`$ model whereas the more complete NLTE treatment produces higher $`T_{\mathrm{kin}}`$ values in the upper atmosphere. For both models, the more complete NLTE treatment gives rise to a $`T_{\mathrm{kin}}`$ structure that is closer to the LTE structure than that of the less complete NLTE treatment. We note that the presence or absence of this NLTE surface heating or cooling in the models will affect the cores of strong lines that form near the top of the atmosphere. NLTE effects on the $`T_{\mathrm{kin}}`$ structure are negligible in the $`M1`$ model. From the right panels of Fig. 5 we see that NLTE effects cause a slight reduction in $`P_{\mathrm{gas}}`$ around $`\mathrm{log}\tau _{\mathrm{cont},5000}=2`$ in the $`M1`$ model and a general reduction throughout the atmosphere in the $`M2`$ model for $`\mathrm{log}\tau _{\mathrm{cont},5000}<2`$. Both NLTE treatments give approximately the same result. For the $`M3`$ model, both NLTE treatments give pressure structures that are close to the LTE structure except for a slight reduction around $`4<\mathrm{log}\tau _{\mathrm{cont},5000}<2`$. By comparing the $`P_{\mathrm{gas}}`$ and $`P_\mathrm{e}`$ structure, we note that the NLTE deviations in $`P_{\mathrm{gas}}`$ in all models mirror those in $`P_\mathrm{e}`$. The NLTE behavior of $`P_\mathrm{e}`$ was discussed above in connection with NLTE H and He. ### 4.3 Flux distribution Figs. 6 to 8 show the distribution of the emergent flux, $`F_\lambda (\tau =0)`$ produced by the models in the observer’s (Eulerian) frame. In the upper panel we show $`F_\lambda `$ at the computed resolution of $`\mathrm{\Delta }\lambda =1.0`$Å for $`\lambda <900`$Å, and $`\mathrm{\Delta }\lambda =0.5`$Å for $`\lambda >900`$Å. The effect of massive line blanketing on the spectrum can be seen for $`\mathrm{log}\lambda <3.5`$. In the lower panel we show the same $`F_\lambda `$ distributions after smoothing with a boxcar function of width $`\mathrm{\Delta }\lambda =100`$Å in the far $`UV`$ and $`\mathrm{\Delta }\lambda =50`$Å in the near $`UV`$, optical and near infrared ($`IR`$). The smoothing approximately reproduces the resolution of intermediate band photometry and wide band spectrophotometry and allows differences in the overall $`F_\lambda `$ level to be more easily discerned. ##### From the lower panels we see that NLTE effects lead to large enhancements in $`F_\lambda `$ for $`\mathrm{log}\lambda <3.1`$ in the $`M1`$ model and for $`\mathrm{log}\lambda <2.95`$ in the $`M2`$ model. By contrast, NLTE effects lead to a reduction in the UV flux in the $`M3`$ model for $`\mathrm{log}\lambda <2.95`$. For all three models, the current NLTE treatment increases the size of the NLTE deviation in $`F_\lambda `$ significantly. The opacity sources that have the largest effect on the emergent flux below the Lyman edge ($`\mathrm{log}\lambda <2.95`$) are H I $`bf`$ absorption and line absorption. We see from Figs. 1 through 4 that the H I concentration is always significantly reduced by the effect of NLTE at depths where $`\tau _{\mathrm{cont},5000}1`$, and is often enhanced higher up in the atmosphere. Furthermore, the current NLTE treatment serves variously to enhance or diminish the size of the NLTE deviation in H I concentration. We may expect that, all else being equal, models in which the emergent flux in the Lyman continuum arises from deeper in the atmosphere will be brighter in the UV in the case of NLTE because the reduction in H I will allow flux to escape from deeper, hotter layers. By contrast, models in which this flux arises from higher up in the atmosphere will be dimmer in the case of NLTE. The extent of the NLTE effect on the emergent flux for particular NLTE treatments will by modified by how that treatment effects the H I concentration. ##### The situation is complicated by the competing effect of line opacity in determining the emergent UV flux. The net effect of NLTE departures on all the metals that contribute significant line opacity may serve either to brighten or dim the flux. From an examination of the concentrations of the dominant Fe ionization stages in Figs. 1 through 4, we see that in the case of $`M1`$ the amount of Fe line blanketing is enhanced by the current NLTE treatment compared to the other two treatments, and in $`M2`$ and $`M3`$ Fe line blanketing is reduced by both NLTE treatments. It is not clear how this can be reconciled with the effects of NLTE on the flux seen in Figs. 6 through 8, and we conclude that the competing influence of NLTE on all the various absorbers that contribute significant opacity is too complex for a simple correlation between the emergent flux and the concentration of any one absorber to be apparent. Finally, we note that generally the flux on the Wien side of the Planck distribution is a sensitive indicator of $`T_{\mathrm{eff}}`$. Therefore, the significant NLTE effects on the UV flux seen here must be included accurately in models. ### 4.4 Completeness Fig. 9 shows the overall $`F_\lambda `$ distribution for the $`M2`$ model with complete line blanketing due to all species that contribute significant line opacity (whether they are treated in LTE or NLTE), and with line blanketing due only to those species treated in NLTE, as indicated in Table 3. A comparison of the two distributions allows an assessment of the completeness of the line opacity that is now treated in NLTE. This is important for assessing the accuracy of the emergent flux spectrum, and also of the equilibrium atmospheric structure because line opacity is an important term in the radiative equilibrium of the atmosphere. ##### The most obvious difference between the two $`F_\lambda `$ distributions is that the Lyman photo-ionization edge at $`\mathrm{log}\lambda =2.96`$ is “softened” by line blanketing to a greater extent when LTE lines are present. The missing NLTE UV line opacity will mostly be accounted for once Fe group species that have a rich line spectrum, such as Cr III and Ni II and III and Co II have been added in NLTE. From Table 1 it can be seen that we already have the facility to compute Ni in NLTE up to stage VI and Co up to stage III, but, as noted in Section 3 and in Table 3, these stages have not been included in these calculations for the sake of computational expediency. Also, there is a noticeable underblanketing in the $`3.1<\mathrm{log}\lambda <3.5`$ range in the treatment with NLTE lines only compared to the fully blanketed treatment. Nevertheless, from the qualitative similarity of the two $`F_\lambda `$ distributions we can also see that the lines treated in NLTE account for the vast majority of the total line opacity. ### 4.5 $`UV`$ and optical spectra Figs. 10 to 12 show the moderate resolution spectrum in three sample regions from the mid-$`UV`$ to the near-IR. Like HSSBSA we find that the quasi-LTE spectrum with a single scattering albedo incorporated into the line source function gives rise to lines that are approximately equal in strength to NLTE lines. The quasi-LTE approach gives rise to inaccurate line strengths for many particular lines, but does not systematically over-predict or under-predict line absorption. Inspection of Figs. 10 and 11 shows that the HSSBA and the current NLTE treatment give rise to different line profiles for many lines, but the differences are not systematic. Differences between the two NLTE treatments are particularly apparent in the $`UV`$ region of the $`M2`$ model as shown in the top panel of Fig. 11. Note that differences in the profiles of individual lines that arise from the NLTE treatment may in general be due to a combination of effects: 1) changes in the line source function, $`S_\nu (\tau )`$, as a result of changes in the populations, $`n_\mathrm{l}`$ and $`n_\mathrm{u}`$, of the levels connected by the transition and by the inclusion of a properly calculated scattering contribution in $`S_\nu (\tau )`$, 2) changes in the background continuous and line opacity due to NLTE effects in other species, and 3) changes in the atmospheric structure such as those discussed in Section 4.2. In general, NLTE effects should be included in as complete a way as possible to accurately calculate any particular line profile. ## 5 Conclusions Generally, we find that the current, more complete NLTE calculation gives rise to qualitatively the same results as the previous extensive NLTE investigation of nova models, that of HSSBSA. NLTE may greatly affect: 1) the chemical concentration of species that are important to the total opacity of the model, such as H I and Fe, or are astrophysically interesting, such as CNO, Mg, Al, and Ca, 2) the overall flux distribution, particularly in the $`UV`$ ($`\lambda <1000`$Å), which is a sensitive $`T_{\mathrm{eff}}`$ indicator, and where NLTE effects increase the flux by several orders of magnitude, 3) the strength of individual lines throughout the observable $`UV`$ to $`IR`$ region where NLTE causes particular lines to be either weaker or stronger than those of a quasi-LTE calculation that crudely incorporates NLTE effects with a single scattering albedo, and 4) to a lesser extent, the atmospheric structure, particularly the $`P_{\mathrm{gas}}`$ structure in models of $`T_{\mathrm{eff}}\mathrm{25\hspace{0.17em}000}`$ K. However, the current NLTE treatment gives rise to specific deviations from LTE values for many of these quantities that differ significantly from those of the HSSBSA treatment. In particular, we note that for the hotter $`M2`$ model ($`T_{\mathrm{eff}}=\mathrm{25\hspace{0.17em}000}`$ K), in which NLTE effects are generally larger than for the cooler $`M1`$ model ($`T_{\mathrm{eff}}=\mathrm{15\hspace{0.17em}000}`$ K), the more complete NLTE treatment gives rise to an H I concentration, $`T_{\mathrm{kin}}`$ and $`P_{\mathrm{gas}}`$ structure, and $`UV`$ flux level that are all significantly closer to the LTE values than those of the less complete HSSBSA NLTE treatment. This result, which may seem counterintuitive at first, results from the increased number of channels through which the gas can thermalize when more species are included in the overall NLTE solution. Note that increasing the number of NLTE species does not necessarily drive the solution for any model atmosphere closer to LTE. Rather, for the $`M2`$ model in particular, the increase in the number of thermalization channels happens to have a dominant effect on the solution when the number of NLTE species is increased from the HSSBSA set to the current set. ##### In general, NLTE should be incorporated in as complete a way as possible for accurate structure and synthetic spectrum calculations for nova models in the 15 000 to 35 000 K $`T_{\mathrm{eff}}`$ range. In regard to the first point above, we draw special attention to the indirect effects that treating any one species in NLTE may have on any other species. The atmospheric structure and the $`e^{}`$ concentration both act as means of coupling all the species, including those that are treated in LTE. Generally, all species that contribute significantly to the $`e^{}`$ reservoir by way of partial ionization, or whose line or continuous opacity is large enough to significantly effect the equilibrium structure of the atmosphere, must be treated in NLTE to insure an accurate result for any particular species. This has a bearing on, for example, attempts to infer model parameters and abundances from fitting the $`UV`$ and optical light curves of novae during the optically thick wind phase of their outburst. This work was supported in part by NASA ATP grant NAG 5-3018 and LTSA grant NAG 5-3619 and NSF grant AST-9720804 to the University of Georgia, and by NSF grant AST-9417242, NASA grant NAG5-3505 and an IBM SUR grant to the University of Oklahoma. Some of the calculations presented in this paper were performed on the IBM SP2 and SGI Origin 2000 of the UGA UCNS, at the San Diego Supercomputer Center (SDSC) and at the National Center for Supercomputing Applications (NCSA), with support from the National Science Foundation, and at the NERSC with support from the DoE. We thank all these institutions for a generous allocation of computer time.
no-problem/9906/astro-ph9906172.html
ar5iv
text
# Global Textures and the Formation of Self-Bound Gravitational Systems ## 1 Introduction The emergence of structures in the universe is one of the most important problems of modern cosmology. The difficulty is to understand how an initially homogeneous mass distribution evolved into its present clumpy state. Generally, it is accepted that structures were initiated by small density fluctuations formed at the early universe, and that the subsequent clustering was produced by gravitational instability (e.g. Kolb & Turner 1990). At the moment, there are two theoretical approaches proposed to explain the origin of the primordial inhomogeneities. In the context of inflationary models, they result of quantum-induced Gaussian fluctuations amplified over an accelerated phase of cosmic expansion. On the contrary, in topological defect scenarios, non-Gaussian fluctuations are produced during spontaneous symmetry breaking at phase transitions in the early universe (e.g. Vilenkin & Shellard 1994). Several types of topological defects can be formed depending on the kind of symmetry that is broken. Here, we are interested in a specific topological defect formed during the symmetry breaking of non-Abelian groups: a global texture. This defect has been proposed as a possible source of galaxy formation in the universe (Turok 1989). The development of structures based on this topological remnant is related to the evolution of textures knots. On scales larger than horizon, textures knots are regions where the scalar field winds around the vacuum manifold in a non-trivial way. As knots come inside the horizon, they collapse at the speed of light down to an infinitesimal scale where they unwind themselves emitting spherical waves of outgoing radiation. This collapse produces an overdense region onto which matter is attracted, possibly leading to the formation of non-linear objects (Turok & Spergel 1990). The most desirable way to study structure formation produced by textures knots is through numerical simulations like those carried out by Park, Spergel & Turok (1989), Spergel et al. (1991) and Cen et al. (1991). All these studies indicate important effects of textures on the distribution of matter, like coherent large-scale flows and clustering of galaxies fairly consistent with the correlation function observed in the nearby universe. However, we should keep in mind that the dynamics of textures is highly non-linear since the decoupling era (Albrecht, Battye & Robinson 1997), and that it is not simple to trace the evolution of the fluctuations over this long time interval. Thus, all these numerical works are strongly dependent on the choice of the initial conditions for the N-body simulations (Deruelle et al. 1997). Actually, this is not the only problem with the texture scenario. The work of Pen, Seljak & Turok (1997) shows that the texture model has strongly suppressed acoustic peaks in comparison to current observations of CMB anisotropies. Also, when COBE normalized, the model produces a low value of the rms density variation on scales of 8 Mpc ($`\sigma _80.23`$), while a higher value of this parameter (closer to unity) is necessary in order to get agreement with the observed galaxy clustering. These results clearly pose serious difficulties to the texture scenario. However, this does not mean that the model should be completely discarded at the present time, since a decisive, model discriminating test will be possible only after the analysis of the high resolution CMB maps which will be available in the next years from the MAP and PLANCK satellite missions. Hence, for the moment, it remains valid to work with texture knots as a viable way to form galaxies in the universe. Indeed, an alternative way to study structure formation via global textures is to develop simple analytic models to describe the general properties of the defect dynamics. For instance, Gooding, Spergel & Turok (1991) studied the formation of bound objects induced by a number of collapsing knots in a $`\mathrm{\Omega }_0=1`$ CDM-dominated universe. An important result of this study is the prediction of early spheroidal formation in the universe, suggesting that by $`z50`$, $``$3% of the mass of the universe has formed non-linear objects with mass greater than $`10^6M_{}`$ and that most objects larger than $`10^{12}M_{}`$ formed by $`z23`$. In the present work, we use a Newtonian approximation to investigate the formation of self-bound gravitational systems seeded by global textures. However, instead of applying a statistical approach to the contribution of many individual knots, we are only interested in the local influence of a single texture knot on the process of galaxy formation. Our aim is just to obtain a quantitative criterion for the collapse of a spherical distribution of matter under a texture field. ## 2 The Collapse Criterion Distributed field gradients induced by a number of global textures are supposed to give a larger contribution to density inhomogeneities than that produced by single isolated knots (e.g. Vilenkin & Shellard 1994). If this is correct, then, the effort to probe the process of structure formation seeded by these defects will require large dynamic range simulations in computers of latest technology. However, the demand for such numerical works does not mean that analytical developments are unnecessary. On the contrary, they are particularly useful to give a supplementary and intuitive picture of the defect dynamical contribution on structure formation. Specifically, in this work, we study the process of accretion of matter onto a single texture knot. This problem has been studied by several authors over the years. But now, our emphasis is on the Newtonian treatment and the possibility of taking a region where the assumption of an isolated texture is reasonable. In order to do that, let’s consider an initially spherical and homogeneous distribution of matter with density $`\rho _b`$ when the texture field is turned on. The Newtonian density associated to the texture can be obtained by using the Einstein equations in the weak field approximation. By numerically solving the Barriola-Vaschaspati equations for a self-gravitating texture, Guéron & Letelier (1997) found that the weak field condition is fulfilled and the corresponding Newtonian density is $$\rho _\mathrm{N}=\frac{8\eta ^2(r^2t^2)}{(r^2+t^2)^2}$$ (1) where $`\eta `$ is the energy scale of symmetry breaking. This density is negligible except when the texture size, $`t`$, is small compared to the physical scale $`r`$. At the same time, the region where is reasonable to take account a single texture configuration, ignoring the effects from other textures, is limitted by $$(\mathrm{\Delta }t^2+r^2)^{1/2}H^1$$ (2) (Nötzold 1991), where $`\mathrm{\Delta }t=t_{}t`$, $`t_{}`$ is the time at which the texture is collapsed and $`H`$ is the cosmic expansion rate. For simplicity, we will only study the behaviour of regions where $`r|t|`$, so we can use an effective density defined as $$\rho _\mathrm{T}8\eta ^2/r^2.$$ (3) For such regions, $`|\mathrm{\Delta }t|<r`$ and so condition (2) is simply $`rH^1/\sqrt{2}`$. The total density within r is $`\rho (r)=\rho _\mathrm{T}+\rho _\mathrm{b}`$ and the density contrast can be expressed by $$\frac{\delta \rho }{\rho }=\frac{\rho (r)\rho _b}{\rho _\mathrm{b}}=\frac{\rho _\mathrm{T}}{\rho _\mathrm{b}}$$ (4) so that the motion of a thin shell of particles located at r is governed by the equation $$\frac{d^2r}{dt^2}=\frac{GM}{r^2},\mathrm{where}M=\frac{4\pi r^3}{3}\rho _\mathrm{b}(1+\overline{\delta })$$ (5) and the average density contrast is $$\overline{\delta }=\frac{3}{4\pi r^3}_0^r\left(\frac{\rho _\mathrm{T}}{\rho _\mathrm{b}}\right)4\pi r^2𝑑r.$$ (6) As usual, it will be assumed here that the mass within the shell is constant. In fact, this is strictly correct only after a time interval of $`t_{}`$, when the texture mass is converted in massless goldstone bosons, leaving behind the other material components: baryonic and dark matter. Thus, we are tacitly assuming that $`t_{}`$ is much smaller than the dynamical time of the system, but large enough to allow the texture mass to dominate the first steps of the gravitational instability. In idealizing the process this way, we hope to make it well-posed mathematically, although a justification of the idealization will be only possible from the agreement (or not) of our predictions with the data. Now, supposing that the peculiar velocities at $`t=t_\mathrm{i}`$ are negligible, the first integral of Eq.(5) will give the total energy of the shell and can be written as $$E=K_\mathrm{i}\mathrm{\Omega }_\mathrm{i}[\mathrm{\Omega }_\mathrm{i}^1(1+\overline{\delta }_\mathrm{i})],$$ (7) (e.g. Padmanabhan 1993). The subscript $`i`$ indicates the initial time $`t_\mathrm{i}`$ when the radius of the shell is $`r_\mathrm{i}`$, the kinetic energy is $`K_\mathrm{i}`$ and the density parameter is $`\mathrm{\Omega }_\mathrm{i}`$. Naturally, the condition $`E<0`$ for a self-bound system is $$\overline{\delta }_\mathrm{i}>(\mathrm{\Omega }_\mathrm{i}^11)\delta _\mathrm{c}.$$ (8) For a dust configuration ($`p=0`$), we have $$\mathrm{\Omega }_\mathrm{i}=\frac{\mathrm{\Omega }_0(1+z_\mathrm{i})}{1+z_\mathrm{i}\mathrm{\Omega }_0}$$ (9) and so $$\delta _\mathrm{c}=\frac{1\mathrm{\Omega }_0}{\mathrm{\Omega }_0(1+z_\mathrm{i})}$$ (10) (e.g. Börner 1988). Therefore, for the case of a flat universe, $`\delta _\mathrm{c}=0`$ and any overdense region with $`\overline{\delta }_\mathrm{i}>0`$ will collapse. On the other hand, for the case of an open universe, only regions where $`\overline{\delta }_\mathrm{i}>\delta _\mathrm{c}`$ can produce self-bound gravitational systems. In this case, introducing definition (3) in Eq.(6) we directly find $$\overline{\delta }_\mathrm{i}=\frac{24\eta ^2}{\rho _\mathrm{b}(t_\mathrm{i})r_\mathrm{i}^2}.$$ (11) Now, it is easy to foresee the destiny of a spherical distribution of matter under a texture field when $`\mathrm{\Omega }_0<1`$. Introducing Eq.(11) in condition (8), it is simple to show that $$r_\mathrm{i}<\sqrt{\frac{24\eta ^2\mathrm{\Omega }_\mathrm{i}}{\rho _\mathrm{b}(t_\mathrm{i})(1\mathrm{\Omega }_i)}}.$$ (12) Recalling that $$\rho _\mathrm{b}(t_\mathrm{i})=\mathrm{\Omega }_\mathrm{i}\rho _\mathrm{c}(t_\mathrm{i})\mathrm{and}\rho _\mathrm{c}(t_\mathrm{i})=\frac{3H_\mathrm{i}^2}{8\pi G},$$ (13) we reach $$r_\mathrm{i}<\frac{8\eta }{H_\mathrm{i}}\sqrt{\frac{G\pi }{1\mathrm{\Omega }_\mathrm{i}}}R_\mathrm{c},$$ (14) where $`R_\mathrm{c}`$ is the critical radius for the collapse. Once more, assuming a Friedmann model with vanishing pressure and with no cosmological constant $$H_\mathrm{i}^2=H_0^2[\mathrm{\Omega }_0(1+z_\mathrm{i})^3+(1\mathrm{\Omega }_0)(1+z_\mathrm{i})^2]$$ (15) (e.g. Sandage 1961), we can finally express the collapse criterion for any overdense region under a texture field at any epoch $`z_\mathrm{i}`$: $$r_\mathrm{i}<\frac{8\eta }{H_0}\sqrt{\frac{G\pi (1+z_\mathrm{i}\mathrm{\Omega }_0)}{(1\mathrm{\Omega }_0)[\mathrm{\Omega }_0(1+z_\mathrm{i})^3+(1\mathrm{\Omega }_0)(1+z_\mathrm{i})^2]}}.$$ (16) ## 3 The Choice of the Energy Scale From Eq.(16) we see that, assuming specific values for $`H_0\mathrm{and}\mathrm{\Omega }_0`$, the critical radius for the collapse will depend basically on the the energy scale of symmetry breaking. Generally, global textures arise after a phase transition during the break of symmetry described by the Grand Unified Theory (GUT), when the strong interaction is unified with the eletroweak force. Current estimates of this unification lead to an energy scale of $`10^{16}`$ GeV (Gleiser 1998). In the same way, if the observed fluctuations in the cosmic background radiation are due to textures, we must have $`(\delta T/T)8G\eta ^2`$ which implies $`\eta 10^{16}`$ GeV in order to obtain an amplitude consistent with the COBE measurements, $`\delta T/T10^5`$ (e.g. Vilenkin & Shellard 1994). In view of these arguments, the energy scale of $`10^{16}`$ GeV seems to be the natural choice of $`\eta `$. However, let’s proceed here to find a new and independent argument for this choice. First of all, assuming $`\mathrm{\Omega }_0=0.2`$ and $`h=0.75`$ $`(h=H_0/100\mathrm{km}\mathrm{s}^1\mathrm{Mpc}^1)`$, we plot the behaviour of $`R_\mathrm{c}`$ as a function of the epoch $`z_\mathrm{i}`$ using $`\eta `$ as a free parameter (see Figure 1). Note that for earlier epochs $`R_\mathrm{c}`$ is smaller (independent of $`\eta `$) indicating that the first bound objects to be formed via textures knots would have small sizes and masses. In fact, isocurvature fluctuations like those produced by topological defects can evolve only after the decoupling ($`z10^3`$), when the Jeans mass is $`10^6M_{}`$. Thus, clumps with this mass should form first and evolve into larger systems through gravity. Obviously, the details of the evolution of clumps up to form typical galaxies and clusters is a rather complex process, involving the thermal history of the gas, star formation phenomena and the mutual interaction between the clumps. It is beyond our aim to address this problem as a whole. Instead, we proceed to simplify it by supposing that a protogalaxy just consist of a regular distribution of matter (possibly sub-galactic clumps) under a texture field where we can apply our collapse criterion. That being the case, an important quantity to be found is the mass $`M_\mathrm{c}`$ related to a given radius $`R_\mathrm{c}(z_\mathrm{i})`$. For a homegeneous universe, the non-relativistic mass within a physical scale $`\lambda `$ is given by $$M(\lambda )=1.45\times 10^{11}(\mathrm{\Omega }_0h^2)\lambda _{\mathrm{Mpc}}^3M_{}$$ (17) (e.g. Padmanabhan 1993). This quantity is conserved during the expansion and corresponds to the uncollapsed mass of a physical scale $`\lambda `$. If we take $`\lambda =R_\mathrm{c}`$, it is possible to study the behaviour of $`M_\mathrm{c}`$ as a function of the epoch $`z_\mathrm{i}`$ and the parameter $`\eta `$. In Figure 2 we see that, for a same epoch $`z_\mathrm{i}`$, the value of $`M`$ will strongly depend on $`\eta `$. From the comparison between $`M_\mathrm{c}`$ and the Jeans mass, we also note that the lowest energy scale which is able to form self-bound systems over all the astrophysical range ($`10^610^{15}M_{}`$) at any pos-recombination time $`z_\mathrm{i}`$ is $`\eta =10^{16}`$ GeV. At this energy scale, $`R_\mathrm{c}`$ is well inside $`H^1/\sqrt{2}`$ (see Figure 1), which defines regions where we can consider the effects of isolated texture knots. This result can be taken as a qualitative argument for the choice of $`\eta =10^{16}`$ GeV and so we are assuming this value for the rest of the work. ## 4 Galaxy Formation In order to apply our collapse criterion to the specific problem of galaxy formation, we should follow the evolution of overdense regions containing typical galaxy masses ($`10^{11}10^{12}M_{}`$). If these respect the condition (8), they will expand to a maximum radius, then collapse and eventually virialize. The dynamical properties of the resultant systems can be estimated from the spherical model for non-linear collapse. Thus, at the epoch $$(1+z_{\mathrm{col}})=0.36\overline{\delta }_\mathrm{i}(1+z_\mathrm{i}),$$ (18) the dissipative component of the system (i.e. baryons) reaches the virialization with the following approximated properties: $$r_{\mathrm{vir}}=163(1+z_{\mathrm{col}})^1\mathrm{\Omega }_0^{1/3}(M/10^{12}M_{})^{1/3}h^{2/3}\mathrm{kpc}$$ (19) and $$\sigma _{\mathrm{vir}}=126(1+z_{\mathrm{col}})^{1/2}\mathrm{\Omega }_0^{2/3}(M/10^{12}M_{})^{1/3}h^{1/3}\mathrm{km}\mathrm{s}^1$$ (20) (e.g. Padmanabhan 1993). Before applying these expressions to galaxy scales, it would be interesting to assure that the initial density contrast $`\overline{\delta }_\mathrm{i}`$ can form a self-bound system before the present epoch, $`z_{col}>0`$. We directly find such a condition from Eq.(18): $$\overline{\delta }_\mathrm{i}>2.78(1+z_\mathrm{i})^1.$$ (21) For the case of a flat model this is a simple condition for the collapse. However, for an open universe, we also should have $`\overline{\delta }_\mathrm{i}>\delta _\mathrm{c}`$ in order to collapse the overdense region. Using Eq.(10) and taking the limitting case when $`\delta _\mathrm{c}=2.78(1+z_\mathrm{i})^1`$, we conclude that $`\mathrm{\Omega }_0<0.26`$ is the necessary condition for the collapse with $`z_{\mathrm{col}}>0`$. In Figure 3, we plot $`\overline{\delta }_\mathrm{i}`$ as a function of $`z_\mathrm{i}`$ for initial radii that correspond to typical galaxy masses. The dotted line is the limit $`2.78(1+z_\mathrm{i})^1`$. Note that for $`M=10^{11}M_{}`$, the collapse is only possible for $`z_\mathrm{i}33`$ ($`z_{\mathrm{col}}9`$) if $`\mathrm{\Omega }_0=0.2`$ and $`z_\mathrm{i}25`$ ($`z_{\mathrm{col}}8`$) if $`\mathrm{\Omega }_0=1.0`$; while for $`M=10^{12}M_{}`$, $`z_\mathrm{i}15`$ ($`z_{\mathrm{col}}5`$) if $`\mathrm{\Omega }_0=0.2`$ and $`z_\mathrm{i}11`$ ($`z_{\mathrm{col}}4`$)if $`\mathrm{\Omega }_0=1.0`$. These ranges in $`z_\mathrm{i}`$ produce associated loci in the $`\sigma r`$ plane (see Figure 4). These loci show that our simple model texture knot$`+`$spherical collapse is able to produce self-bound gravitational systems with dynamical properties quite similar to real galaxies. In fact, the comparison with data compiled from the literature shows a significant agreement between our predictions and typical galaxies, while it is also clear in Figure 4 an important trend towards larger radii and lower velocity dispersions in our bound systems. At the same time, the epoch when such objects are presumed to reach the virial equilibrium ($`z_{\mathrm{col}}`$) is consistent with the results of several studies which point out an epoch of galaxy formation by $`z5`$ (e.g. Peebles 1993; Pahre 1998). Now, by extending the approach developed in this section to a wider astrophysical range ($`10^610^{14}M_{}`$), it is possible to depict an intuitive picture of the structure formation process when it is seeded by textures knots. The results for both open and flat universes are summarized in Tables 1 and 2, respectively, where the columns are as follows: (1) is the mass of the systems in solar units; (2) is the initial redshift we take; (3) is the initial radius of the region to be collapsed in Mpc; (4) is the initial average density contrast; (5) is the redshif at which the systems reach the virialization; (6) is the time interval of the collapse and virialization process (the dynamical time); finally, (7) and (8) are the radius in kpc and the velocity dispersion in km/s of the virialized systems. A simple analysis of the data assembled in these tables reveals some points we should emphasize. First of all, note that structure formation ceases at earlier times in the low $`\mathrm{\Omega }_0`$ universe in comparison to the flat case. This is expected as long as matter has a smooth distribution in high $`\mathrm{\Omega }_0`$ universes so that it contributes to the expansiong rate but does not contribute to the gravitational potential and so it slows down the growth of fluctuations. At the same time, note that the process of structure formation in our model begins at a much earlier epoch in comparison to standard CDM models, where most of the objects with $`10^6M_{}`$ are formed at the epoch $`z=50/b`$ ($`b`$ is the biasing factor) (e.g. Peebles 1993). Therefore, even an unbiased CDM model would present a less advanced structure formation stage for a given redshift in comparison to our model. Such an early self-bound system formation may have serious consequences. In particular, we form the first generation of $`10^6M_{}`$ objects at $`z=135`$ ($`\mathrm{\Omega }_0=0.2`$) and $`z=112`$ ($`\mathrm{\Omega }_0=1.0`$). These objects could be related to an early star formation phase in the universe and its consequent reionization. At the same time, these early spheroids may develop massive black holes onto which matter would be accreted, possibly forming an early generation of active galaxies. Still concerning to the chronology of structure formation, Gooding, Spergel & Turok (1991) found that by $`z50`$, $`3`$% of the mass of the universe has formed non-linear objects of mass greater than $`10^6M_{}`$ and that most of objects larger than $`10^{12}M_{}`$ form by $`z23`$. These results should be compared to our findings, while in the present work we have no statistics on the fraction of matter which is collected in bound objects of a specific mass in a given redshift. Note, however, that by $`z50`$ we already formed systems of $`10^610^8M_{}`$ and by $`z3`$ we form objects with $`10^{13}10^{14}M_{}`$. Thus, in some measure, our results are consistent with Gooding, Spergel & Turok (1991) On the other hand, a disadvantage of our model refers to the velocity dispersions of the collapsed systems which are sistematically lower than those observed in astrophysical objects of same mass. Up to galaxy scales this effect is accompanied with larger radii and so it can be taken as a simple consequence of the virial theorem. On the contrary, the lower velocity dispersion effect seems to be particularly serious in larger scales, like galaxy clusters, where the model predicts objects with radii similar to those found in real clusters. Actually, this is the opposite one could expect in a texture scenario, where the positive skewness of the mass fluctuations induces higher velocity dispersions for objects of a given mass than (as an example) in standard CDM models (e.g. Bartlett, Gooding & Spergel 1993). We think this trend is probably due to the fact we are considering only the dissipative baryonic component of matter inside $`r_{\mathrm{vir}}`$. By adding dark matter within this radius we could recover values of velocity dispersion nearer to reality. For instance, in the case of a galaxy cluster scale ($`10^{14}M_{}`$), if the total mass (dissipative plus non-dissipative components) inside $`r_{\mathrm{vir}}`$ reached $`10^{15}M_{}`$ the velocity dispersion of the collapsed system would be 583 km/s ($`\mathrm{\Omega }_0=0.2`$) or 1475 km/s ($`\mathrm{\Omega }_0=1.0`$), which could be taken as more “normal” values. However, if this is the right solution to the problem, we would have to assume specific distributions for the luminous and dark matter components when considering different objects. At the same time, we cannot discard the possibility that the simplicity of our model texture knot$`+`$spherical collapse is not taking account all of the physics relevant to the structure formation process. Indeed, we are ignoring interactions between clumps, star formation and hydrodynamical effects. These physical mechanisms are probably important over the systems evolution and their inclusion would change (to some extent) the outcomes of the model. ## 5 Discussion and Summary In this work, we developed a quantitative criterion for the collapse of a spherical distribution of matter under a single texture field. Further, we applied this criterion to form galaxies in an idealized smooth universe where texture knots can be taken as isolated sources for gravitational collapse. Despite the simplicity of the model, the properties of the systems which collapse and virialize respecting our criterion are similar to real astrophysical objects. A drawback of the model is the trend towards larger radii and lower velocity dispersions in our bound systems in comparison to typical values found in the literature for real objects of same mass. This trend is specifically significant for velocity dispersions and could be related to the quantity of dark matter inside the virialization radius. However, we would like to emphasize that the most important achievement in this work is the determination of a collapse criterion using simple arguments and the possibility of applying it in a direct way to predict astrophysical objects. This means that local effects of isolated knots may be important to the development of structures in the universe and suggests that individual textures (not only their dynamical collective effect) should be taken account in galaxy formation models. In the following, we summarize the main results of this work: 1. We showed that, under specific conditions, the destiny of an overdense region under a single texture field will basically depend on its initial size and the energy scale of symmetry breaking $`\eta `$ (if $`\mathrm{\Omega }_0<1`$). For those regions smaller than $`R_\mathrm{c}`$, the system will expand to a maximum radius and then inevitably collapse. 2. We found an additional argument for the choice of $`\eta `$ as $`10^{16}`$ GeV. This is simply the lowest energy scale which allows the formation of structures on scales relevant to astrophysical objects ($`10^610^{15}M_{}`$) at any time $`z_\mathrm{i}<1000`$. 3. We showed the loci of objects with masses $`10^{11}10^{12}M_{}`$ in the $`\sigma r`$ plane agree well with the observed properties of real galaxies. 4. Our results also indicate an epoch of galaxy formation fairly consistent with other independent studies which point out it should have been by $`z5`$. 5. Our model produces a chronology of structure formation which begins at earlier times in comparison to other models, in particular CDM models. 6. The existence of spheroids by $`z100`$ suggests the possibility of associated phenomena like early star formation process, reionization of the universe, production of black holes and active galaxies. ## Acknowledgments We thank the referee A. Liddle for useful suggestions. We also thank R. de Carvalho and R. Coziol for the critical reading of the paper. Finally, we thank the finantial support of the Brazilian FAPESP and the CNPq.
no-problem/9906/astro-ph9906354.html
ar5iv
text
# The origin and formation of cuspy density profiles through violent relaxation of stellar systems ## 1 Introduction Recent Hubble Space Telescope observations (Crane et al. 1993; Ferrarese et al. 1994; Jaffe et al. 1994; Lauer et al. 1995; Byun et al. 1996; Gebhardt et al. 1996; Faber et al. 1997) have revealed that elliptical galaxies do not have constant density cores but have cusps which continue toward the centre until the resolution limit. In addition, it has been found that the density profiles of the cores are divided into two types: one is a shallow cuspy core represented by $`\rho r^n`$ with $`0.5n1`$ for bright elliptical galaxies, and the other is a steep cuspy core described by approximately $`\rho r^2`$ for faint ones (Merritt & Fridman 1996). Several cosmological simulations, based on a standard cold dark matter scenario, have demonstrated that cuspy density profiles with $`\rho r^1`$ might be a natural end-product of hierarchical clustering (Dubinski & Carlberg 1991; Navarro, Frenk & White 1996, 1997). Recently, Fukushige & Makino (1997) have pointed out from their high resolution simulation with about $`8\times 10^5`$ particles that the resulting central density profiles of dark matter haloes are steeper than $`\rho r^1`$ in a hierarchical clustering scenario, and that merging processes among clumps with different binding energies would play a significant role in the formation of such steep density cusps. Moore et al. (1998) have also demonstrated that very high resolution simulations with nearly $`3\times 10^6`$ particles yield steep inner density profiles with $`\rho r^{1.4}`$ (see however Kravtsov et al. 1998). On the other hand, shallower density cusps could be formed through merging between galaxies which individually have a massive central black hole, as shown by Makino & Ebisuzaki (1996). For this kind of merging, Nakano & Makino (1999) have explained why a galaxy that has swallowed a massive black hole leaves a shallow cusp in the centre. However, the physical mechanism to generate steep cusps still remains unsolved. From the simulations mentioned above, we might conclude that clumps are needed to produce steep cusps while merging between galaxies with massive central black holes is required to generate shallow cusps. However, it might be too early to abandon a simple dissipationless collapse picture, which includes neither clumpiness nor massive central black holes, for the formation of cusps. We know that the collapse of uniform-density spheres results in large constant density cores (van Albada 1982; Fujiwara 1983), while little is known about the details of the collapse of power-law density spheres. Of course, there are cases for the collapse of non-uniform density spheres. For example, Hozumi, Fujiwara & Kan-ya (1996) have carried out collapse simulations of power-law density spheres only in the initial contraction phases in order to demonstrate the predominance of the growth of the tangential velocity dispersion over that of the radial velocity dispersion until the most contracting phase. In addition, Londrillo, Messina & Stiavelli (1991) examined the collapse of generalized Plummer models; their models have large cores, and steep density haloes with $`\rho r^5`$ or $`\rho r^7`$, so that the collapse behaviour over most parts of the system is considered similar to that of a uniform-density sphere. Burkert (1990) paid attention to degenerate cores arising from the dissipationless collapse of spherical stellar systems with different initial density gradients and virial ratios. In another instance, Cannizzo & Hollister (1992) studied the relation between the power-law index of initial density distributions and the final axis ratios of the systems. However, the structures at sufficiently small radii where the density cusps are observed in real galaxies were not a main concern in these examples. Although many dissipationless collapse simulations have been carried out so far, there has been no sufficiently high resolution to determine fine structures at very small radii, such as density cusps. In addition, the systems in numerical studies were assumed to have isotropic velocity distributions at the beginning. That is, the initial models for dissipationless collapse simulations have never included anisotropic velocity distributions, despite the fact that stellar systems can support velocity anisotropy unlike gaseous systems. In this paper, we examine the influence of initial velocity anisotropy on the relaxed density profiles in the core, assigning enough resolution to analyse the innermost region, and show that radially anisotropic models could produce different cuspy density profiles. In Section 2, we describe the initial models in which anisotropic velocity distributions are taken into account, and the numerical method. In Section 3, the results of collapse simulations are presented. In Section 4, we explain the resulting density profiles from a viewpoint that the maximum phase-space density accessible to the relaxed system is determined at each radius by that of the initial system. Conclusions are given in Section 5. ## 2 Models and Method We carry out collapse simulations of power-law density spheres on the assumption of spherical symmetry. The initial virial ratios, $`\eta 2T/|W|`$, used are $`\eta =0.1`$ and $`\eta =10^{1.5}`$, where $`T`$ is the kinetic energy, and $`W`$ is the potential energy. According to existing simulations (Polyachenko 1981, 1992; Merritt & Aguilar 1985; Min & Choi 1989; Aguilar & Merritt 1990; Londrillo et al. 1991; Cannizzo & Hollister 1992; Udry 1993) like those studied here, our models would be affected by the radial orbit instability if spherical symmetry were not assumed, because of the small initial virial ratios. However, Aguilar & Merritt (1990) have shown that the origin of the radial orbit instability is different from that of the universal density profile observed in elliptical galaxies, that is, the de Vaucouleurs $`R^{1/4}`$ law in projection (see also Burkert 1993). Therefore, we are allowed to assume spherical symmetry as long as we focus only on the spherically averaged density profiles and related physical quantities of collapsed objects. The velocity anisotropy is introduced into the initial models by adopting the distribution function, $`f`$, as $$f=\frac{\rho _0(r)}{\sqrt{2}\pi ^{3/2}\sigma _r\sigma _{}^2}\mathrm{exp}\left(\frac{u^2}{2\sigma _r^2}\right)\mathrm{exp}\left(\frac{j^2/r^2}{\sigma _{}^2}\right),$$ (1) where $`\rho _0`$ is the density, $`\sigma _r`$ is the radial velocity dispersion, and $`\sigma _{}`$ is the tangential velocity dispersion with $`r`$, $`u`$, and $`j`$ being the radius, radial velocity, and angular momentum, respectively. We begin with a moderate power-law density profile given by $$\rho _0(r)=(M/2\pi R_0^2)r^1,$$ (2) where $`M`$ is the total mass and $`R_0`$ is the radius of the sphere. We measure the velocity anisotropy by $`\alpha 2T_r/T_{}`$, where $`T_r`$ and $`T_{}`$ are the kinetic energies in radial and tangential motions, respectively. In our models, $`\alpha `$ is equivalent to $`2\sigma _r^2/\sigma _{}^2`$, as is found from equation (1). We take $`\alpha =0.2,0.5,1,2,6,`$ and 10 for each value of the virial ratios. We use Fujiwara’s (1983) phase-space solver which integrates the collisionless Boltzmann equation directly, with the help of a splitting scheme (Cheng & Knorr 1976). Since this method leads to the results that are not subject to random fluctuations, we can obtain smooth density and velocity dispersion profiles. We employ $`(N_r,N_u,N_j)=(300,301,50)`$, where $`N_r,N_u`$, and $`N_j`$ are the numbers of mesh points along the radius, radial velocity, and angular momentum, respectively. A logarithmic grid is used in the radial direction. An equal interval grid is assigned to the radial velocity while a quadratic interval is adopted for the angular momentum. The units of mass, $`M`$, and the gravitational constant, $`G`$, are taken to be $`M=G=1`$. The unit of length is determined from the relation such that $`\eta \times R_0=1`$. This choice of $`R_0`$ leads to approximately the same core sizes for uniform-density spheres with isotropic velocity distributions (Fujiwara 1983). As a characteristic time-scale for the collapse, we adopt the free-fall time, defined by $`t_{\mathrm{ff}}=\pi \sqrt{R_0^3/8GM}`$. We obtain $`t_{\mathrm{ff}}=35.1`$ for $`\eta =0.1`$ and $`t_{\mathrm{ff}}=197`$ for $`\eta =10^{1.5}`$ in our system of units. The minimum and maximum radii assigned to grid points are set to be $`R_{\mathrm{min}}=0.01`$ and $`R_{\mathrm{max}}=40.0`$, respectively, for $`\eta =0.1`$, and $`R_{\mathrm{min}}=0.01`$ and $`R_{\mathrm{max}}=49.9`$, respectively, for $`\eta =10^{1.5}`$. Test simulations have shown that resulting density and velocity dispersion profiles at small radii do not depend on the replacement of $`R_{\mathrm{min}}=0.01`$ by $`R_{\mathrm{min}}=0.005`$ with more finely divided mesh points in the angular momentum. Hozumi & Hernquist (1995) and Hozumi (1997) traced the orbits of stars on grid points backward to $`t=0`$ every suitable time step in order to avoid the numerical diffusion generated by the repeated interpolation required by the splitting scheme. However, we do not adopt this method in our simulations. This is because we are interested not in the details of the phase-space distributions but mainly in low-order moments of the distribution function such as density and velocity dispersion, in addition to saving computational time. ## 3 Results We stopped the simulations at $`t=120`$ for $`\eta =0.1`$ and at $`t=500`$ for $`\eta =10^{1.5}`$. These times correspond to about $`3t_{\mathrm{ff}}`$. They are, nevertheless, sufficiently long that the resulting cores have relaxed completely. Since our main concern is the core regions, these times satisfy our requirements for analysing the relaxed density and velocity dispersion profiles near those regions. There were practically no escapers for $`\eta =0.1`$, while about 20 per cent of the total mass with positive energies expanded beyond $`R_{\mathrm{max}}`$ and escaped for $`\eta =10^{1.5}`$. Consequently, the total energy was conserved to better than 0.39 per cent for all the models with $`\eta =0.1`$ except for the model with $`\alpha =0.2`$ in which the total energy within $`R_{\mathrm{max}}`$ changed by 1.8 per cent owning to a small fraction of escapers. On the other hand, the total energy within $`R_{\mathrm{max}}`$ changed by at worst 10.7 per cent for all the models with $`\eta =10^{1.5}`$ because of a large fraction of escapers. In Figs. 1a and 1b, we show the resulting density profiles for $`\eta =0.1`$ and $`\eta =10^{1.5}`$, respectively. The half-mass radii, $`r_\mathrm{h}`$, for $`\eta =0.1`$ are about 2.6 to 3.2, while they are about 7.8 to 9.7 for $`\eta =10^{1.5}`$. Thus, our adopted minimum radii correspond to $`R_{\mathrm{min}}0.003r_\mathrm{h}`$ for $`\eta =0.1`$ and $`R_{\mathrm{min}}0.001r_\mathrm{h}`$ for $`\eta =10^{1.5}`$. If the half-mass radius is regarded practically as the effective radius, $`r_\mathrm{e}`$, of the de Vaucouleurs law, the minimum radii used are sufficiently small and correspond to roughly 10 pc, because $`r_\mathrm{e}`$ is typically 3 kpc for bright ellipticals and is smaller for faint ones (Kormendy 1977). We point out that most galaxies are not well resolved inside to 10 pc according to the data cited in the paper of Gebhardt et al. (1996). We can see from Fig. 1 that the density distributions at small radii become more cuspy as $`\alpha `$ increases, and that their slope becomes steeper with increasing $`\alpha `$ for both virial ratios. As a result, the core region, specified by a nearly constant density distribution, decreases with increasing $`\alpha `$. In particular, it is to be noticed that there is no constant density core for the most radially anisotropic models, that is, the models with $`\alpha =10`$. On the other hand, the density profiles at intermediate radii are well-approximated by $`\rho r^{2.1}`$, regardless of the virial ratios and the velocity anisotropy. The relaxed velocity dispersion profiles for $`\eta =0.1`$ and $`\eta =10^{1.5}`$ are shown in Figs. 2a and 2b, respectively. These figures indicate that the relaxed velocity distributions no longer remain anisotropic in the inner region of the cores even though we begin with a model which has a large velocity anisotropy at all radii. Thus, it seems very difficult to produce and maintain velocity anisotropy in the cores only by gravitational collapse. The isotropic region that roughly corresponds to the core becomes smaller, as the velocity anisotropy increases. It is found from Figs. 2a and 2b that the density of the inner regions necessarily decreases with radius because the velocity dispersions rise with radius in the cores. In particular, for $`\alpha =10`$, the velocity dispersions in the cores increase almost linearly with radius in logarithmic scale. Then, the hydrostatic equation requires that the corresponding density profiles should follow a power-law distribution. In reality, as shown by Fig. 2, the imperfect linearity of the velocity dispersion against radius in logarithmic scale does not realize a perfect power-law density distribution. ## 4 Discussion ### 4.1 Phase-space constraint Collisionless stellar systems suffer the constraint that the phase-space density accessible to the relaxed state is less than the maximum phase-space density of the initial state. On the basis of this phase-space constraint, we can explain the density bend which occurs around the edge of the cores. As far as the core generated through violent relaxation is concerned, the phase-space density does not decrease substantially from the initial value at least for spherically symmetric systems (Fujiwara 1983; Hozumi 1997). Our adopted initial virial ratios are small enough to induce a radial orbit instability, unless spherical symmetry is imposed. Nevertheless, it may be unlikely that the phase-space density in the core suffers a substantial decrease after violent relaxation according to the results of May & van Albada (1984): they have shown that the phase-space density in the core is almost conserved after violent relaxation even for aspherical collapses, although they examined this behaviour only for initially homogeneous density distribution models. The effects of aspherical collapses through the radial orbit instability are left to be investigated as a future problem. Here, admitting the non-decreasing nature of the phase-space density in the core, we postulate a local phase-space constraint that the maximum phase-space density in the relaxed system is determined by the initial conditions at each radius in the core regions for power-law density spheres. From equation (1), the local maximum phase-space density at the beginning, $`f_\mathrm{m}`$, is defined by $$f_\mathrm{m}(r)\frac{\rho _0(r)}{\sqrt{2}\pi ^{3/2}\sigma _r\sigma _{}^2}.$$ (3) If the initial density distribution is given by a power law like $`\rho _0r^n`$, the local maximum phase-space density of the initial system is represented in terms of the anisotropic parameter, $`\alpha `$, and the virial ratio, $`\eta `$, by $$f_\mathrm{m}(r)=\frac{1}{8\sqrt{2}\pi ^{5/2}}\frac{(52n)^{3/2}}{(3n)^{1/2}}\frac{(\alpha +2)^{3/2}}{\alpha ^{1/2}}\eta ^nr^n.$$ (4) Notice that equation (4) holds even though $`\alpha `$ depends on radius. The cumulative mass within $`r`$ is $`M(r)=(r/R_0)^{3n}`$. For the special case where $`\alpha `$ is independent of $`r`$, we find the cumulative mass with phase-space density greater than $`f_\mathrm{m}`$ by solving equation (4) with respect to $`r`$, $`M(f_\mathrm{m})`$ $`=`$ $`{\displaystyle \frac{1}{[8\sqrt{2}\pi ^{5/2}]^{(3n)/n}}}{\displaystyle \frac{(52n)^{3(3n)/(2n)}}{(3n)^{(3n)/(2n)}}}`$ (5) $`\times {\displaystyle \frac{(\alpha +2)^{3(3n)/(2n)}}{\alpha ^{(3n)/(2n)}}}f_\mathrm{m}^{(3n)/n}.`$ This equation shows that $`M(f_\mathrm{m})`$ is independent of $`\eta `$, because we have determined $`R_0`$ from the relation such that $`\eta \times R_0=1`$. Thus, our discussion described below holds whatever value the virial ratio will be. From equation (5), we obtain the relation between $`M`$ and $`f_\mathrm{m}`$ as $$M(f_\mathrm{m})=\frac{27}{256\pi ^5}\frac{(\alpha +2)^3}{\alpha }f_\mathrm{m}^2\mathrm{for}n=1,$$ (6) and $$M(f_\mathrm{m})=\frac{1}{2^{7/4}\pi ^{5/4}}\frac{(\alpha +2)^{3/4}}{\alpha ^{1/4}}f_\mathrm{m}^{1/2}\mathrm{for}n=2.$$ (7) Since our initial models have $`\rho _0r^1`$, equation (6) predicts the relation $`M(f_\mathrm{m})f_\mathrm{m}^2`$. The resulting $`M(f_\mathrm{m})`$ after violent relaxation cannot exceed $`M(f_\mathrm{m})`$ of the initial system. As the relaxed density distributions show $`\rho r^{2.1}`$ at intermediate radii, regardless of the values of $`\alpha `$, $`M(f_\mathrm{m})f_\mathrm{m}^{1/2}`$ would hold approximately at such radii. As illustrated in Fig. 3, for large values of $`f_\mathrm{m}`$ there unavoidably emerges a region in which $`M(f_\mathrm{m})`$ for $`\rho r^{2.1}`$ becomes larger than that for $`\rho r^1`$. Since large $`f_\mathrm{m}`$ correspond to small $`r`$ as indicated by equation (4), we find from Fig. 3 that the density profile with $`\rho r^{2.1}`$ cannot continue down to sufficiently small radii owning to the lack of phase-space density. Therefore, the density bend necessarily arises as long as the initial density distribution is shallower than $`\rho _0r^{2.1}`$. As a piece of evidence that the local phase-space constraint is valid, we present in Fig. 4 the plots of phase particles on the radial mesh points at the final state for $`\eta =0.1`$ with $`\alpha =0.2`$, 1, and 10 in the $`(f,r)`$plane. This figure reveals that phase particles with high phase-space density are not scattered in a wide range of radius. Thus, at least in the central region, mixing of high phase-space density particles with low phase-space ones does not occur efficiently through violent relaxation. If we pick out the maximum phase-space density at each radius from that distribution of the phase particles which is shown in Fig. 4, we can obtain, in a practical sense, the local maximum phase-space density, $`f_\mathrm{m}`$, in the relaxed state. Fig. 5 is depicted in such a way to show the relaxed $`f_\mathrm{m}`$ against $`r`$ near the central region. We can see from this figure that the relaxed $`f_\mathrm{m}`$ is well-approximated by $`f_\mathrm{m}r^2`$ at $`r0.1`$, although the slope of the model with $`\alpha =10`$ is closer to $`1.6`$ rather than $`2`$. The relation $`f_\mathrm{m}r^2`$ is naturally derived from equation (4) for the density profile $`\rho r^2`$ which is in fact obtained at intermediate radii in our simulations, though it is precisely $`\rho r^{2.1}`$. Therefore, our explanation for the density bend based on Fig. 3 is justified qualitatively. We now turn to the tendency that larger values of $`\alpha `$ result in more cuspy density distributions. When we begin with $`\rho _0r^1`$, equation (6) indicates that $`M(f_\mathrm{m})`$ is proportional to $`(\alpha +2)^3/\alpha `$. This function has a minimum at $`\alpha =1`$. Hence, $`M(f_\mathrm{m})`$ increases with increasing $`\alpha `$ if $`\alpha >1`$. This means that $`M(f_\mathrm{m})`$ for $`\rho r^2`$ intersects with $`M(f_\mathrm{m})`$ for $`\rho r^1`$ with $`\alpha =10`$ at a larger $`f_\mathrm{m}`$ than that with $`\alpha =1`$ (see Fig. 3). Thus, $`\rho r^{2.1}`$ can continue down to smaller radii with increasing $`\alpha `$, provided that the relaxed density profiles at intermediate radii do not change greatly among the models with different values of $`\alpha `$, as is demonstrated by our simulations. In this way, we can understand that the large $`\alpha `$ models produce nearly cuspy density profiles down to sufficiently small radii. On the other hand, $`M(f_\mathrm{m})`$ decreases with increasing $`\alpha `$ if $`\alpha <1`$. Thus, if $`\alpha <1`$, smaller values of $`\alpha `$ should also, in principle, lead to a density bend closer to the centre (see Fig. 3). In reality, this is not the case, as illustrated in Fig. 1. This can be understood from Fig. 6 which shows the fractional mass, $`dM(j)/dj`$, with the angular momentum, $`j`$, for the initial models. These fractional mass distributions remain unchanged throughout the evolution because spherical symmetry is assumed. Fig. 6 shows that the mass of stars with small values of the angular momentum decreases as $`\alpha (<1)`$ decreases for $`j0.2`$. If the $`\rho r^2`$ profile could continue down to the very centre, and if most stars would be on circular orbits, $`dM(j)/dj`$ would be equal to $`\sqrt{4\pi \rho _{}}`$, where $`\rho _{}`$ is the reference density at $`r=1`$. The values of the imagined $`dM(j)/dj`$ are 0.45 for $`\eta =0.1`$ and 0.26 for $`\eta =10^{1.5}`$. For these values of $`dM(j)/dj`$, Fig. 6 indicates that the smaller $`\alpha `$ models contain less stars with small $`j`$. Since the stars with small values of the angular momentum can pass close to the centre, the deficiency of small $`j`$ stars means that the smaller $`\alpha `$ models cannot populate a sufficient number of stars down to smaller radii. Therefore, the radius of the density bend moves outward as $`\alpha `$ decreases if $`\alpha <1`$, even though there is sufficient phase-space density available. If spherical symmetry were not imposed, some degree of mixing in angular momentum could be expected (May & van Albada 1984), and so, $`\rho r^{2.1}`$ might continue down to smaller radii as $`\alpha `$ decreases when $`\alpha <1`$. We can infer from Figs. 1 and 6 that the models with $`\alpha >1`$ would not suffer the limitation, described above, on the central density distribution arising from the amount of small angular momentum stars. This is because we can explain the shift of the density bend toward smaller radii with increasing $`\alpha `$ simply from the local phase-space constraint like that illustrated in Fig. 3. We have found that the velocity anisotropy could be effective in the formation of cuspy density profiles. However, it is unclear how the system can acquire such a large radial anisotropic velocity dispersion. It may be difficult to realize a large velocity anisotropy in each stellar system before collapse. In a cosmological situation based on cold dark matter, clumps are first formed and then collapse to merge into a large system. Thus, a system consisting of clumps may have anisotropic velocity dispersions. Therefore, it will be important to study the velocity distribution just after such a system has decoupled from the Hubble flow. ### 4.2 Relation between initial and final density powers We have found that more radially anisotropic velocity distributions lead to more cuspy density profiles by gravitational collapse. Here, we examine how seriously the initial power index of density distributions influences the final one. We carry out collapse simulations of density distributions $`\rho r^n`$ with $`n`$=0.5, 1, and 2, again on the assumption of spherical symmetry using the collisionless Boltzmann code. The anisotropic parameter and virial ratio are chosen to be $`\alpha =10`$ and $`\eta =0.1`$, respectively. The other numerical parameters are the same as those used in Section 2. The final density profiles are shown in Fig. 7. The half-mass radii, $`r_\mathrm{h}`$, are about 2.3 to 3.2. We can see that the density distributions at intermediate radii are well-approximated by $`\rho r^{2.1}`$, and that initially steeper density distributions result in steeper cuspy density distributions, although the final power-law indices within the radius of the density bends become always shallower than the initial ones. In particular, the density power for the model with $`\rho r^2`$ remains almost unchanged after the collapse, so that practically no density bend is found. This fact is in good agreement with the results of Burkert (1990), and proves the validity of the local phase-space constraint that we have postulated in the previous subsection, because the relaxed density distribution with $`\rho r^{2.1}`$ can almost continue toward the very centre without the phase-space constraint if the initial density distribution is $`\rho r^2`$, as indicated by Fig. 3. Taking into consideration our results obtained here, the steep cusps with $`\rho r^2`$ observed in faint ellipticals could originate from the fluctuation spectrum that realizes density distributions with $`\rho r^2`$. According to observations, low-luminosity ellipticals that have the steep cusps are isotropic in the sense that they show $`(V/\sigma )^{}1`$, where $`(V/\sigma )^{}`$ is the ratio of the rotation parameter $`V/\sigma `$ to the value for an isotropic oblate spheroid flattened by rotation, $`V`$ is the maximum rotation velocity, and $`\sigma `$ is the mean velocity dispersion inside one-half of the effective radius (Kormendy & Bender 1996; Faber et al. 1997). In addition, these ellipticals are rapidly rotating. Unfortunately, we cannot discuss $`(V/\sigma )^{}`$ values for our end-products, because our models include no net rotation. Therefore, we will need to investigate the effects of rotation on the steep cusps using three-dimensional collapses in order to make a detailed comparison with observations. The resulting power-law indices of the density distributions within the radius of the density bend correlate with those of the initial density distributions. The difference in power-law index of the cusps originates from a difference in that of initial density distributions. In cosmological simulations with high resolution, based on a standard cold dark matter scenario, the final density distributions show a rather steep density cusp, such as $`\rho r^{1.4}`$ (Fukushige & Makino 1997; Moore et al. 1998). Probably, this kind of violent relaxation would not produce shallow cusps observed in bright ellipticals, so that we might require a merger between galaxies each of which contains a massive central black hole (Makino & Ebisuzaki 1996; Nakano & Makino 1999). However, if the circumstances for the formation of bright ellipticals are somehow very different from those of faint ellipticals, it is conceivable from our results that bright ellipticals could have been formed from the fluctuation spectrum that corresponds to $`\rho r^n`$ with $`0.5n1`$ approximately. ## 5 Conclusions Collapse simulations of power-law density spheres such as $`\rho r^1`$ with anisotropic velocity distributions have been performed on the assumption of spherical symmetry. We have found that the resulting density profiles are well-approximated by $`\rho r^{2.1}`$ at intermediate radii, regardless of the velocity anisotropy and the virial ratio. However, this universal density profile cannot continue down to sufficiently small radii, and so, the density bends necessarily. We have explained this density bend from the $`local`$ phase-space constraint that the phase-space density accessible to the relaxed system is determined at each radius by the maximum phase-space density of the initial system. As the radial velocity anisotropy increases, the density profiles of the end-products after violent relaxation become more cuspy. In particular, the extremely radially anisotropic models with $`\alpha =10`$ have practically no flat cores. Thus, our results imply that the velocity anisotropy could play a substantial role in the formation of the density cusps observed in elliptical galaxies as long as we rely on a simple dissipationless collapse picture. In reality, it will be difficult to produce a large velocity anisotropy in individual systems. Concerning this problem, however, we point out that a system composed of clumps, as can be envisaged in the standard cold dark matter scenario, might have a large velocity anisotropy when it has decoupled from the Hubble expansion. In addition, we have examined the relation between the initial and final power-law indices of the density profiles by performing collapse simulations for $`\rho r^n`$ models with $`n`$=0.5, 1, and 2. In these cases, the resulting density profiles at intermediate radii are also well-approximated by $`\rho r^{2.1}`$, regardless of the initial density power indices. The results show that initially steeper density profiles result in steeper cuspy density profiles. Therefore, the difference in the power index of the cusps between faint and bright elliptical galaxies might arise from the difference in the power index of initial density profiles. In particular, the steep cusps with $`\rho r^2`$ observed in faint ellipticals could result from initial density distributions with $`\rho r^2`$, because such density distributions can avoid the local phase-space constraint. As a by-product, we have also found that the resulting cores no longer show velocity anisotropy even though we start with a model which has a large velocity anisotropy at all radii. This implies that simple gravitational collapse could have difficulty in generating and maintaining velocity anisotropy in the cores. Our results shown here are obtained on the assumption of spherical symmetry. If aspherical collapses are allowed, the radial orbit instability that arises from small initial virial ratios might affect the density distributions of end-products. The study of three-dimensional collapses is in progress. ## Acknowledgments We are grateful to Drs. J. Makino and T. Fukushige for enlightening discussion. We are also grateful to the anonymous referee for valuable comments on the manuscript. SH thanks Max-Planck-Institut für Astronomie in Heidelberg for its hospitality during this research. This work was supported in part by the Grant-in-Aid for Scientific Research from the Ministry of Education, Science, Sports and Culture of Japan (09740172).
no-problem/9906/astro-ph9906167.html
ar5iv
text
# Detection of the 62 𝜇m crystalline H2O ice feature in emission toward HH 7 with ISO-LWS0footnote 00footnote 0 ## 1. Introduction Herbig-Haro objects (HH; Haro H50 (1950), Herbig H51 (1951)) are emission-line objects acting as signposts for the shock regions (Draine, Roberge & Dalgarno DRD83 (1983); Hollenbach & McKee HM89 (1989)) originating at the interface between stellar winds accelerated by Young Stellar Objects (YSOs) and the circumstellar or cloud ambient material. The temperature of dust grains in these shock regions can raise to only a few hundred degrees at most in dissociative shocks (Hollenbach & McKee HM89 (1989)), so that their thermal emission cannot be detected below 10 $`\mu `$m. The instruments on board the Infrared Space Observatory satellite (ISO, Kessler et al. Ketal96 (1996)) opened unprecedented possibilities for far infrared continuum studies of cold objects, including HH objects. In this Letter we present the data obtained with the Long (LWS, Clegg et al. Clegg96 (1996)) and Short (SWS, de Graauw et al. dGetal96 (1996)) Wavelength Spectrometer towards HH 7, the leading bow-shaped shock of the HH 7-11 chain emanating from the YSO SVS 13, in the star forming region NGC 1333 in Perseus (d=350 pc). Details about the observations and data reduction are given elsewhere (Molinari et al. Metal99 (1999)). In Sect. 2 the additional data analysis procedures we adopted to derive a reliable continuum spectrum for HH 7 are discussed. Nomenclature for the ten LWS detectors is described in the ISO Data User Manual (http://www.iso.vilspa.esa.es/manuals/lws\_idum5; detectors SW1 to SW5 (43-90 $`\mu `$m), and LW1 to LW5 (80-197 $`\mu `$m), are sometimes referred to as “short” and “long” wavelength detectors respectively. ## 2. Results The LWS beam centered on HH 7, also includes object HH 8 somewhat 20<sup>′′</sup> off-axis and HH 10 at the edge of the beam; these objects are however fainter than HH 7 at 2 $`\mu `$m (Molinari et al. Metal99 (1999)) and the beam profile suppresses their possible contribution even more. Apart from a contamination by the strong nearby source SVS 13, which will be discussed in detail in Sect. 2.2, it is plausible to assume that HH 7 dominates the observed spectrum. ### 2.1. The 62 $`\mu `$m feature In Fig. 2.1 the complete spectra observed towards HH 7 and SVS 13 are shown. A broad feature extending from roughly 50 to 70 $`\mu `$m is clearly visible in the HH 7 spectrum. For clarity the SW3 detector spectrum, with an offset applied to align it with the continuum of the adjacent detectors, is shown with plus symbols. The primary concern was to make sure that the 50-70$`\mu `$m feature was not a result of a residual instrumental effect and that it is intrinsic to HH 7. The passband calibration is known to be somewhat inaccurate for detector SW1 (it may be worse than 50% both in absolute and relative terms) which is the most sensitive to transients. Detectors SW2, SW3 and SW4 however, are stable in relative (passband) terms and a conservative figure of 30% can be assumed for their absolute calibration accuracy. The possibility that the broad feature visible on Fig. 2.1 may be due to the near-IR leaks present in the LWS detectors filters (http://isowww.estec.esa.nl/notes/lws\_0197.html) can also be excluded. Finally, this feature might result from contamination effects from the bright nearby source SVS 13, which is the candidate exciting source for HH 7; these effects will be discussed in detail in the Sect. 2.2. Fig. 1 – LWS averaged spectra observed towards HH 7 (bottom) and SVS 13 (top). The different portions are the 10 LWS detectors, labeled with the names used in the text. The plus symbols for HH 7 represent detector SW3 with an applied offset of 5 Jy to bring it in line with the adjacent detectors. Here we just note here that if the 50-70 $`\mu `$m feature seen on the HH 7 spectrum was due to a fraction of the flux emitted by the nearby contaminating source SVS 13, we would expect to see the same feature at a comparable “line-to-continuum” ratio on the SVS 13 continuum spectrum; Fig. 2.1 clearly shows that this is not the case. We conclude that the 50-70 $`\mu `$m feature is real and intrinsic to HH 7 and we consider it as a 2$`\sigma `$ detection, assuming that the noise of the feature is equal to half the gap between detectors SW3 and SW2. We identify this as the 62$`\mu `$m feature due to the longitudinal acoustic modes of crystalline water ice (Bertie, Labbe & Whalley BLW69 (1969)), observed for the first time by Omont et al. (Oetal90 (1990)) in the expanding envelopes of post-AGB stars; ISO-LWS has detected this feature in the spectra of similar objects (Barlow B98 (1998)) and of a few Herbig Ae/Be stars (Waters & Waelkens WW98 (1998), Malfait et al. Malfait (1999)). It is the first time this feature has been detected towards Herbig-Haro objects. ### 2.2. The 2-200 $`\mu `$m Continuum Several SWS line scans were used to estimate the continuum at different wavelengths for $`\lambda <`$40 $`\mu `$m. The line-free portions of each individual spectrum were used to build flux histograms, and a gaussian was then fitted to the core of the distribution to obtain the centroid (average flux) and the standard deviation; the latter was then divided by the square root of the number of points in histogram to get an estimate of the uncertainty. These uncertainties reflect the internal accuracy of the estimates; infact the true uncertainty may be higher. In particular, all data shortward of $``$ 10 $`\mu `$m are at the detection limit of the SWS and they will be treated as 1$`\sigma `$ upper limits. LWS scans were averaged using the ISO Spectral Analysis Package (ISAP, http://www.ipac.caltech.edu/iso/isap/isap.html) using a median clipping algorithm which rejected outlying data points due to incomplete removal of glitches from cosmic rays hits. However, contamination effects in the HH 7 spectrum due to the nearby ($``$ 70<sup>′′</sup> NW) candidate exciting source SVS 13 must be taken into account to assess if the observed continuum (Fig. 2.1 is all due to HH 7; for this purpose we used an irregularly sampled raster map of Mars taken as part of the LWS calibration programme. We first used the LIA (http://www.ipac.caltech.edu/iso/lia/lia.html) routine INSPECT\_RASTER to determine the relative position of HH 7 and SVS 13 in the \[Y,Z\] spacecraft frame of reference. For each Mars raster position along the direction connecting the two sources, we averaged the Mars spectra detector by detector; each detector average at the various off-axis positions was then ratioed to the analogue detector average of the on-axis spectra, and a set of ten contamination factors for each position was obtained. A spline interpolation was then used to estimate these factors (one per detector) at an off-axis distance of 68<sup>′′</sup>, the distance between HH 7 and SVS 13; we find values 0.024, 0.025, 0.037, 0.05, 0.075, 0.09, 0.10, 0.11, 0.13 and 0.15, which increase with wavelength as expected due to diffraction. We multiplied the observed SVS 13 spectra by these numbers, and the resultant spectrum was subtracted from the observed HH 7 spectrum. This method for estimating the contamination from nearby sources inevitably suffers from the irregular sampling of the Mars raster map. Raster points did not lay exactly along the HH 7-SVS 13 direction on the focal plane, and the interpolation between the correction factors estimated at each relevant raster position introduced an additional uncertainty. At the end of the procedure (see Fig. 3) we find worse alignment between adjacent LW detectors, and detector SW5 goes to negative flux values (not reported on Fig. 3). We believe this may be due to the fact that as the magnitude of the applied contamination correction increases with wavelength, so it does the associated uncertainty; in particular, detector SW5 is in the critical region where the contamination correction starts to be high (2-3 times higher than for SW2 and SW4) while the observed signal is still low (similar to SW2), so that a slightly overestimate of the contamination fraction is enough to bring its corrected values below zero. The LWS LW detectors are those for which the contamination fractions are higher and for which diffraction and source’s extension effects (which here have been neglected) are more severe. An additional complication, which does not affect the SW detectors, is that heavy fringing is observed (which has been here removed using standard tools available in ISAP); this has the effect of modulating the beam size as a function of wavelength even within individual detector bands. We did not take this into account, instead deriving a single contamination factor per detector; multiplying or dividing by a constant, however, has the effect of changing the slope. Hence, although the observed continuum levels are high enough to ensure that an important part of the observed signal can confidently be assigned to HH 7, the exact absolute fluxes and spectral shape in this wavelength range remain highly uncertain. On the short wavelength side of the LWS spectrum on the other hand, the existence of intrinsic continuum emission from HH 7 depends more critically on the particular value of the contamination factors; doubling this factors would lower the spectrum to negative flux levels. If, however, the $`\lambda `$ 80 $`\mu `$m flux observed towards HH 7 were due to contamination, we would expect to see the 62 $`\mu `$m emission feature also on the continuum spectrum of SVS 13 and with a comparable “line/continuum”; instead, no trace of such a feature is seen on the SVS 13 continuum (see Fig. 2.1). We conclude that FIR emission which is intrinsic to HH 7 has been detected, showing for the first time that Herbig-Haro objects cease to be exclusively emission-line objects at FIR wavelengths. ## 3. Discussion In the following discussion we will assume that the FIR and mm continuum arise from the same region of space, based on the spatial coincidence between the HH objects and the mm emission distribution; currently available 50-100 $`\mu `$m data on the region (Harvey et al. Hetal84 (1984), Hetal98 (1998); Jennings et al. Jetal87 (1987)) however, do not have enough spatial resolution or sensitivity to support this claim. The spectral energy distribution (SED) from HH 7 will be modeled as thermal emission from dust grains composed of a silicate core and a water ice mantle. Absorptivities between 2 and 300 $`\mu `$m were computed with Mie theory in the formulation of Wickramasinghe (W67 (1967)), using the complex refractive indices for crystalline water ice (Bertie et al. BLW69 (1969)) and silicate (Draine D85 (1985)). Longward of 300 $`\mu `$m the silicate absorptivities by Draine (D85 (1985)) were adopted; inclusion of water ice mantles may steepen the slope of the Q<sub>abs</sub> vs $`\lambda `$ relationship (Aannestad A75 (1975)), possibly leading to underestimation of the dust mass. Radiative transfer is approximated with an analytical treatment where dust is distributed on a sphere which is characterised by radial density and temperature gradients (Noriega-Crespo, Garnavich & Molinari NCGM98 (1998)). The presence of a cold dust clump centered on the location of HH 7 and extending over a larger area than the one traced by the optical or near-IR emission (Lefloch et al. Letal98 (1998)), justifies a treatment where a dust clump is centrally heated by the HH 7 shock. Fig. 3 presents the complete SED towards HH 7, after removal of the contamination from SVS 13 (see Sect. 2.2). The LWS detectors SW1 to SW4 have been rescaled to a common level preserving their original mean value; no other rescaling was made to the rest of the LWS spectra. The continuum points from the SWS spectra and an estimate for the 1.25 mm flux integrated on the model fit area (see below) are also reported; the latter point has to be considered an upper limit since it is contaminated by SVS 13 just in the same way it happens for the LWS observations. We find that it is impossible to globally fit the observed SED; considering the discussion in Sect. 2.2, we then decided to give higher priority to the short wavelength LWS detectors spectra also because these are the ones showing the 62 $`\mu `$m feature, and more closely match each other (not considering the applied rescaling). While the overall appearance of the continuum depends on the radius of the dust clump, its mass and temperature gradient, the “line/continuum” ratio of the 62 $`\mu `$m feature depends on the relative size of the ice mantle with respect to the grain core. The 62 $`\mu `$m water ice feature is well fitted by adopting a core radius of 0.07 $`\mu `$m and a total core + mantle radius of 0.1 $`\mu `$m. The model also suggests that the short wavelength end of the LWS spectrum can be identified with one side of the 45 $`\mu `$m water ice feature (see also Dartois et al. Detal98 (1998)), also in emission; although detector SW1 is the one that experiences the most severe problems in calibration-related aspects, the further agreement of the model with the $`\lambda `$38.4 $`\mu `$m SWS point makes this a plausible possibility. The importance of the detection of the 62 $`\mu `$m feature is that it is specific to crystalline water ice, as opposed to the 45 $`\mu `$m feature which is instead predicted for both crystalline and amorphous ice (e.g. Moore et al. Metal94 (1994)). Most of the laboratory studies (e.g., Smith et al. Setal94 (1994) and references therein) show that ice deposited onto grains at T $``$$`<`$ 100 K is in amorphous state (but see Moore et al. Metal94 (1994) who instead obtain crystalline ice in these conditions) also for deposition rates as high as 20 $`\mu `$m h<sup>-1</sup> (Léger et al. Letal83 (1983)); the 62$`\mu `$m feature is then an important indicator of the thermal history of the grains because it appears at T $``$$`>`$ 100 K when an irreversible transition from amorphous to crystalline state commences. Fig. 2 – Complete spectral energy distribution towards HH 7. The full lines represent the LWS spectra after applying the contamination corrections discussed in the text. Detectors SW1 to SW4 have been rescaled to a common level preserving their original mean value. The fit plotted in Fig. 3 is obtained adopting a dust clump radius of 0.06 pc, which approximates the radius of the average LWS beam size at a distance of 350 pc, and the size of the 1.25 mm continuum emission area (Lefloch et al. Letal98 (1998)). The density is assumed constant and equal to 6$`\times 10^4`$ cm<sup>-3</sup> while the temperature varies from $``$10 to 200 K with a $``$0.4 power-law radial gradient. The bolometric luminosity obtained integrating the model fit is L<sub>fit</sub>= 3.7 L, while integration of the contamination-corrected SED (Sect. 2.2) yields L$`{}_{sed}{}^{}<\mathrm{\hspace{0.33em}4.5}`$ L (since $`\lambda <10`$ $`\mu `$m and $`\lambda `$= 1.25 mm data are to be considered upper limits); this is about a factor 30 higher than the cooling via atomic and molecular lines observed towards HH 7 (Molinari et al. Metal99 (1999)). The fitted model predicts that the bulk of the 62 $`\mu `$m feature is emitted by dust at temperatures T $``$$`>`$ 30 K concentrated inside a 4<sup>′′</sup>-radius region centered on HH 7, a size comparable to that of the optical and near-IR emission which traces the shock front. Since these grains should have experienced a rise in temperature to values $``$$`>`$ 100 K in order for ice mantles to be in crystalline state, it is plausible that the 62 $`\mu `$m feature originates from dust which has been processed by the HH 7 shock. The dust mass in the 4<sup>′′</sup>-radius region centered on HH 7 amounts to 5$`\times 10^5`$ M (the total dust mass implied by the model fit is $``$ 0.035 M); the relative proportion of core and mantle (the core has 70% of the total grain radius) implies a water ice mass of $`2\times 10^5`$ M, or a H<sub>2</sub>O column density $`1.1\times 10^{18}`$ cm<sup>-2</sup>. H<sub>2</sub> pure rotational lines (Molinari et al. Metal99 (1999)) suggest N(H<sub>2</sub>) $`4.4\times 10^{20}`$ cm<sup>-2</sup> in the same 4<sup>′′</sup>-radius region, implying \[H<sub>2</sub>O\]/\[H\]$`1.25\times 10^3`$, or a factor $``$ 4 higher than the interstellar O gas-phase abundance (Meyer et al. MJC98 (1998)). However, this number should be regarded as an upper limit because our model assumes that all dust grains are coated with ice mantles, which is not necessarily true; besides, we cannot exclude the presence of cold H<sub>2</sub> (T $``$$`<`$ 100 K) which our ISO observations (Molinari et al. Metal99 (1999)) would not trace. This water abundance, even if considered only as an order-of-magnitude estimate, is however much higher than the gas phase water abundance (\[H<sub>2</sub>O\]/\[H\]$`<10^5`$) deduced from FIR lines (Molinari et al. Metal99 (1999)), and it would essentially imply that most of the oxygen is locked into water ice. Due to the uncertainty about the 45 $`\mu `$m feature (see above), it hard to tell from the observational viewpoint whether gas-phase water produced behind the HH 7 shock front (Kaufman & Neufeld KN96 (1996)) is deposited onto bare warm grains, or pre-existing ice mantles are warmed-up during the passage of a relatively gentle shock front. The ice optical constants that we used (Bertie et al. BLW69 (1969)) are from laboratory samples obtained by direct deposition at T=173 K and subsequent cooling to 100 K; hence the good simultaneous fit of the 62 and 45$`\mu `$m features would tend to support the first scenario. The physical condition behind low velocity (v$`{}_{s}{}^{}<`$40 km s<sup>-1</sup>), non-dissociative, shocks are favourable for the rapid gas-phase incorporation of atomic oxygen into water (Draine, Roberge & Dalgarno DRD83 (1983), Kaufman & Neufeld KN96 (1996)). Subsequent freezing onto grains (Bergin, Neufeld & Melnick BNM98 (1998), BNM99 (1999)) in the cooling post-shock region could produce the observed water ice mantles, also explaining the minor role played by gas-phase water in the cooling of the HH 7 shock (Molinari et al. Metal99 (1999)). However, non-dissociative shocks are unable to raise the grain temperature to the T $``$$`>`$ 100 K (Draine, Roberge & Dalgarno DRD83 (1983)) needed to explain the presence of crystalline ice mantles; a dissociative component, whose generated intense UV field (Hollenbach & McKee HM89 (1989)) is much more efficient in heating the grains, is needed. The co-existence of both types of shock is indeed an expected feature of bow shocks like HH 7 (Smith & Brand SB90 (1990)) and it is independently supported by FIR lines studies (Molinari et al. Metal99 (1999)). However, we point out that this relies on the general result that crystalline ice requires relatively high temperatures to form; if crystalline ice can also form at low (T$`<`$ 20 K) temperature (Moore et al. Metal94 (1994)) then a dissociative shock component may not be needed. The alternative possibility that pre-existing amorphous water ice mantles are heated up during the passage of the shock front seems unlikely since the mantles are easily destroyed by grain-grain collisions once the shock velocities exceed $``$ 15 km s<sup>-1</sup> (Caselli, Hartquist & Havnes CHH97 (1997)). We can derive an order-of-magnitude estimate for the timescale $`\tau _{ice}`$ of the formation of water ice mantles, dividing the linear size of the post-shock region by the shock velocity; assuming a maximum shock velocity of $``$ 40 km s<sup>-1</sup> for the HH 7 shock (Solf & Böhm SB87 (1987), Molinari et al. Metal99 (1999)) and a linear extent of $``$ 10<sup>′′</sup> for the HH 7 post-shock region as estimated from H<sub>2</sub> 2.12$`\mu `$m images (Garden et al. Getal90 (1990), Everett E97 (1997), Molinari et al Metal99 (1999)), we estimate $`\tau _{ice}`$ 400 yrs, a factor $``$ 250 less compared to theoretical predictions for ice mantles formation behind non-dissociative shocks (Bergin, Neufeld & Melnick BNM99 (1999)). This very short timescale for ice mantle formation would make the gas-phase water cooling practically irrelevant as a shock diagnostic. Acknowledgements: We thank B. Swinyard for providing us with the calibrated data for the Mars raster map we used to estimate the contamination corrections, and B. Lefloch for providing us with the integrated flux from his published 1.25 mm map of NGC 1333. We also thank A. Noriega-Crespo and E. Sturm for sharing their insight in SWS data. The ISO Spectral Analysis Package (ISAP) is a joint development by the LWS and SWS Instrument Teams and Data Centers. Contributing institutes are CESR, IAS, IPAC, MPE, RAL and SRON.
no-problem/9906/cond-mat9906118.html
ar5iv
text
# Toward a Theory of Orbiton Dispersion in LaMnO3 ## I introduction The fascinating properties of Manganite perovskites have recently been rediscovered . LaMnO<sub>3</sub> has Mn<sup>3+</sup> ions with electronic configuration $`3d^4`$ in the high-spin $`t_{2g}^3e_g^1`$ state. The single electron in the doublet $`e_g`$ level is Jahn-Teller (JT) active. A cubic to orthorhombic distortion occurs at $`T_{\mathrm{JT}}`$=750K which lifts the degeneracy with alternating sign in alternating cells, and drives orbital ordering : $`x`$ and $`y`$-oriented $`e_g`$ orbitals alternate in the $`xy`$ plane. This cooperative JT transition yields a layer structure which then determines at the much lower Neel temperature ($`T_\mathrm{N}`$=240K) an antiferromagnetic “A-type” (AFA) spin-ordering with ferromagnetic spin order in the $`xy`$ planes. We believe that the principal term in the Hamiltonian which drives the cooperative JT transition is coupling of Mn orbitals to motions of oxygen atoms along the directions of the (approximately 180) Mn-O-Mn bonds. In a previous paper we have discussed a model for the JT transition and used this model both to describe the $`T=0`$ phase of pure LaMnO<sub>3</sub> and the formation of small polarons when hole-doped as in La<sub>1-x</sub>Ca<sub>x</sub>MnO<sub>3</sub>. Our model is related to the Rice-Sneddon model for BaBiO<sub>3</sub>, and was introduced for LaMnO<sub>3</sub> by Millis . Our calculations are done at $`T`$=0 and $`x`$=0. Previous authors have studied orbital dynamics using models which omit JT and electron-phonon effects, focussing on dynamics induced by hopping. This leads to complicated orbital dynamics strongly dependent on spin order because of Hubbard and Hund energies. Our model emphasizes a very different kind of physics. Because we take $`U=\mathrm{}`$ and $`x`$=0, there is no hopping and the spins are decoupled from other degrees of freedom. The remaining electron-phonon physics is still very rich, and we think it will be dominant at small $`x`$. Other treatments of the small $`x`$ orbital dynamics exist (see ref. and papers cited there), but the large ratio $`T_{\mathrm{JT}}/T_\mathrm{N}`$ argues against omitting the electron-phonon physics at small $`x`$. ## II model and ground state Each cubic cell has 5 degrees of freedom, two electronic and 3 lattice dynamical. The electronic degrees of freedom are occupancies of the Mn $`e_g`$ orbitals. In Van Vleck notation $`\psi _2`$ and $`\psi _3`$ are the familiar $`x^2y^2`$ and $`3z^2r^2`$ functions. These can be rotated by 45 to an alternative orthogonal basis, the symmetrical pair $`\psi _X`$, $`\psi _Y=(\psi _3\pm \psi _2)/\sqrt{2}`$. These orbitals point in $`x`$ and $`y`$ directions respectively, with a moderately strong lobe in the $`z`$ direction. Lattice dynamical degrees of freedom are labeled by $`u(\mathrm{}\pm x)`$, $`u(\mathrm{}\pm y)`$, and $`u(\mathrm{}\pm z)`$, which refer to oxygen displacements along the $`x`$, $`y`$, and $`z`$ directions of the oxygen atoms half-way between the Mn atom at site $`\mathrm{}`$ and the next atom in the $`\pm x`$, $`\pm y`$, and $`\pm z`$ directions, respectively. The notation $`u(\mathrm{}x)`$ is equivalent to $`u(\mathrm{}^{}+x)`$ for the Mn site $`\mathrm{}^{}`$, which is translated from $`\mathrm{}`$ by one unit in the $`x`$-direction. It is convenient to use both the Cartesian labeling given above, and a Van Vleck labeling: $`\sqrt{3/2}Q_1(\mathrm{})`$ $`=`$ $`[u(\mathrm{}+x)u(\mathrm{}x)]+[u(\mathrm{}+y)u(\mathrm{}y)]+[u(\mathrm{}+z)u(\mathrm{}z)]`$ (1) $`Q_2(\mathrm{})`$ $`=`$ $`[u(\mathrm{}+x)u(\mathrm{}x)][u(\mathrm{}+y)u(\mathrm{}y)]`$ (2) $`\sqrt{3}Q_3(\mathrm{})`$ $`=`$ $`2[u(\mathrm{}+z)u(\mathrm{}z)][u(\mathrm{}+x)u(\mathrm{}x)][u(\mathrm{}+y)u(\mathrm{}y)]`$ (3) Our starting Hamiltonian $``$ has a lattice term $`_\mathrm{L}`$ and an electron-phonon term $`_{\mathrm{ep}}`$. The lattice term does not have a useful expression in the Van Vleck notation, and is best written as $$_\mathrm{L}=\underset{\mathrm{},\lambda =x,y,z}{}\left(\frac{K}{2}u^2(\mathrm{}+\lambda )+\frac{1}{2M}p^2(\mathrm{}+\lambda )\right).$$ (4) On the other hand, the electron-phonon term is best written in Van Vleck operators, $`_{\mathrm{ep}}`$ $`=`$ $`g{\displaystyle \underset{\mathrm{}}{}}(c_X^{}(\mathrm{}),c_Y^{}(\mathrm{}))\left(\begin{array}{cc}\hfill Q_2(\mathrm{})& \hfill Q_3(\mathrm{})\\ \hfill Q_3(\mathrm{})& \hfill Q_2(\mathrm{})\end{array}\right)\left(\begin{array}{c}c_X(\mathrm{})\\ c_Y(\mathrm{})\end{array}\right)`$ (10) $`\sqrt{2}g{\displaystyle \underset{\mathrm{}}{}}Q_1(\mathrm{})\left(c_X^{}(\mathrm{})c_X(\mathrm{})+c_Y^{}(\mathrm{})c_Y(\mathrm{})\right).`$ In the undoped case, each Mn site is singly occupied; we stay in the subspace where the operator relation $$c_X^{}(\mathrm{})c_X(\mathrm{})+c_Y^{}(\mathrm{})c_Y(\mathrm{})=1$$ (11) holds. The degenerate $`e_g`$ orbitals are defined to have zero energy. In adiabatic approximation, the ground state of this Hamiltonian has a doubled unit cell and a net average oxygen displacement $`(Q_2(\mathrm{}),Q_3(\mathrm{}))`$ of the type $`Q(\mathrm{cos}\theta ,\mathrm{sin}\theta )\mathrm{exp}(i\stackrel{}{q}\stackrel{}{\mathrm{}})`$. The wave-vector $`\stackrel{}{q}=(\pi ,\pi ,\pi )`$ is favored, but the angle $`\theta `$ in the $`(Q_2,Q_3)`$-plane is undetermined. The observed distortion, of the $`Q_2`$-type ($`\theta =0`$), is fixed by terms left out of our Hamiltonian. Once a $`Q_2`$-type order parameter is established, the wave-vector $`\stackrel{}{q}`$ can equally well be $`(\pi ,\pi ,0)`$, and this is what is observed. The lattice is now bi-partite. We label sites by A when $`\mathrm{exp}(i\stackrel{}{q}\stackrel{}{\mathrm{}})`$ is 1, and by B when it is -1. According to Eqn. 10, the A sites will be occupied by X orbitals and the B sites by Y orbitals. The alternating occupancy of these orbitals on A and B sublattices is shown schematically in Fig. (1). Oxygen atoms move off-center by $`\pm 2g/K`$, giving a JT distortion $`Q_2=8g/K\mathrm{exp}(i\stackrel{}{q}\stackrel{}{\mathrm{}})`$. This lowers the energy of each occupied orbital by $`8g^2/K=\mathrm{\Delta }`$ and raises the empty orbitals by the same amount. The distortion raises the lattice energy by $`4Ng^2/K`$, which is half the energy lowering of $`8Ng^2/K`$ from Jahn-Teller energy. The lowest electronic excited states (“orbitons”) are misoriented orbitals which cost $`16g^2/K=2\mathrm{\Delta }`$2eV. ## III orbiton Hamiltonian We are interested in the lowest excited states of the model. We allow vibrations of oxygen atoms around their distorted positions. Therefore we are no longer making the adiabatic approximation. Lattice vibrations will renormalize both the ground state and the orbiton excitations. By flipping an orbital we create an orbital defect. We can define a creation operator $`a^{}(\mathrm{})`$ for the orbital defect on the A site as $`c_Y^{}(\mathrm{})c_X(\mathrm{})`$, and similarly on the B site $`b^{}(\mathrm{})=c_X^{}(\mathrm{})c_Y(\mathrm{})`$. One can show that these operators obey Fermi commutation relations in the subspace of singly occupied sites. Furthermore one can show that on the A sublattice number the operator of the misoriented orbitals $`c_Y^{}(\mathrm{})c_Y(\mathrm{})`$ is the same as $`a^{}(\mathrm{})a(\mathrm{})`$; similarly on the B sublattice $`c_X^{}(\mathrm{})c_X(\mathrm{})`$ is the same as $`b^{}(\mathrm{})b(\mathrm{})`$. Using these definitions we can rewrite the Hamiltonian Eqns. 4, 10 in terms of the lattice fluctuations $`u(\mathrm{}+\lambda )`$ defined as $`u_0(\mathrm{}+\lambda )+\delta u(\mathrm{}+\lambda )`$, where $`u_0(\mathrm{}+\lambda )=\pm 2g/K`$ in $`x`$ or $`y`$ direction and $`0`$ in the $`z`$ direction. $`4Ng^2/K+_{\mathrm{ep}}+_\mathrm{L}=`$ $`2`$ $`\mathrm{\Delta }\left({\displaystyle \underset{\mathrm{}A}{}}a^{}(\mathrm{})a(\mathrm{})+{\displaystyle \underset{\mathrm{}B}{}}b^{}(\mathrm{})b(\mathrm{})\right)`$ (12) $`+`$ $`{\displaystyle \underset{\mathrm{},\lambda =x,y,z}{}}\left({\displaystyle \frac{K}{2}}\delta u^2(\mathrm{}+\lambda )+{\displaystyle \frac{1}{2M}}p^2(\mathrm{}+\lambda )\right)`$ (13) $`+`$ $`2g\left({\displaystyle \underset{\mathrm{}A}{}}\delta Q_2(\mathrm{})a^{}(\mathrm{})a(\mathrm{}){\displaystyle \underset{\mathrm{}B}{}}\delta Q_2(\mathrm{})b^{}(\mathrm{})b(\mathrm{})\right)`$ (14) $`+`$ $`g\left({\displaystyle \underset{\mathrm{}A}{}}\delta Q_3(\mathrm{})\left(a^{}(\mathrm{})+a(\mathrm{})\right)+{\displaystyle \underset{\mathrm{}B}{}}\delta Q_3(\mathrm{})\left(b^{}(\mathrm{})+b(\mathrm{})\right)\right)`$ (15) The next step is into Fourier space. Define Fourier variables by $`a(\mathrm{})`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{N}}}{\displaystyle \underset{k}{}}a_ke^{i\stackrel{}{k}\stackrel{}{\mathrm{}}}`$ (16) $`b(\mathrm{})`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{N}}}{\displaystyle \underset{k}{}}b_ke^{i\stackrel{}{k}\stackrel{}{\mathrm{}}}`$ (17) $`\delta u(\mathrm{}+\lambda )`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{N}}}{\displaystyle \underset{k}{}}\sqrt{{\displaystyle \frac{\mathrm{}}{2M\omega }}}\delta u_{k_\lambda }e^{i\stackrel{}{k}\stackrel{}{\mathrm{}}}`$ (18) where $`\delta u_{k_\lambda }=f_{k_\lambda }+f_{k_\lambda }^{}`$, if $`\mathrm{}`$ is on the A sublattice, and otherwise $`\delta u_{k_\lambda }=g_{k_\lambda }+g_{k_\lambda }^{}`$. The Fourier-transformed Hamiltonian has the form $$=_{\mathrm{orb}}^0+_{\mathrm{ph}}^0+_3(\delta Q_3)+_2(\delta Q_2).$$ (19) The first two terms are unperturbed dispersionless orbitons and phonons with energy $`2\mathrm{\Delta }`$ and $`\mathrm{}\omega `$. $`_{\mathrm{orb}}^0`$ $`=`$ $`2\mathrm{\Delta }{\displaystyle \underset{k}{}}\left(a_k^{}a_k+b_k^{}b_k\right)`$ (20) $`_{\mathrm{ph}}^0`$ $`=`$ $`\mathrm{}\omega {\displaystyle \underset{k,\lambda =x,y,z}{}}\left(f_{k_\lambda }^{}f_{k_\lambda }+g_{k_\lambda }^{}g_{k_\lambda }\right)`$ (21) The last two terms give the coupling of orbiton density to quantum fluctuations of the lattice of the $`\delta Q_3`$-type and $`\delta Q_2`$-type. $`_3(\delta Q_3)`$ $`=`$ $`\kappa _3{\displaystyle \underset{k}{}}\left[\left(a_k^{}+a_k\right)\left(A_1(k)+A_1^{}(k)\right)+\left(b_k^{}+b_k\right)\left(A_2(k)+A_2^{}(k)\right)\right]`$ (22) $`_2(\delta Q_2)`$ $`=`$ $`\kappa _2{\displaystyle \underset{k}{}}\left[\left(B_1(k)+B_1^{}(k)\right)\rho _a(k)+\left(B_2(k)+B_2^{}(k)\right)\rho _b(k)\right]`$ (23) We are now explicitly using the $`(\pi \pi 0)`$ wavevector for Jahn-Teller oxygen distortions. New operators $`A`$, $`B`$, are introduced to represent the Fourier transformed $`\delta Q_3(k)`$ and $`\delta Q_2(k)`$ variables, and new operators $`\rho _a`$ and $`\rho _b`$ are used for orbiton densities on A and B sites: $`A_1(k)`$ $`=`$ $`2f_{k_z}\left(1e^{ik_z}\right)f_{k_x}f_{k_y}+g_{k_x}e^{ik_x}+g_{k_y}e^{ik_y}`$ (24) $`A_2(k)`$ $`=`$ $`2g_{k_z}\left(1e^{ik_z}\right)g_{k_x}g_{k_y}+f_{k_x}e^{ik_x}+f_{k_y}e^{ik_y}`$ (25) $`B_1(k)`$ $`=`$ $`f_{k_x}f_{k_y}g_{k_x}e^{ik_x}+g_{k_y}e^{ik_y}`$ (26) $`B_2(k)`$ $`=`$ $`g_{k_x}g_{k_y}f_{k_x}e^{ik_x}+f_{k_y}e^{ik_y}`$ (27) $`\rho _a(k)`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{N}}}{\displaystyle \underset{q}{}}a_{k+q}^{}a_q`$ (28) $`\rho _b(k)`$ $`=`$ $`\sqrt{{\displaystyle \frac{2}{N}}}{\displaystyle \underset{q}{}}b_{k+q}^{}b_q`$ (29) The term due to $`\delta Q_3`$ couples orbitons to the phonon subspace with coupling constant $`\kappa _3=g\left(\mathrm{}/6M\omega \right)^{1/2}=`$ $`\left(\mathrm{}\omega \mathrm{\Delta }/48\right)^{1/2}`$. Since the phonon energy is much smaller than the orbiton energy $`(\mathrm{}\omega \mathrm{\Delta })`$ this term can be treated perturbatively. The term due to $`Q_2`$ couples phonon vibrations to orbiton densities with the coupling constant $`\kappa _2=2g(\mathrm{}/2M\omega )^{1/2}=(\mathrm{}\omega \mathrm{\Delta }/4)^{1/2}`$. This term describes oxygen vibrations around new positions which relax by an amount proportional to the orbiton density $`\rho `$ in order to lower the energy of the orbiton excitation. To deal with term we make a displaced oscillator transformation. In place of the phonon operators $`f_{k_\lambda },g_{k_\lambda }`$ defined above, we will use the following operators: $`F_{k_x}`$ $`=`$ $`f_{k_x}+{\displaystyle \frac{\kappa _2}{\mathrm{}\omega }}\left(\rho _a(k)+\rho _b(k)e^{ik_x}\right)`$ (30) $`F_{k_y}`$ $`=`$ $`f_{k_y}{\displaystyle \frac{\kappa _2}{\mathrm{}\omega }}\left(\rho _a(k)+\rho _b(k)e^{ik_y}\right)`$ (31) $`G_{k_x}`$ $`=`$ $`g_{k_x}{\displaystyle \frac{\kappa _2}{\mathrm{}\omega }}\left(\rho _a(k)e^{ik_x}+\rho _b(k)\right)`$ (32) $`G_{k_y}`$ $`=`$ $`g_{k_y}+{\displaystyle \frac{\kappa _2}{\mathrm{}\omega }}\left(\rho _a(k)e^{ik_y}+\rho _b(k)\right)`$ (33) $`F_{k_z}`$ $`=`$ $`f_{k_z}`$ (34) $`G_{k_z}`$ $`=`$ $`g_{k_z}`$ (35) These new operators are bosonic. They were chosen to complete squares in the $`_{\mathrm{ph}}^0+_2(\delta Q_2)`$ part of the Hamiltonian, which becomes $`_{\mathrm{ph}}^0`$ $`+`$ $`_2(\delta Q_2)=\mathrm{}\omega {\displaystyle \underset{k,\lambda =x,y,z}{}}\left(F_{k_\lambda }^{}F_{k_\lambda }+G_{k_\lambda }^{}G_{k_\lambda }\right)`$ (36) $``$ $`{\displaystyle \frac{(\kappa _2)^2}{\mathrm{}\omega }}{\displaystyle \underset{k,\lambda =x,y}{}}\left[\left(\rho _a(k)+\rho _b(k)e^{ik_\lambda }\right)\times \left(\mathrm{HC}\right)+\left(\rho _a(k)e^{ik_\lambda }+\rho _b(k)\right)\times \left(\mathrm{HC}\right)\right]`$ (37) where HC stands for Hermitean conjugate. We now truncate the Hilbert space to the subspace with only one orbiton present. Then the product of two adjacent orbiton annihilation operators gives zero. Then Eqn. 37 can be simplified further by noting that $`_k[\rho _a(k)\rho _a(k)+\rho _b(k)\rho _b(k)]`$ is the same as $`_k[a^{}(k)a(k)+b^{}(k)b(k)]`$; cross terms are zero in our subspace. The second term of Eqn. 37 then combines with $`_{\mathrm{orb}}^0`$ Eqn. (21) to renormalize the orbiton energy by $`4(\kappa _2)^2/\mathrm{}\omega =\mathrm{\Delta }`$. The orbiton energy is now exactly half of the bare energy $`2\mathrm{\Delta }`$. The reason is obvious in a site representation. Flipping an orbital creates a local region where an orbital is surrounded in the $`xy`$ plane by other orbitals of the same orientation. Oxygen atoms adjacent to the central (flipped) orbital will now prefer to locate in undistorted positions. This means that rather than raising the local energy from $`\mathrm{\Delta }`$ to $`+\mathrm{\Delta }`$ with the flip, we have instead raised it from $`\mathrm{\Delta }`$ to 0. This is the sole influence of $`\delta Q_2`$-type fluctuations in the one-orbiton subspace. Now we transform the rest of the Hamiltonian to the displaced oscillator operators. Substituting new phonon operators (35) into $`_3(\delta Q_3)`$ generates the same term as in (23), but with old phonon operators replaced with new ones, plus additional terms coupling orbiton operators to orbiton density operators. The new Hamiltonian has the following form: $$=_{\mathrm{orb}}+_{\mathrm{ph}}+_1+_2+_3$$ (38) where the first two terms are the new bare orbitons and phonons $`_{\mathrm{orb}}`$ $`=`$ $`\mathrm{\Delta }{\displaystyle \underset{k}{}}\left(a_k^{}a_k+b_k^{}b_k\right)`$ (39) $`_{\mathrm{ph}}`$ $`=`$ $`\mathrm{}\omega {\displaystyle \underset{k,\lambda =x,y,z}{}}\left(F_{k_\lambda }^{}F_{k_\lambda }+G_{k_\lambda }^{}G_{k_\lambda }\right).`$ (40) There are now three coupling terms left, $`_1`$ $`=`$ $`\kappa _3{\displaystyle \underset{k}{}}\left(a_k^{}A_1(k)+b_k^{}A_2(k)+\mathrm{HC}\right)`$ (41) $`_2`$ $`=`$ $`\kappa _3{\displaystyle \underset{k}{}}\left(a_k^{}A_1^{}(k)+b_k^{}A_2^{}(k)+\mathrm{HC}\right)`$ (42) $`_3`$ $`=`$ $`\kappa _2^0{\displaystyle \underset{k}{}}\left(\mathrm{cos}k_x\mathrm{cos}k_y\right)\left(a_k^{}\rho _b(k)+b_k^{}\rho _a(k)+\mathrm{HC}\right).`$ (43) Here the $`A`$ and $`B`$ operators are the same as defined in (29) after substition of the new phonon operators $`F`$ and $`G`$ in place of $`f`$ and $`g`$. The new coupling constant $`\kappa _2^0`$ is $`4\kappa _1\kappa _2/\mathrm{}\omega =\mathrm{\Delta }/\sqrt{12}`$. The perturbation part of the Hamiltonian $`_1+_2+_3`$ couples orbitons and phonons to states with energies lying above or below by the order of $`\mathrm{\Delta }`$. These terms will be handled in the next section. ## IV Low energy Effective Hamiltonian In the Hamiltonian (40,43) orbitons and phonons are coupled directly through the $`_1`$ term. The term $`_3`$ does not couple to the phonons, but couples the one orbiton subspace to the two orbiton subspace. The remaining $`_2`$ piece couples the one orbiton and one phonon states to states with one more orbiton and one more phonon. Thus the Hamiltonian matrix acting on the one phonon and one orbiton subspace has the structure: $$\mathrm{\Psi }=\left(\begin{array}{ccccc}\mathrm{}\omega \widehat{I}& _1& _{2a}& 0& 0\\ _1& \mathrm{\Delta }\widehat{I}& 0& _3& _{2b}\\ _{2a}& 0& \left(\mathrm{\Delta }+2\mathrm{}\omega \right)\widehat{I}& 0& 0\\ 0& _3& 0& 2\mathrm{\Delta }\widehat{I}& 0\\ 0& _{2b}& 0& 0& \left(2\mathrm{\Delta }+\mathrm{}\omega \right)\widehat{I}\end{array}\right)\left(\begin{array}{c}\mathrm{\Psi }(1\mathrm{p}\mathrm{h})\hfill \\ \mathrm{\Psi }(1\mathrm{o}\mathrm{r}\mathrm{b})\hfill \\ \mathrm{\Psi }(1\mathrm{o}\mathrm{r}\mathrm{b}+2\mathrm{p}\mathrm{h})\hfill \\ \mathrm{\Psi }(2\mathrm{o}\mathrm{r}\mathrm{b})\hfill \\ \mathrm{\Psi }(2\mathrm{o}\mathrm{r}\mathrm{b}+1ph)\hfill \end{array}\right)$$ (44) The operator $`_2`$ has matrix elements $`_{2a}`$ when it couples the 1 phonon to the 1 orbiton plus 2 phonon subspace, and matrix elements $`_{2b}`$ when it couples the 1 orbiton to the 2 orbiton plus 1 phonon subspace. The operator $`_1`$ couples 1 orbiton to 1 phonon. It also couples 1 orbiton plus 2 phonons to 2 orbitons plus 1 phonon. However, this block does not influence the 1 phonon or 1 orbiton answers except in higher order, so these elements of the Hamiltonian matrix are omitted. This Hamiltonian can be perturbatively reduced to two separate low-energy effective Hamiltonians, one for single orbitons and one for single phonons: $`_{\mathrm{orb}}^{\mathrm{eff}}`$ $`=`$ $`_{\mathrm{orb}}+_1\left(E\mathrm{}\omega \right)^1_1+_{2\mathrm{b}}\left(E\left(2\mathrm{\Delta }+\mathrm{}\omega \right)\right)^1_{2\mathrm{b}}+_3\left(E2\mathrm{\Delta }\right)^1_3`$ (45) $`_{\mathrm{ph}}^{\mathrm{eff}}`$ $`=`$ $`_{\mathrm{ph}}+_1\left(E\mathrm{\Delta }\right)^1_1+_{2\mathrm{a}}\left(E\left(\mathrm{\Delta }+2\mathrm{}\omega \right)\right)^1_{2\mathrm{a}}`$ (46) The $`2\times 2`$ effective Hamiltonian $`_{\mathrm{orb}}^{\mathrm{eff}}`$ acts on wavefunctions built from two types of orbitons $`(a_q^{},b_q^{})0,0>`$. Similarly the $`6\times 6`$ effective Hamiltonian $`_{\mathrm{ph}}^{\mathrm{eff}}`$ acts on six-component wavefunctions $`(F_{q_\lambda }^{},G_{q_\lambda }^{})0,0>`$,where $`\lambda =x,y,z`$. The unperturbed ground state $`0,0>`$ with zero orbitons and phonons couples through the $`_2`$ term to states of one phonon and one orbiton. This leads to a ground energy shift $`\delta E_{GS}=<0,0_2\left(E\left(\mathrm{\Delta }+\mathrm{}\omega \right)\right)^1_20,0>`$. The problem is now reduced to computation and diagonalization of a $`2\times 2`$ and a $`6\times 6`$ matrix. Note that we could have taken into account the $`_1`$ part of the perturbation Hamiltonian by leaving it as an off-diagonal coupling term, giving an $`8\times 8`$ effective Hamiltonian to compute and diagonalize. The error we make by folding it down as in (46) has the same order of magnitude $`(\mathrm{}\omega /\mathrm{\Delta })^2`$ as the other errors we make in our perturbation theory. Our next task is to evaluate the matrix elements of the effective Hamiltonians (46). After straight-forward algebra the result (with the ground energy shift $`\delta E_{GS}=12\kappa _3^2N/\left(\mathrm{\Delta }+\mathrm{}\omega \right)`$ subtracted out) is: $`_{\mathrm{orb}}^{\mathrm{eff}}`$ $`=`$ $`\left(\mathrm{\Delta }{\displaystyle \frac{(\kappa _2^0)^2}{\mathrm{\Delta }}}\right)\widehat{I}+\left({\displaystyle \frac{\kappa _3^2}{\mathrm{\Delta }\mathrm{}\omega }}+{\displaystyle \frac{\kappa _3^2}{\mathrm{\Delta }+\mathrm{}\omega }}\right)\left(\left(128\mathrm{cos}k_z\right)\widehat{I}2\left(\mathrm{cos}k_x+\mathrm{cos}k_y\right)\widehat{\sigma }_x\right)`$ (47) $`_{\mathrm{ph}}^{\mathrm{eff}}`$ $`=`$ $`\mathrm{}\omega \widehat{I}+\left({\displaystyle \frac{\kappa _3^2}{\mathrm{\Delta }\mathrm{}\omega }}+{\displaystyle \frac{\kappa _3^2}{\mathrm{\Delta }+\mathrm{}\omega }}\right)\left(\begin{array}{cc}_a& _b\\ _b^{}& _a\end{array}\right)`$ (50) Where $`\widehat{\sigma }_x`$ is a Pauli matrix. The $`3\times 3`$ matrixes $`_a`$ and $`_b`$ are $`_a`$ $`=`$ $`\left(\begin{array}{ccc}2& 1+e^{i\left(k_xk_y\right)}& 2+2e^{ik_z}\\ 1+e^{i\left(k_yk_x\right)}& 2& 2+2e^{ik_z}\\ 2+2e^{ik_z}& 2+2e^{ik_z}& 88\mathrm{cos}k_z\end{array}\right)`$ (54) $`_b`$ $`=`$ $`\left(\begin{array}{ccc}2\mathrm{cos}k_x& e^{ik_y}e^{ik_x}& 2e^{ik_x}2e^{i\left(k_xk_z\right)}\\ e^{ik_y}e^{ik_x}& 2\mathrm{cos}k_y& 2e^{ik_y}2e^{i\left(k_yk_z\right)}\\ 2e^{ik_x}2e^{i\left(k_zk_x\right)}& 2e^{ik_y}2e^{i\left(k_zk_y\right)}& 0\end{array}\right)`$ (58) By solving (50) for eigenvalues and eigenvectors we find the dispersion law of orbitons and phonons. In our approximation it is consistent to neglect $`\mathrm{}\omega `$ relative to $`\mathrm{\Delta }`$ in the denominators of (50). Then we obtain two dispersive modes for orbitons and two for phonons, while 4 phonon modes remains degenerate with eigenvalues $`E_{ph}^{3,4,5,6}=\mathrm{}\omega `$. The dispersive modes are: $`E_{\mathrm{orb}}^1`$ $`=`$ $`{\displaystyle \frac{11}{12}}\mathrm{\Delta }+{\displaystyle \frac{\mathrm{}\omega }{12}}\left(64\mathrm{cos}k_z+\mathrm{cos}k_x+\mathrm{cos}k_y\right)`$ (59) $`E_{\mathrm{orb}}^2`$ $`=`$ $`{\displaystyle \frac{11}{12}}\mathrm{\Delta }+{\displaystyle \frac{\mathrm{}\omega }{12}}\left(64\mathrm{cos}k_z\mathrm{cos}k_x\mathrm{cos}k_y\right)`$ (60) $`E_{\mathrm{ph}}^1`$ $`=`$ $`\mathrm{}\omega {\displaystyle \frac{\mathrm{}\omega }{12}}\left(64\mathrm{cos}k_z+\mathrm{cos}k_x+\mathrm{cos}k_y\right)`$ (61) $`E_{\mathrm{ph}}^2`$ $`=`$ $`\mathrm{}\omega {\displaystyle \frac{\mathrm{}\omega }{12}}\left(64\mathrm{cos}k_z\mathrm{cos}k_x\mathrm{cos}k_y\right)`$ (62) The Jahn-Teller distortion $`Q_2`$ with a $`(\pi ,\pi ,0)`$ wavevector creates a simple tetragonal Brillouin zone, show in Fig. 2. Fig. 3 shows the orbiton and phonon eigenvalues along principle symmetry lines. An interesting result is that the energy of one of the phonon branches goes to zero at wavevector $`\stackrel{}{k}=(0,0,\pi )`$. The ground state energy of our Hamiltonian is degenerate in $`(Q_2,Q_3)`$ plane. It depends only on magnitude of the vector $`\stackrel{}{Q}=(Q_2,Q_3)`$, but not on its direction. We have chosen a $`Q_2`$-type distortion as is observed in experiment. In our model, an infinitesimal distortion of the $`Q_3`$ type does not cost any energy. This Goldstone mode occurs necessarily at the $`Z`$ point of the Brillouin zone because only a $`\stackrel{}{q}=(\pi ,\pi ,\pi )`$ order parameter is independent of the direction of ordering in $`(Q_2,Q_3)`$-space. ## V conclusion In this work we discuss the dispersion of the orbiton excitation caused by coupling to phonons. We solve a simple but reasonably realistic model, which has previously been used to describe the insulating nature and magnetic ordering of lightly doped LaMnO<sub>3</sub>. The solution makes a single occupancy approximation, allowing only one electron per Mn site. We restrict consideration to creation of only a single orbiton defect. We show that the electron-phonon interaction reduces the energy of the excitation by a factor of two, which is due to lattice relaxation. We also obtain a zero energy excitation or Goldstone mode which is due to a broken continuous symmetry of the Hamiltonian. We ignored two other sources of dispersion of the orbiton – the hopping terms ($`t^2/U`$), and direct quadrupolar Coulomb coupling. These effects would probably cause a similar size dispersion to the one we considered, but would not have the major effect of reducing the JT gap by a factor of two which occurs because of coupling to phonons. ## ACKNOWLEDGMENTS We thank A. J. Millis, J. P. Hill and S. Ishihara for help. This work was supported in part by NSF grant no. DMR-9725037.
no-problem/9906/cond-mat9906090.html
ar5iv
text
# Electrophoresis simulated with the cage model for reptation ## I Introduction A widely used tool to separate mixtures of DNA molecules by length is DC electrophoresis. The DNA is confined to an agarose gel, and an electric field is applied. Since DNA is negatively charged, it moves towards the positive electrode as a result of this electric field. As the drift velocity depends on the length, DNA fragments with different lengths end up in different bands, and can therefore easily be separated. Since DNA fragments are usually much longer than the typical spacing between the gel strands, they are unable to move sideways. De Gennes described the motion of a polymer in such an environment, and termed it reptation: the polymers move by diffusion of ‘defects’ along the chain of monomers. Each defect contracts the polymer by a certain amount of length, called its stored length. When a defect passes a monomer, the monomer is moved by this distance. Figure 1 shows an example where a defect travels past monomer B. Two models are widely used to simulate reptation: the repton model, introduced by Rubinstein , and the cage model, introduced by Evans and Edwards . In both models, monomers reside on sites of a simple cubic lattice (or in two dimensions a square lattice), and are connected by bonds; the dynamics consist of single-monomer moves. In the repton model, stored length consists of zero-length bonds. For this model, it was already proposed by Rubinstein , and later proven by Prähofer and Spohn that the diffusion constant $`D`$ of the polymer in the limit of long polymer length $`L`$ obeys the scaling $`L^2D=1/3`$. For finite lengths, the diffusion constant is known numerically exact up to length 20 and from Monte Carlo simulations up to length 250 . The repton model has been adapted for the study of DC electrophoresis by Duke . This Rubinstein-Duke model has been studied numerically for lengths up to 400 . Simulations of this model are easy because it can, without loss of generality, be reduced to a one-dimensional model. In the cage model, stored length consists of a pair of anti-parallel nearest-neighbour bonds, called ‘kinks’. The polymer diffusion constant in this model has been determined numerically exact for small $`L`$ , and with Monte Carlo simulations for polymers up to length 200 . As in the repton model, the polymer diffusion constant scales as $`DL^2`$. In this paper, we extend the cage model to simulate a charged polymer in an electric field. Interestingly, we find differences in the scaling of the polymer drift velocity as compared to the Duke-Rubinstein model. In section II we describe the cage model and present how the model can be extended to simulate reptation in a non-zero electric field. We discuss in sections II C to II E how efficient simulations can be achieved with multispin coding; these sections are not needed for understanding other parts of the paper. In section III we discuss the simulation approach in detail. In section IV we discuss scaling arguments for the drift velocity. The results are presented in section V which includes statements about the polymer shapes and comparisons to previous reports. ## II Cage model The cage model describes a polymer of $`L`$ monomers, located on the sites of an infinite cubic lattice. The monomers are connected by $`L1`$ bonds with a length of one lattice spacing. A single step of the Monte Carlo simulation consists of selecting randomly a monomer and, if it is free to move, moving it to a randomly selected location (possibly the current location). The monomers at both ends of the polymer are always free to move, but monomers in the interior of the polymer are only free to move when the two neighbours along the chain are located on the same adjacent lattice site. Other movements might result in an acceptable polymer configuration, but are ruled out because they would allow the polymer to move sideways, which is not reptation. One possible interior move is shown in figure 2. Every possible move occurs statistically with unit rate, setting the time scale. A single elementary move thus corresponds to a time increment of $`\mathrm{\Delta }t=1/(2dL)`$, where $`d`$ is the dimensionality of the lattice; in our case, $`d=3`$. ### A Electric field In solution, DNA becomes negatively charged with a fixed charge per unit length. We incorporate this into the cage model by assigning a negative charge $`q`$ per monomer. The polymer is located in a homogeneous electric field $`\stackrel{}{E}`$, that acts on these charged monomers. For two monomer positions $`\stackrel{}{r}_1`$ and $`\stackrel{}{r}_2`$, separated by a displacement $`\stackrel{}{r}_{12}=\stackrel{}{r}_2\stackrel{}{r}_1`$, the difference in potential energy is given by $`U=q\stackrel{}{E}\stackrel{}{r}_{12}`$. The ratio of the corresponding Boltzmann probabilities is $$P_1/P_2=e^{U/(k_BT)}=e^{qEr/(k_BT)}\text{,}$$ (1) where $`E=|\stackrel{}{E}|`$ is the electric field strength, and $`r=\stackrel{}{E}\stackrel{}{r}_{12}/|\stackrel{}{E}|`$ is the displacement parallel to the field. In a Monte Carlo simulation, this ratio determines at which rates the monomers are to be moved along the field or against it. We choose the direction of the electric field along one of the body diagonals of the unit cubes, because then the $`x`$, $`y`$ and $`z`$ directions are equivalent, and within one elementary move, the displacement $`r`$ takes only the two values $`\pm 2/\sqrt{3}`$ times the lattice spacing. For convenience, the units are chosen in such a way that $`qr/k_BT=\pm 1`$. Each monomer moves with a rate $`R^+=\mathrm{exp}(E)`$ for moves which lower the energy and $`R^{}=\mathrm{exp}(E)`$ for moves which raise the energy. When a monomer is selected and is able to move to lower (higher) energy, it will do so with a probability $`P^+`$ ($`P^{}`$), given by $$P^+=\frac{1}{d}\frac{e^E}{e^E+e^E}\text{,}P^{}=\frac{1}{d}\frac{e^E}{e^E+e^E}\text{.}$$ (2) The time increment corresponding to one elementary Monte Carlo move is thus equal to $$\mathrm{\Delta }t=\frac{1}{dL}\frac{1}{e^E+e^E}\text{.}$$ (3) ### B Bond Representation As described in section II the monomers are connected by bonds, where each bond has one of $`2d`$ possible orientations. One way of describing the polymer configuration is by specifying the location of the first monomer and the orientation of all bonds. The advantage of this notation is that only the position of one monomer has to be stored plus the orientations of all bonds. The polymer in figure 2, for example, is described by the position of the first monomer, on the left side of the figure, and `+x-y+z+x-x+z`. The dynamics can be described in terms of bonds. The bonds that are located on both ends of the polymer are always free to move. The internal bonds are free to move only when they are part of a pair of oppositely oriented neighbouring bonds (a kink). The first and last monomer in figure 2 can change to any new bond: `+x`, `+y`, `+z`, `-x`, `-y` or `-z`. The kink configuration `+x-x` can change in any new kink: `+x-x`, `+y-y`, `+z-z`, `-x+x`, `-y+y` or `-z+z`. ### C Multispin Coding With multispin coding, many polymers can be simulated in parallel. We used an approach similar to the one by Barkema and Krenzlin . The simulation that we performed used 64-bit unsigned integers to simulate 64 different polymers in parallel. As described in section II B, there are six directions a bond can point to, so each bond can be encoded using three binary digits. It is now possible to encode 64 bonds in three integers $`x`$, $`y`$ and $`z`$, as shown in table I. In each iteration of the inner loop of the algorithm a random monomer $`i`$, $`0i<L`$, is selected. When an inner monomer is selected the two surrounding bonds are compared; if they are opposites, they are replaced by a randomly generated pair of opposite bonds. Section II E describes how to generate those bonds. The first and last monomers are special cases, which are described in section II D. To find the kinks in all of the 64 polymers, we use Equation (4): $$\begin{array}{ccc}\hfill k_i& =& (x_{i1}x_i)\hfill \\ & & (y_{i1}y_i)\hfill \\ & & (z_{i1}z_i)\text{.}\hfill \end{array}$$ (4) Monomer $`i`$ is surrounded by bonds $`i1`$ and $`i`$. Bit $`j`$ of $`k_i`$ is $`1`$ if the surrounding bonds of monomer $`i`$ of polymer $`j`$ are in opposite directions. If a monomer can be moved, it will be relocated using a list of random kinks encoded in $`\widehat{x}`$, $`\widehat{y}`$ and $`\widehat{z}`$. Bonds $`i1`$ and $`i`$ that surround monomer $`i`$ are replaced by $`\widehat{x}`$, $`\widehat{y}`$ and $`\widehat{z}`$ and their binary complements, respectively. Equation (5) shows how this can be done: $$\begin{array}{ccc}\hfill x_{i1}& =& (\neg k_ix_{i1})(k_i\widehat{x})\hfill \\ \hfill y_{i1}& =& (\neg k_iy_{i1})(k_i\widehat{y})\hfill \\ \hfill z_{i1}& =& (\neg k_iz_{i1})(k_i\widehat{z})\hfill \\ \hfill x_i& =& (\neg k_ix_i)(k_i\neg \widehat{x})\hfill \\ \hfill y_i& =& (\neg k_iy_i)(k_i\neg \widehat{y})\hfill \\ \hfill z_i& =& (\neg k_iz_i)(k_i\neg \widehat{z})\text{.}\hfill \end{array}$$ (5) These 27 logical operations replace the kinks near monomer $`i`$ in all 64 polymers; polymers that have no kink near monomer $`i`$ are left unaltered. ### D First and last monomer The first and last monomers are always free to move. When one of those monomers is selected we can just replace the bonds with randomly generated bonds. To keep track of the position of the first monomer we have to calculate the distances traveled in the $`x`$, $`y`$ and $`z`$ directions. Since those directions are symmetric we only calculate $`r=x+y+z`$. For this we only need to know whether the first bonds point at a negative direction, which is one of `-x`, `-y` and `-z`. This is done using the following equation: $$d=(x_0y_0)(y_0z_0)(z_0x_0)\text{.}$$ (6) Now the new random bonds are inserted: $$\begin{array}{c}x_0=\neg \widehat{x}\\ y_0=\neg \widehat{y}\\ z_0=\neg \widehat{z}\text{.}\end{array}$$ (7) Monomer $`0`$ only has one bond, which is number $`0`$. Section II C tells us that we have to use the binary complement of the random kink. We now calculate once more whether the first bond points at a negative or positive direction, and with this information we can calculate the new positions of the first monomers: $$r_i=r_i2d_{\text{before}}^{(i)}+2d_{\text{after}}^{(i)}\text{.}$$ (8) When the last monomer is selected, it can simply be replaced with random new bonds: $$\begin{array}{c}x_{L2}=\widehat{x}\\ y_{L2}=\widehat{y}\\ z_{L2}=\widehat{z}\text{.}\end{array}$$ (9) ### E Generation of random kinks The algorithm described above relies on the availability of random kinks. These kinks should be generated with the probabilities as given in equation (2). Since the two bonds in a kink have opposite directions only one bond has to be generated; the bond on the other side of the monomer is easily derived. The properties of detailed balance are used to create those bonds correctly. Certain properties must be enforced: first of all the `x`, `y` and `z` directions should occur with the same probability; secondly the ratio of the probabilities for `+` and `-` bonds is given by quotient of $`P^{}`$ and $`P^+`$, as given in equation (2); this quotient is given by $$P^{\text{rel}}=P^+/P^{}=e^{2E}\text{.}$$ (10) The first property is enforced by rotating some of the bonds (we used 50%) the following way: `x`$``$`y`, `y`$``$`z` and `z`$``$`x`. Using the randomly generated bit pattern $`r`$ the following statements are used to rotate the bonds: $$\begin{array}{c}\stackrel{~}{x}=(r\widehat{x})(\neg r\widehat{y})\hfill \\ \stackrel{~}{y}=(r\widehat{y})(\neg r\widehat{z})\hfill \\ \stackrel{~}{z}=(r\widehat{z})(\neg r\widehat{x})\text{.}\hfill \end{array}$$ (11) The second property is then enforced by inverting some of the bonds. With 50% probability, the negative bonds are inverted, with $`P^{\text{rel}}`$ times 50%, also the positive bonds are inverted. To make sure that all random kinks are independent we create a list of those and reshuffle this list regularly. ## III Simulations The simulation algorithm described in sections II C to II E was implemented using the C programming language. The random number generator we used is a lagged (24, 55) additive Fibonacci generator. The simulations are done on a Silicon Graphics Origin 200 (180 Mhz) and on a DEC Alpha (466 Mhz) computer. The latter is faster and takes about 1.1 $`\mu `$s per Monte Carlo step. The polymers where initialized in a U-shape with both ends in the direction of the electric field. At regular intervals we checked whether a polymer has moved at least its own size, which is the maximum distance between any two monomers. When this has occurred for a polymer, we assume that the polymer has thermalized, the real measurement starts when this thermalization has finished. The measurement is stopped when all polymers have thermalized and the average distance traveled by all polymers is a few times their own size. We assume that measurements are statistically independent when a polymer has traveled a distance equal to its own size. We have performed simulations for lengths $`3`$ up to $`200`$. The time taken to calculate the drift velocity varied from a few seconds for small polymers up to about 17 hours for the longest polymers ($`L=200`$) in the smallest electric field ($`E=0.001`$). Simulations of longer polymers take too much time to compute the drift velocity for small electric fields. ## IV Scaling arguments for the drift velocity The velocity of a polymer in a small electric field behaves according to the Nernst-Einstein relation, $`v=FD`$, where $`F=qLE`$ is the force. The diffusion constant can thus be calculated from the drift velocity by $`D=v/(qLE)`$ in the limit $`E0`$. De Gennes found the diffusion constant to be proportional to $`DL^2`$. This means the drift velocity is: $`vqE/L`$. For slightly larger electric fields the Nernst-Einstein relation breaks down. Barkema, Marko and Widom give an intuitive explanation of the dependence of the drift velocity on the electric field strength. The argument goes as follows. A random polymer will have an end-to-end length around $`h=\sqrt{L}`$. When an electric field is applied, the polymer is stretched in the direction of the electric field. When the electric field exceeds a certain level, the polymer as a whole does no longer resemble a random walk: $`h>\sqrt{L}`$. One may cut the polymer into $`n_b`$ pieces (‘blobs’) of length $`L_b=L/n_b`$, that each still look like a random walk; the average end-to-end distance of the blobs is equal to $`h_b=\sqrt{L_b}`$. Two forces work on the blob. The elastic force tries to contract a stretched polymer and is proportional to the size of the blob and inversely proportional to the length of the part of the polymer that forms the blob: $`F_{\text{elastic}}h_b/L_b`$. The electric force tries to stretch the polymer and is proportional to the size of the blob as well as the electric field: $`F_{\text{electric}}h_bE`$. These two forces have to be in balance which implies that the blob size is $`L_bE^1`$. The Nernst-Einstein relation now applies to the blobs, so $`v=F_bD_b=qL_bED_b`$. Again, if the blob size is large enough, $`D_bL_b^2`$ which makes the speed of the polymer quadratic in the electric field: $`vE/L_bE^2`$. This effect has already been observed in the Rubinstein-Duke model by Barkema, Marko and Widom. ## V Results The results of our simulations are presented in figure 3. The short polymers, up to length $`20`$, show a behavior different from the longer polymers. These short polymers have no superlinear dependence on the velocity on the electric field. When a small force, $`EL1`$, is applied to the polymers, the velocity of the polymers varies linearly with the electric field. When a force around $`EL1`$ is applied to the longer polymers, the polymer velocity depends superlinearly on the electric field. We show in figure 8 that the dependence becomes quadratic for long polymer chains, as derived in section III. For much larger electric fields, the velocity decreases to zero. For polymer length $`L=100`$ we performed some short simulations to get insight in the typical movement of the center of mass of the polymer. In figure 4 the position of the center of mass, scaled with a factor of $`1/E`$, is plotted as a function of time, for different field strengths. The starting point of the polymers are chosen in a way that the graphs do not overlap. For the smallest electric fields the movement is just like one would expect from a diffusing particle, it moves randomly, but with some preferential direction. For the electric field in the middle range, the diffusion effect becomes relatively smaller. This results in a smoother behaviour. In high electric fields the movement of the center of mass sometimes halts, when the force on the ends of the polymer pulls the polymer into a U-shape. When this happens the polymer has to untangle itself before its center of mass can move forward again. ### A Polymer shapes The polymer shape in a small electric field resembles a random walk, as shown on the left side of figure 5. When the electric field is increased, the shape becomes stretched parallel to the electric field ; the configuration may be viewed as a set of blobs which move with independent speed, as discussed in section IV. As shorter polymers move more quickly in a given electric field, the blob-configuration moves faster than a random walk configuration which results in a superlinear increase of speed when the electric field is changed. When the electric field is increased above a certain value the shape may transforms into an U-shape, as shown in figure 6. With higher electric fields it becomes more difficult to escape from this U-shape. Since the polymer cannot move sideways it is trapped in the lattice for a long time compared to the time it moves. Figure 6 shows polymers in different configurations. The first polymer is stretched in the direction of the electric field. This configuration may be viewed as a large number of very small blobs. As such, the polymer has a high velocity, which may also be seen in figure 4 near $`5.810^7`$ Monte Carlo steps. The second polymer is a transition configuration between the fast-moving cigar-like configuration as described above and the U-shape configuration. The polymer forms a ‘knot’ which is later passed by the trailing end of the polymer. The third polymer has a typical U-shape. The polymer in this configuration is almost fully stretched. This means that it has only a small number of kinks, which means that there is almost no stored length. Just before the polymer escapes from the U-shape, as is the fourth polymer, its configuration is very much stretched and has almost no stored length. This state transforms quickly in a state that resembles the state of the first polymer in figure 6. As described before, the number of kinks in a polymer decreases when the polymer is stretched. To check the dependence on the electric field we have performed some short simulations to find the average number of kinks on each location along the polymer. The simulations consisted of $`10^9`$ Monte Carlo steps after $`210^8`$ steps of thermalization, starting with a random configuration. Every $`10^6`$ Monte Carlo steps the kinks are counted. The fraction of time that a kink exists on a certain location is displayed in figure 7. For small electric fields the polymer configuration is known to resemble a random walk in three dimensions. The average number of kinks is thus expected to be $`1/6`$. For higher electric fields the U-shape configuration becomes more frequent. In this configuration the kinks are likely to diffuse towards the ends of the polymer, which means that the average number of kinks in the middle of the polymer decreases. When this happens we can no longer apply the blob argument as described in section IV. The mobility of the blobs in the middle of the polymer decreases as the average number of kinks in that region decreases. For a longer U-shaped polymer it becomes more difficult to escape from this configuration. The probability of a kink moving from one end of the polymer to the other end decreases with the length of the polymer. This means that longer polymers spend a longer time in the U-shape configurations. When the density of kinks becomes less than $`1/6`$ per monomer, the entropic force that contracts the polymer is no longer in balance with the electric force. The polymer itself now transports the force along the chain, which may be better explained by the continuous model of Deutsch and Madden . ### B Comparisons to previous reports The results of the Duke-Rubinstein model have been compared to actual experiments. For longer polymers, the data is well described by $$\frac{L^2v}{\alpha }=\left[\left(\frac{LE}{\beta }\right)^2+\left(\frac{LE}{\beta }\right)^4\right]^{1/2}$$ (12) This function is equivalent to the function $`v^2=aE^2+bE^4`$, where $`a`$ and $`b`$ are functions of $`\alpha `$, $`\beta `$ and $`L`$. To check whether our results show the same scaling behaviour, we collapsed our data to the function $`v^{}=\sqrt{E_{}^{}{}_{}{}^{2}+E_{}^{}{}_{}{}^{4}}`$ in figure 8, where $`v^{}=(\sqrt{b}/a)v`$ and $`E^{}=(\sqrt{b/a})E`$. The data in the third region is discarded in calculating $`a`$ and $`b`$. Experiments have been performed on DC electrophoresis , both articles confirm the existence of regions where $`vE`$ and $`vE^2`$. The diffusion constant is equal to $`D=\sqrt{a}/L`$. The scaling found by Barkema and Krenzlin is given by $`DN^2=0.173+1.9N^{2/3}`$, where $`N=L1`$. Figure 9 shows our results compared to their scaling function. Our results agree within statistical errors. ## VI Conclusions The cage model is extended to simulate DC electrophoresis, and the drift velocity of polymers in a gel is measured as a function of polymer length $`L`$ and electric field strength $`E`$. The polymers behave differently in three regimes of the electric field: in a small electric field the velocity depends linearly on the electric field, in a high electric field the polymers are likely to be trapped in an U-shape. The regime in between shows a superlinear dependency on the electric field. We showed that this dependency becomes quadratic in the electric field for $`L\mathrm{}`$. We have shown that the average number of kinks is not equally distributed over the polymer in high electric fields, because the polymers tend to get stretched in the direction of the electric field. Hence the stored length disappears from the polymer. ## Acknowledgement We thank G.T. Barkema and A.G.M. van Hees for useful discussion. The High-performance computing group of Utrecht University is gratefully acknowledged for ample computer time.